Anda di halaman 1dari 31

1 hcAinencan Asscicialioii ol Petroleum Geologisis Bulleun

V 6K, No. (HJunc 1984), P. 7!.V743,27 Figs., 1 Table

Geothermal Gradients, Hydrodynamics, and


Hydrocarbon Occurrences, Alberta, Canada'
BRIAN HITCHON

result of the second Laramide orogenic pulse was, therefore, to reverse the geothermal gradient pattern resulting
Using coal rank data and the demonstrated genetic link mainly from compaction flow, and to impose a topobetween the present hydrodynamic regime and geothermal graphically controlled hydrodynamic regime with flow in
gradient pattern, it is possible to reconstruct the geother- the same direction but with a 9C/km (0.5F/100 ft)
mal history of Alberta over the past 60 m.y. Palinspastic decreased geothermal gradient in western Alberta and a
adjustment for tectonic compression caused by the major 4C/km (0.2F/100 ft) increased geothermal gradient in
Laramide thrusting shows that the predeformation eastern Alberta.
isoreflectance lines increased logarithmically with depth.
As erosion proceeded during the Tertiary, the entire
In the late Paleocene, the geothermal gradient was about Alberta basin warmed up until the present temperature
23C/km (1.25F/100 ft) in the eastern Alberta Plains, regime was attained. Study of some major hydrocarbon
compared to about 30C/km (1.65F/100 ft) in the west- occurrences shows a genetic link between the hydrocarbon
ern Alberta Plains, a regional trend opposite to the post- occurrence and the present fluid regime which itself is conLaramide trends. Reconstruction of the early Eocene trolled by the present topography. For at least the signifisurface indicates western uplands with geothermal gradi- cant Upper Cretaceous shallow gas reserves in
ents as low as 21C/km (1.15F/100 ft) and eastern low- southeastern Alberta, however, their position may already
lands with geothermal gradients of 27C/km (1.5F/100 have been determined as early as the beginning of the
ft). Compared to the present situation, this represents an Eocene by the topographic surface at that time. Thereenhanced topographic surface and a subdued geothermal fore, the extent to which the present topographic surface is
gradient pattern.
responsible for hydrocarbon accumulation through
The genetic relations of topography (water-table eleva- gravity-induced cross-formational flow remains unknown
tion), hydrodynamic regime, and geothermal gradient because this process was operative as long ago as the early
pattern in both the early Eocene and the present conform Eocene. What does not appear to be in doubt is the fact
to a model developed for any compacted sedimentary that gravity-induced cross-formational flow does control
basin with subaerial relief. In this model, on a regional the hydrodynamic regime, which in turn influences the
scale, high topographic areas have high water-table eleva- geothermal gradient pattern and the accumulation of
tions with correspondingly high potentiometric surfaces; hydrocarbons.
these areas control the regional recharge of cold meteoric
water and hence have low geothermal gradients. Areas of
INTRODUCTION
medium elevation exhibit regional lateral flow and intermediate geothermal gradients. Regional topographic lows
A common axiom in geology is that the present is a key to
correspond with low potentiometric surfaces, and the
the past. This axiom is most easily shown for active procregional discharge of warm formation waters from deep in
esses such as sedimentation and volcanism. The demonthe basin results in high geothermal gradients. Local topostration of its validity for relatively passive processes like
graphic features cause perturbations in this picture. Study
heat flow is more difficult. Terrestrial heat flow is the
of both the regional pattern and fine detail of the early
transfer of thermal energy from the interior of the Earth to
Eocene and present topographic surfaces and geothermal
the surface, where it is dissipated. It is expressed through
gradient patterns shows the validity of the model. The
measurement of the geothermal gradient, which is the rate
of increase of temperature with depth. The principles governing the physics of heat flow have been invariate over
Copyright 1984. The American Association of Petroleum Geologists. All
rights reserved.
geologic time, but the geothermal gradient has certainly
^Manuscriptreceived, Decembers, 1982; accepted, June 13,1983.
varied with both time and place.
^Alberta Research Council, Edmonton, Alberta, T6G 2C2, Canada. Contribution No. 1193, Alberta Research Council.
Many reasons have been suggested for regional variaThe author acknowledges, with appreciation and thanl<s, the critical and
tions in geothermal gradients in compacted and undervaluable reviews of J. Tdth (Department of Geology, University of Alberta,
compacted sedimentary basins. Among the most pertinent
Edmonton, Alberta), J. R. Nurl(owski (Alberta Geological Survey, Alberta
Research Council), R. G. McCrossan (Esso Resources Canada Limited, Calare (1) differences in heat-flow rate from the source in the
gary, Alberta), and A . M . Jessop (Division of Seismology and Geothermal Studbasement beneath the basin; (2) differences in vertical and
ies, Department of Energy, Mines and Resources, Ottawa, Ontario). The
lateral thermal conductivity of the rocks owing to a variety
paleogeothermal maps were computer-contoured by A. T Lytviak (Groundwater Department, Alberta Research Council), whose assistance is acknowlof causes such as rock type, cementation, degree of comedged with thanks. AAPG reviewers included Paul H. Jones (Consultant,
paction, and permeability; (3) sources of heat within the
Baton Rouge, Louisiana) and H. W. McGee (Consultant, Lakewood, Colorado),
rocks owing to chemical reactions, mineralogic transforwhose kind comments were much appreciated.
ABSTRACT

713

714

Geothermai Gradients, Hydrodynamics, and Hydrocarbons, Alberta

mations, or radioactivity; (4) volcanism; (5) effects of


Fluids in compacted and undercompacted sedimentary
moving water and other fluids, including juvenile water basins are governed by a well-recognized system of energy
from the basement; (6) tectonic friction causing heating; potential fields (fluid, thermal, electro-osmotic, and
(7) surface effects; (8) anthropogenic anomalies such as chemico-osmotic). Of these, fluid potential is the most
drilling; and, of course, (9) erroneous temperature mea- important and the best known. Fluid potential is defined
surements. Not all are of equal validity and importance, as the amount of work required to transport a unit mass of
and for some sedimentary basins most of the suggested fluid from an arbitrary chosen datum (usually sea level)
reasons can be dismissed because they are either negligible and state to the position and state of the point considered.
or inapplicable.
Electro-osmotic fluid movement is generated by electric
For more than 100 years, individual geologists have potentials resulting from the existence of telluric currents,
observed an association between hydrocarbon occur- and is generally assumed to have a negligible effect on
rences and positive temperature anomalies as reflected in fluid movement. Chemico-osmotic potential is more
higher local geothermai gradients compared to the important, however, in those sedimentary basins with both
regional geothermai gradient. Roberts (1979) has reviewed a wide range in concentration of dissolved salts in formamuch of this literature, and Klemme (1975) has presented tion waters and shales which act as semi-permeable memevidence to show that present and past geothermai gradi- branes. Thermal potential is probably neghgible as a force
ents in various hydrocarbon-bearing basins appear to have to move fluids in most compacted sedimentary basins, but
had an influence on the relative magnitude of hydrocar- it is clearly significant in geothermai regions, including
bon occurrence, especially with respect to many major active sea-floor spreading centers. Whatever the dominant
(giant) hydrocarbon accumulations. Despite this wealth of energy potential that is causing fluid movement, flow is
knowledge, the genetic link between geothermai gradi- always from regions of high energy to regions of lower
ents, hydrodynamics, and hydrocarbon occurrences is not energy. An aqueous continuum occurs throughout all sedimentary basins below the water table, and, with the possialways recognized by explorationists.
In this paper an attempt is made to explain both the ble exception of thick evaporite beds, all sedimentary
regional and fine (local) detail of the geothermai gradient rocks and sediments may be classified as aquifers or
map of Alberta (U.S. Geological Survey and AAPG, aquitardsnone are aquicludes.
1976) in terms of basinal hydrodynamics; a model is develThe classic work of Hubbert (1940) on the theory of
oped which has application to any compacted sedimentary ground-water motion was the first published account of
basin with subaerial relief. Evaluation of the geothermai the basin-wide flow of fluids that considered the problem
gradient pattern and hydrodynamic regime, with respect in exact mathematical terms as a steady-state phenometo major hydrocarbon accumulations, shows a strong non. This work and subsequent studies by Toth (1962,
genetic link between these factors for both the Arabian 1963), Freeze (1966, 1969), and Freeze and Witherspoon
Gulf and Alberta. Using coal rank data, the geothermai (1966, 1967, 1968) showed conclusively that the relief of
regimes during the late Paleocene (before the second Lara- the water table (commonly a subdued replica of the land
mide orogenic pulse) and at the time of maximum burial surface) is the main agent in generating the suite of interde(early Eocene) are reconstructed. Phenomena which con- pendent systems of formation water flow in compacted
trol the present geothermai gradient pattern were appar- sedimentary basins. Computer models and field studies
ently the same that controlled the early Eocene geothermai show these mature, hydraulically continuous environgradient pattern, both differing from the pre-Laramide ments exhibit major meteoric water recharge in major
pattern. For a relatively passive process like heat flow, upland areas, lateral migration under areas of medium eletherefore, the present is indeed a key to the past, and there vation, and major regional discharge in major lowland
has always been a genetic hnk between geothermai gradi- regions. This regional basin-wide recharge-lateral flowents, hydrodynamics, and hydrocarbon occurrences.
discharge regime may be modified by both permeability
differences and near-surface flow controlled by the local
topography (water table). The dominant fluid potential in
HYDRODYNAMICS IN COMPACTED SEDIMENTARY
any part of the basin, therefore, corresponds closely to the
BASINS
fluid potential at the topographic surface in that part of
The flow of fluids through porous media may be consid- the basin, except where modified by large local topoered as a transient (unsteady-state) or steady-state phe- graphic features or channel flow that has been developed
nomenon. Problems concerned with the flow of fluids in relatively highly permeable beds (Hitchon, 1969a, b).
from a borehole or a small group of boreholes through Large rivers commonly exert significant drawndown
confined layers, or if released from storage, are probably effects as exemplified by the South Saskatchewan River in
best solved if they are analyzed as transient phenomena. In Saskatchewan, Canada (Hitchon, 1971, Figure 4). The
studies of small drainage basins or flow regimes across simplest method of determining the regional flow system
large sedimentary basins, it may be assumed that a pseudo is through the construction of hydraulic-head cross secsteady state has been reached. This does not imply static tions and hydraulic-head slice maps (parallel to the sea
conditions, but rather that any recharge to the flow regime level datum). This method may be more instructive in outis negligible in relation to the vast amount of fluid in the lining the true flow regime than constructing hydraulicsystem. In addition, the law of conservation is operative, head maps of individual stratigraphic units because of the
and the recharge is equal to the discharge. In short, we are extensive cross-formational flow present in all sedimentary basins. This method was first used in the western Canconsidering a case of "quasi equilibrium."

Brian Hitchon
ada sedimentary basin where there is abundant data
(Hitchon, 1969a, b), but it is also applicable to less
explored basins.
T6th (1980) advanced the theory that in these geologically mature basins, gravity-induced cross-formational
flow is the principal agent in the transport and accumulation of hydrocarbons. This theory is based on the observation that where flow systems meet or part, quasi-stagnant
zones develop, flow directions change abruptly, and there
is a propensity for hydrocarbon accumulations (Toth,
1970, 1978, 1979). The mechanism becomes operative
after the sediments are compacted, the concomitant primary migration ceases, and subaerial topographic relief
develops. At this stage, hydrocarbons from source or carrier beds are moved along well-defined migration paths
toward discharge foci of converging flow systems, and

Geothdrmal
gradient
^ ^

> 5CC/Km

j - 60akni
| - 3 0 - | 20 - W O K m
|:g;:i:;:| < 20C/km
..- llmM of data
SH - SWEETQRASS HLLS
0
I
0

'

100km
i
60 mi

715

may accumulate en route in hydraulic or hydrodynamic


traps.
Hydrocarbon accumulations are expected and observed
to be preferentially associated with ascending limbs and
quasi-stagnant zones of flow systems. These locations in
the flow systems are characterized by relative potentiometric minima; downward increase in hydraulic heads,
possibly reaching artesian conditions; reduced or zero
lateral hydraulic gradients; and relatively high formation
water salinity. According to Toth's (1980) theory, continuous flow of formation water imports hydrocarbons into
the traps until the trap capacity is reached; then the excess
becomes source material for new accumulations. A temporal change in surface topography causes a proportionate but delayed readj ustment of the flow pattern (transient
state) and, according to Toth (1980), redistribution of the
hydrocarbons; nevertheless, some hydrocarbons may
remain in place, constituting residual deposits in discharge
and quasi-stagnant regions of relict flow systems.
This review has been concerned only with hydrodynamics in compacted sedimentary basins because this is most
relevant to the Alberta part of the western Canada sedimentary basin being considered in this paper. An appraisal
of fluid flow in sedimentary basins in general is also pertinent.
Early in the compaction history of any sedimentary
basin, before subaerial topographic relief has developed,
fluid flow will be upward and updip in the stratigraphic
section because the potential energy of the formation fluids increases downward owing to gravitational settling of
the solid particles. Major unconformities, highly permeable beds, and faults may act as conduits for channel-type
flow. If areally extensive, thick sequences of shales and
siltstones are present, geopressure zones develop. In these
zones, part or all of the weight of the superincumbent rock
column is reflected in the fluid pressure gradients, which
then approach lithostatic gradients. Extensive crossformational flow occurs, and this will be enhanced if a
geopressure zone is present. Once subaerial topographic
relief is developed, the extent to which meteoric water
recharge alters the existing compaction flow will depend
largely on the relief of the water table. The previously
described model will become operative only when compaction has effectively ceased and subaerial topographic
relief has developed.
Provided that thermal maturation conditions are appropriate for the generation of hydrocarbons, the association
of hydrocarbon accumulations with the ascending limbs
and quasi-stagnant zones of flow systems should be
observable in both compacted and undercompacted sedimentary basins. In both basin types, the relative timing of
hydrocarbon generation, trap formation, and development of an appropriate flow system will be crucial to
migration and accumulation of hydrocarbon fluids.
GEOTHERMAL GRADIENT PATTERN, ALBERTA

Figure 1Simplified geotliermal gradient map of Alberta, from


Geotliermal Gradient Map of Nortli America (U.S. Geological
Survey and AAPG, 1976), with soft metric conversion. Only
most eastern, continuous (nonanomalous) 30C/km (1.65F/
100 ft) isogradient contour is shown.

Introduction
The subject of this study is the geothermal gradient pattern of Alberta as shown in the Geothermal Gradient Map

716

Geothermal Gradients, Hydrodynamics, and Hydrocarbons, Alberta

of North America prepared by the American Association


of Petroleum Geologists and the United States Geological
Survey (U.S. Geological Survey and AAPG, 1976). The
original map contains much fine detail, and one way to
simplify and illustrate the regional trends for Alberta is to
plot only the ver> low geothermal gradient areas (<
20C/km, l.TF/MK) ft) and the very high geothermal
gradient regions (> iOC/km, 2.2F/100 ft) (Figure 1).
Before considering the geothermal gradient pattern in
detail, however, it is pertinent to evaluate the applicability
of the 9 effects previously noted, which may cause
regional variations in geothermal gradients.
Geothermal gradient patterns in sedimentary basins are
related to differences in the rate of heat flow through the
basin, which in turn may be due to differences in heat generation in the basement rocks beneath the basin, hence the
concept of heat-flow provinces. Majorowicz and Jessop
(1981) have evaluated the distribution of heat generation
in the Precambrian basement beneath the southern part of
the western Canada sedimentary basin, using published
data on the uranium, thorium, and potassium content of
basement rocks. They found a random distribution, unrelated to the geothermal gradient pattern, with a wide range
of heat generation values ( < 1/tWm"' to close to
14fiWm"'). They concluded that although the heat generation distribution pattern for the Precambrian basement of
the southern part of the western Canada sedimentary
basin does not appear to be exceptional in comparison
with other Precambrian areas, the lack of correlation with
the geothermal gradient pattern rules out a causal relation.
Although Majorowicz and Jessop's (1981) conclusion is
probably correct, it was based on only 112 samples from
the entire area studied. Some tentative trends are discernable from their heat generation distribution map (Majorowicz and Jessop, 1981, Figure 13). For example, in
Alberta, all seven values greater than 8/t Wm"' are in the
region with the greatest concentration of "hot-spots"
(geothermal gradients > 50C/km, 2.75F/100 ft). Further, the greatest concentration of heat generation values
in the range of 3 to 6 /iWm"' is in the area of the Peace
River arch. These tentative trends suggest that the acquisition and contouring of a large data base (e.g., one comparable to the size of the AAPG temperature data file for
Alberta; see U.S. Geological Survey and AAPG, 1976)
might show a closer relation between the heat generation
distribution and the geothermal gradient pattern, especially when the fine detail of both maps is compared.
Assuming that heat generation in the Precambrian basement is unrelated to the geothermal gradient pattern, the
latter might then be due to differences in thermal conductivity of the sedimentary rocks because of differing rock
types, cementation, or degree of compaction or permeability. Majorowicz and Jessop (1981) compared maps of
the temperature distribution in the crystalline basement
surface; the distribution of calculated average thermal
conductivity for the Cenozoic, Mesozoic, and younger
Paleozoic rocks; and the geothermal gradient pattern for
the Mesozoic sequence to the geothermal gradient pattern
for the entire stratigraphic section. They concluded that
the estimated mean values of thermal conductivity for the
entire section of Mesozoic and younger Paleozoic rocks do

not show any major regional variation that would explain


the significant regional pattern of geothermal gradient.
It seems most improbable that sources of heat within the
rocks due to chemical reactions, mineralogic transformations, or radioactivity could exert more than a minor perturbation on the geothermal gradient pattern. The
distribution of shales and lignitic coals with above average
radioactivity and the regions of highest geothermal gradient in northeastern Alberta appear to be mutually exclusive. Clay mineral transformations would be confined to
the deeper strata, however, which occur in the western part
of the basin where geothermal gradients are the lowest. No
record of recent volcanism occurs in the high geothermal
gradient region of northeastern Alberta. The most recent
volcanic rocks are those of the Crowsnest Member of late
Early Cretaceous or early Late Cretaceous age, which is of
limited extent and confined to the southernmost Alberta
Foothills. In the area of the Crowsnest Member, geothermal gradients range from less than 15C/km (0.8F/100
ft) to greater than 35C/km (1.9F/100 ft), with a "hot
spot" centered over the area of greatest known thickness
around Crowsnest Pass. It will be shown later, however,
that a more plausible explanation for the "hot spot" exists
than relict volcanic heat. Minor Eocene minette and acidic
porphyry intrusions are found south of the Milk River
close to the United States border, and are part of a larger
intrusive suite centered on the Sweetgrass Hills of northern Montana. Geothermal gradients over this intrusive
complex are characteristically low, for reasons which will
become obvious later. If tectonic friction did cause significant heating, it would be confined primarily to the disturbed bek where geothermal gradients are the lowest;
therefore, it is unlikely that this phenomenon has had any
effect on the geothermal gradient pattern.
Surface effects, anthropogenic anomalies such as drilling, and erroneous temperature measurements can be
ruled out. After Majorowicz and Jessop (1981) carefully
selected only about 25% of the data from the AAPG temperature data file based on criteria such as length of time
after cessation of circulation, their geothermal gradient
pattern was effectively the same as that of the Geothermal
Gradient Map of North America (U.S. Geological Survey
and AAPG, 1976). Their map lacked the fine detail, possibly owing to the fewer data points used and the selection
process chosen. Majorowicz and Jessop and this writer
recognize the limitations of the information in the AAPG
temperature data file; but despite these limitations this
paper will show that even the fine detail of the Geothermal
Gradient Map of North America is explainable in terms of
the dominant process controlling the geothermal gradient
pattern in Alberta.
There remains one possible effect which may cause
regional variations in the geothermal gradient pattern of
Albertathe movement of fluids. Effects of hot juvenile
water from the basement can be excluded, unequivocally,
based on the study of stable isotopes of hydrogen and oxygen in the formation waters (Hitchon and Friedman,
1969). This study showed that the observed isotope ratios
result from the mixing of surface water and diagenetically
modified sea water, accompanied by a process which
enriches the heavy oxygen isotope. There was no evidence

Brian IHitchon

of juvenile water. In their study of regional heat flow patterns in the western Canada sedimentary basin, Majorowicz and Jessop (1981) concluded that,
Comparison of heat flow patterns with some of the hydrogeological
phenomena suggests the significant influence of fluid flow in the basin
formations on geothermal features. Low geothermal gradient areas coincide with water recharge areas and high hydraulic head distribution
regions. The phenomenon of upward water movement in the deep strata
and downward flow through much of the Cenozoic and Mesozoic strata
seems to control the regional heat flow distribution in the basin.

Their conclusions relied on the hydrodynamic regime


determined by this writer (Hitchon, 1969a, b). The present
paper expands on the association of geothermal gradient
pattern and hydrodynamics observed by Majorowicz and
Jessop (1981), and it shows that this association operates
at both the regional and local level. In addition, it presents

Figure 2Pliysiographic features of Alberta (from Canadian


Society of Petroleum Geologists, 1975) and generalized topograpliy of province.

717

a demonstrable association of geothermal gradients,


hydrodynamics, and hydrocarbon occurrences.
Hydrodynamic Control of Regional Gleothermal
Gradient Pattern
One of the two main variables affecting the fluid potential distribution is topography (Hitchon, 1969a, b); the
main physiographic features of Alberta, taken from the
Geological Highway Map of Alberta (Canadian Society of
Petroleum Geologists, 1975), are shown in Figure 2,
together with a generalized topographic map of the province. The major upland region, with elevations more than
1,000 m (3,300 ft), comprises the continuous region of the
Rocky Mountains, Rocky Mountain Foothills, western
Alberta Plains, Porcupine Hills, and the Milk River
Ridge, with isolated uplands at the Cypress Hills and Swan
Hills. The major lowland region is in northern Alberta at
the downstream end of the Peace River and Athabasca
River drainage basins, and the Slave River drainage basin,
all with elevations generally less than 500 m (1,650 ft). The
major upland regions are the major regions of meteoric
water recharge, and the major lowland regions are the
major regions of discharge. In general, the region with elevations between 500 and 1,000 m (1,650 and 3,300 ft) represents the region of lateral migration of formation
waters.
Geothermal gradients range from less than 20 C/km
(l.TF/lOO ft) to greater than 50C/km (2.75F/100 ft).
Areas with geothermal gradients less than 20C/km
(I.1F/100 ft) are effectively confined to the Rocky
Mountain Foothills, Porcupine Hills, and the southern
end of the Buffalo Head Hills Upland. The isolated area
with a geothermal gradient of less than 20C/km (1.1 F/
I(X) ft) in the Clearwater Lowland has no data points and is
clearly a computer-contoured extrapolation. However, it
may still represent the true geothermal gradient in that
area because hydraulic-head cross sections (Hitchon,
1969a, Figure 8; Hitchon et al, 1969, Figure 3) and
hydraulic-head maps of individual stratigraphic units
(Hackbarth, 1978a) all show a reversal of the dominant
updip flow from the Alberta basin caused by westward
flow from the recharge area in the Muskeg Mountain
Upland. Relatively cold meteoric water recharging in the
Muskeg Mountain Upland meets the relatively warm formation waters moving updip from the Alberta basin; this
results in a lowered temperature regime where these flows
meet and discharge in the Clearwater Lowland. The
ground-water temperature study of Hackbarth (1978b,
Figure 3) would support this suggestion. With only minor
exceptions, all regions with geothermal gradients of more
than 40C/km (2.2F/100 ft) are found in the northeast
part of the province where most of the elevations are less
than 750 m (2,500 ft), and a significant area is less than 500
m (1,650 ft). The broad region of lateral flow of formation
waters, where elevations range from 500 to 1,000 m (1,650
to 3,300 ft), has geothermal gradients around 30C/km
(1.65F/100 ft). Numerous temperature gradient anomalies exist in this region of lateral flow of formation waters,
and the 30C/km (1.65F/100 ft) contour shown in Figure 1 is the most eastern, continuous (and therefore non-

718

Geothermal Gradients, Hydrodynamics, and Hydrocarbons, Alberta

anomalous) contour. It represents the true regional


geothermal gradient trend in Alberta.
From this comparison of the geothermal gradient pattern and fluid flow regime in Alberta, it is clear that the
major upland areas correspond to major meteoric water
recharge areas with low geothermal gradients; the region
of lateral migration under areas of medium elevation has
medium geothermal gradients; and the major regional discharge in major lowland regions is associated with high
geothermal gradients. Therefore, on a regional basis,
there is a genetic link between the geothermal gradient pattern and the hydrodynamic regime.
Hydrodynamic Control of Local Geothermal
Gradient Pattern
Local topographic features may exert a strong influence
on the regional hydrodynamic regime. It may be anticipated, therefore, that this influence will extend to the fine
detail of the regional geothermal gradient pattern. For
example, iii' the major regional discharge area of the
Alberta basin in northern and northeastern Alberta (Fig-

I elevations > 2000 ft

hydrogeologic cross
sections (see Fig. 5)

ures 1 and 2), the local uplands (recharge areas) comprising the Cameron Hills, Caribou Mountains, Buffalo Head
Hills, Birch Mountains, Stony Mountain, and Mostoos
Hills are dominantly areas of intermediate geothermal
gradient (25-35C/km, 1.35-1.9F/100 ft) within the
regional high geothermal gradient trend (> 40C/km,
2.2''F/100 ft). Furthermore, all "hot spots" with geothermal gradients greater than 50C/km (2.75F/1(X) ft) are
primarily limited to the low topographic (discharge) areas
within the regional high geothermal gradient trend (such
as the Hay River Plain, Fort Nelson Lowland, Fort Vermilion Lowland, Algar Plain, Methy Portage Plain and
the extreme northeastern area of the eastern Alberta
Plains). To confirm the hydrodynamic control of the local
geothermal gradient pattern, two areas are discussed. The
first is the Red Earth region of northern Alberta primarily
within the high geothermal gradient trend and influenced
by the low fluid-potential drain identified by Hitchon
(1969a). The second is the southern part of Alberta in a
region of intermediate and low geothermal gradient and
generally beyond the main effects caused by the low fluidpotential drain.

50 km
0

30 mi

Figure 3Topognphy and location of hydrogeologic cross sections (from Toth, 1978, Figure 1), Red Earth region, northern Alberta.

Brian Hitchon

T6th (1978, 1979) has studied the gravity-induced crossformational flow system in the Red Earth region of northern Alberta, most of which lies within the regional high
geothermal gradient area (see Figures 1 and 2 for location
of study area). Figure 3 shows the present surface topography of the Red Earth region; Figure 4 shows the geothermal gradient pattern sketched from the Geothermal
Gradient Map of North America (U.S. Geological Survey
and AAPG, 1976); and Figure 5 shows the three fluidflow cross sections constructed by Toth (1978, 1979).
The topography of the Red Earth region comprises a Ushaped upland consisting of the Buffalo Head Hills, tlie
northern part of the Utikuma Hills, Trout Mountain, and
the southern end of the Birch Mountains, all with elevations over 6(X) m (2,000 ft). Lowland areas include the valley of the Peace River to the west, the valleys of the Loon
and Wabasca Rivers between the U-shaped upland region,
and the Algar Plain with the Athabasca River to the east
(outside the study area). Geothermal gradients range from
about 15C/km (0.8F/100 ft) over the southern end of
the Buffalo Head Hills to more than 50C/km (2.75F/
100 ft) in two separate areas of the Algar Plain. In general.

15 20 25 30 35 40 45 50 55
Geothermal gradient (C/km)

719

upland regions with elevations greater than 600 m (2,000


ft) have geothermal gradients less than 30C/km (1.65F/
100 ft). Flow directions in the three cross sections (Figure
5) indicate recharge regions at the Buffalo Head Hills,
Trout Mountain, the Birch Mountains, and the Utikuma
Hills. At the Buffalo Head Hills, the meteoric recharge
water penetrates deep into Devonian strata, whereas at
Trout Mountain, the Birch Mountains, and the northern
part of the Utikuma Hills it is deflected near the base of
the Cretaceous section by upward or lateral moving formation waters in Carboniferous or Devonian rocks. The
thickness of the total section overlying the Precambrian is
approximately the same in all these recharge regions.
Therefore, the difference in geothermal gradient between
the southern end of the Buffalo Head Hills ( 15C/km,
0.8F/100 ft) and the other recharge areas in the line of
section (Trout Mountain, = 25C/km, 1.35F/100 ft;
southern end of the Birch Mountains, = 30C/km,
1.65F/100 ft; northern part of the Utikuma Hills,
25C/km, 1.35F/100 ft) can be attributed to the deeper
penetration of meteoric recharge water at the Buffalo
Head Hills. The three other recharge areas are more

hydrogeologic cross
sections (see Fig. 5)

50 km

30 mi

Figure 4Geothermal gradient map of Red Earth region, northern Alberta, sketched from Geothermal Gradient Map of North
America (U.S. Geological Survey and AAPG, 1976), with soft metric conversion.

720

Geothermal Gradients, Hydrodynamics, and Hydrocarbons, Alberta

2-2'

1
Peace
"'"*'

1,000 r

1'

3-3'

I
BUFFALO HEAD HILLS

Loon
River

TROUT
MOUNTAIN

Wabasca
River

-13,000

4 , / ' / - 1,500

500 r,

sea
level

-1,000

c
.0

- -1,500 I

I I t '- J

50 km

H -3,000

30mi

-4,SO0

-1,500 -

* direction of fluid flow


in line of section
Note: for flow directions
toward front or back
of diagrams see
Totti (1978)
K

Cretaceous
and Cenozoic

Carboniferous
and Permian

Devonian

Pe

Precambrian
-1,500

4,500
-1,500

FiguresHydrogeologic cross sections, Red Earth region, northern Alberta (simplified from T6th, 1978, Figures 4,5,6).

Brian Hitchon

affected by warmer formation waters moving updip in the


basin in the Carboniferous and Devonian rocks of the low
fluid-potential drain.
In cross section 1-1 (Figures 4,5), two rivers of comparable size have different geothermal gradient regimes;
both are underlain by similar thicknesses of strata at the
line of section. At the Loon River, the geothermal gradient
is about aS^C/km (1.8F/100 ft), whereas at the Wabasca
River it is greater than 45''C/km (2.45F/100 ft) and close
to a "hot spot" with a geothermal gradient greater than
SS^C/km (S.^F/lOO ft). The Loon River is situated in a
relatively narrow valley between major recharge areas at
the Buffalo Head Hills and Trout Mountain; this has
resulted in cell-type flow with deep penetration of
recharge waters into Devonian rocks on the east side of the
Buffalo Head Hills. As a result, the geothermal gradient in
this discharge area is higher than in the adjacent recharge
regions. In contrast, the Wabasca River, also in a discharge
area, is close to the Trout Mountain recharge region, but it
is on the west side of the wide Algar Plain. Extending
about 80 km (50 mi) beyond the edge of the map area (Figure 3), the Algar Plain contains the Athabasca River. The
relatively warmer formation waters from Devonian rocks
penetrate closer to the surface, and the geothermal gradient at the line of section is 12C/km (0.65F/100 ft) higher
than at the Loon River. These examples show that, even
within the regional high geothermal gradient discharge
area, local recharge and discharge areas strongly influence
the local geothermal pattern.
The second area comprises that portion of Alberta
between lat. 52N and the United States border (Figures 1,
2), together with the adjacent area of northern Montana
which is part of the same geothermal gradient province.
The main physiographic features for the Alberta portion
of the study area are shown in Figure 2. The generalized
topography for the entire area is illustrated in Figure 6A,
and the more detailed topography for southern Alberta is
in Figure 7. The Rocky Mountains, with elevations generally greater than 1,8(X) m (6,000 ft), form the western
boundary (Figure 6A). Contiguous to the east is a wide
band of country with elevations between 900 and 1,800 m
(3,000-6,000 ft). Within this band, the Rocky Mountain
Foothills and the Porcupine Hills generally are over 1,200
m (4,000 ft) -and are continuous with the Rocky Mountains; the Milk River Ridge (> 1,200 m, 4,000 ft) and the
Sweetgrass Hills (> 1,800m, 5,900 ft) are isolated uplands
separated from the Rocky Mountain Foothills Beh and the
Porcupine Hills. Land between 900 and 1,200 m (3,0004,000 ft) comprises the western Alberta Plains and their
extension into northern Montana. To the east, in Alberta,
the eastern Alberta Plains have elevations between 600
and 900 m (2,000-3,000 ft); within these plains are several
small isolated uplands, including the Wintering Hills and
the Hand Hills (elevations > 900 m, 3,000 ft), and the
large upland mass of the Cypress Hills with a maximum
elevation of about 1,470 m (4,820 ft). South of the Cypress
Hills, in northern Montana, the wide valley of the Milk
River separates the Cypress Hills from the Bearpaw
Mountains (maximum elevation 2,105 m, 6,906 ft). The
lowest elevations (< 600 m, 2,000 ft) are in Alberta, close
to the confluence of the Red Deer River and the South Sas-

721

3000 - 6000 ft

< 3000 ft

I > 6000ft

B
Saskatchewan

\ 1

British

Columbia

JCTP
closed high geothermal gradient anomalies
closed low geothermal gradient anomalies
regional Isogradlent contour (C/km)

Figure 6(A) Generalized topography of southern Alberto and


northern Montana. BM = Bearpaw Mountains; HH = Hand
Hills; MRR = MiHc River Ridge; PH = Porcupine Hife; SH =
Sweetgrass Hills; WH = Winterii^ Hflls. (B) Regional geothermal gradient map of southern Alberta and northern Montana,
from Geothermal Gradient Map of North America (U.S. Geological Survey and AAPG, 1976), with soft metric converrion.
Only major geothermal gradient anomalies are shown.

722

Geothermal Gradients, Hydrodynamics, and Hydrocarbons, Aiberta

katchewan River adjacent to the Alberta-Saskatchewan


border.
In Alberta, the 3,(XX)-ft (900-m) contour shows a marked
reentrant on the otherwise generally northwest-trending
boundary of the western Alberta Plains (the reentrant is
occupied by the downstream portions of the drainage
basins of the Red Deer, Bow, and Oldman Rivers and by
the South Saskatchewan River). With respect to the Oldman River, this reentrant is also reflected in the 4,000-ft
and 6,000-ft (1,200-m and 1,800-m) contours near its
headwaters. An area of internal drainage between the
Cypress Hills and the Milk River Ridge is occupied by
Pakowki Lake. Hydraulic-head slice maps (Hitchon,
1969a, Figures 5,6,7) show that southern Alberta is essentially outside the main influence of the low fluid potential
drain, which only truly manifests itself north of the Upper
Devonian southern Alberta shelf complex. The only published hydraulic-head cross section is one approximately
along the valleys of the Oldman and South Saskatchewan
Rivers (Hitchon, 1969a, Figure 8). It is a relatively featureless cross section illustrating the dominant northward flow
as shown by the hydraulic-head distributions in the Upper
Devonian Beaverhill Lake Formation and Woodbend
Group, and the Lower Cretaceous Mannville Group and
Viking Formation (Hitchon, 1969a, Figures 1-4). In

southern Alberta, the fluid flow pattern is dominated by


meteoric water recharge at the Rocky Mountain Foothills
along the western boundary, and in the upland area comprising the Milk River Ridge and the Cypress Hills along
the southern boundary. No data have been published on
the hydrodynamic regime in northern Montana, but it is
assumed that the Sweetgrass Hills and the Bearpaw Mountains are the major recharge regions, together with the
extension of the Rocky Mountain Foothills into that state.
The regional geothermal gradient pattern for the entire
study area is shown in Figure 6B. The numbered contours
are the most eastern, continuous, and therefore nonanomalous, values for each isogradient line; as such they
reflect the regional geothermal gradient. The regional geothermal gradient pattern shows a general easterly or
northeasterly increase in values from less than 20C/km
(l.rF/100 ft) to greater than 33C/km (1.8F/100 ft).
The Rocky Mountain Foothills are characterized by geothermal gradients generally less than Zl'C/km (1.2F/
100 ft); this contour (Figure 6B) shows a major reentrant
associated with the headwaters of the Red Deer and Bow
Rivers, and to the south it extends outward from the
Rocky Mountain Foothills in a trend stretching between
the Porcupine Hills and the Milk River Ridge. The 25C/
km (1.35F/100 ft) isogradient contour shows a similar

well penetrating Precambrlan on


hydrologic cross section
50 km
30 mi

Figure 7Topography of southern Alberta.

Brian Hitchon
reentrant in the middle reaches of the Bow River (and to
some extent where it crosses the valley of the Red Deer
River). This reentrant follows the same trend between the
Porcupine Hills and the Milk River Ridge; it exhibits a
marked reentrant where it crosses the valley of the Oldman
River.
In the northern part of the map area, the 30C/km
(1.65F/100 ft) isogradient contour is extensively convoluted, perhaps related in part to the isolated uplands of the
Wintering Hills and Hand Hills, which have a northeasterly trend, like the isogradient contour. Farther south, this
same isogradient contour shows a sharp reentrant where it
crosses the valley of the South Saskatchewan River. It then
loops around the portion of the Cypress Hills with topographic elevations greater than 1,200 m (4,000 ft); it forms
a narrow reentrant along the valley of the Milk River; and
it forms an extension of the geothermal gradient from the
Porcupine Hills through the Milk River Ridge to the
Sweetgrass Hills. The contour shows a wide reentrant in
the valley of the Marias River before leaving the map area
on the northern side of the valley of the Missouri River.
The highest regional geothermal gradient contour (33C/
km; 1.8F/100 ft) runs parallel and close to the AlbertaSaskatchewan border.
It is now apparent that much of the detail in the regional

723

geothermal isogradient contours can be explained in terms


of local topography Geothermal gradient anomalies
should also be related to topographic features. Figure 8 is a
regional geothermal gradient map of southern Alberta,
with the regional isogradient contours shown in heavy
lines and the high geothermal gradient and low geothermal gradient closed anomalies shown screened. The more
important anomalies are shown also in Figure 6B; they are
numbered in Figure 8 for ease of reference in the following
discussion.
One of the strongest features on the regional geothermal
gradient map (Figure 6B) is the low geothermal gradient
trend stretching along the meteoric water recharge region
from the Porcupine Hills through the Milk River Ridge to
the Sweetgrass Hills. There are two low geothermal gradient anomalies (1,2) along this trend; the strongest is centered over the Porcupine Hills, with geothermal gradients
less than 15C/km (0.8F/100 ft). The two low geothermal gradient anomalies (3, 4) on the northeast side of the
Milk River Ridge may be associated with this recharge
area, or more specifically with recharge from the ridges
north of Etzikom Coulee (Figure 7). A low geothermal
gradient anomaly (5) is present close to the high elevation
(> 1,200 m, 4,000 ft) part of the Cypress Hills. Pakowki
Lake is situated between Milk River Ridge and the Cypress

52'

Closed high geothermal


gradient anomalies
closed low geothermal
gradient anomalies
regional isogradient
contour (F/100ft)
50 km
30 mi

112

Figure 8Geothermal gradient map of southern Alberta, from Geothermal Gradient Map of North America (U.S. Geological Survey and AAPG, W76). Numbers refer to geothermal gradient anomalies discussed in text.

724

Geothermal Gradients, Hydrodynamics, and l-iydrocarbons, Aiberta

Topography (water table)


Elevation

Hydrodynannics

Temperature Regime

Potentiometric surface

Geothermal gradient

o
en

regional
recharge

;-SSv;- ^liQh value

intermediate value

ROCKY
I FOOTHILLS
MOUNTAINS

low value

L = local effect

INTERIOR PLAINS
major regional drainage basin

regional
recharge

regional lateral flow


local
discharge

local
recharge

regional
discharge

local
discharge

Figure 9Schematic model of association of topography (water table), hydrodynamics, and temperature regime. Alberta, Canada.

Brian Hitchon

725

Hills. It is an area of internal drainage and therefore a


local discharge feature. A well-defined high geothermal
gradient anomaly (6) is almost centered over Pakowki
Lake. The low geothermal gradient anomaly (7) north of
the Cypress Hills may be related to recharge from the
Cypress Hills. In addition to the reentrant features of the
regional isogradient contours crossing the valleys of the
Oldman and South Saskatchewan Rivers, this local discharge area contains three high geothermal gradient
anomalies (8, 9, 10). Anomaly 8 lies in a wide valley
bounded by the 2,500-ft (760-m) contour (Figure 7) on
three sides, and is directly over the junction of two major,
deeply buried valleys (Stevenson and Borneuf, 1977).
Anomaly 9 hes at the confluence of the Bow and Oldman
Rivers. Geothermal gradients up to 36C/km (2.0F/100
ft) have been found at anomaly 10, which is centered over
a narrow valley at the headwaters of the Oldman River and
which is well defined by the 4,000-ft and 6,000 ft (1,200-m
and 1,800-m) contours in Figure 7. This topographic situation is conducive to a strong discharge feature and the
associated high geothermal gradient anomaly; it is unrelated to nearby volcanic rocks of the Cretaceous
Crowsnest Member. The S-shape of anomaly 1 is almost
certainly a reflection of the Oldman River discharge
regime.
Southeast of the Wintering Hills are two large low geothermal gradient anomalies (11, 12), both apparently
related to the divide between the Red Deer River and Bow
River drainage basins. Almost between them is a high geothermal gradient anomaly (13) along the lower reaches of
the Red Deer River. The large reentrant in the 25C/km
(1.35F/10O ft) regional isogradient contour near the middle reaches of the Bow River is associated with a strong
high geothermal gradient anomaly (14). Most of the
remaining anomalies in the map area (Figure 8) are in situations where a topographic explanation would be more
speculative. In northern Montana (Figure 6B), the Bearpaw Mountains are associated with a low geothermal gradient anomaly, and the Milk, Marias, and Missouri Rivers
are associated with only high geothermal gradient anomalies.
From this study, it is concluded that both the fine detail
of the regional geothermal isogradient contours and most
of the high geothermal gradient and low geothermal gradient anomalies are closely associated with variations in
local topography. Lower geothermal gradients reflect
higher topographic features which are also local recharge
areas, and higher geothermal gradients reflect lower topographic features which are dominantly discharge areas.

major rock units, and (3) hydraulic continuity throughout


the basin.
At the regional level, high topographic areas have high
water-table elevations with correspondingly high potentiometric surfaces; these areas control the regional recharge
of cold meteoric water and hence have low geothermal
gradients. Areas of medium elevation exhibit regional
lateral flow and intermediate geothermal gradients. The
regional topographic low corresponds with low potentiometric surfaces, and the regional discharge of warm formation waters from deep in the basin results in high
geothermal gradients. Local topographic features cause
perturbations in this picture. River valleys result in drawndown potentiometric surfaces and reentrants on the
regional geothermal isogradient contours, sometimes creating high geothermal gradient anomaUes along the length
of the valleys. Closed drainage basins have closed relatively low potentiometric surfaces and closed high geothermal gradient contours. Contiguous local uplands may
result in closed low geothermal gradient anomalies along
their lengths, analogous to the effects of river valleys. Isolated local uplands show local recharge features and
locally lower geothermal gradients. All these effects may
be modified by the regional and local lithology and permeability, but the general features of the model are evident
throughout Alberta.
Although the writer is convinced of the relations between
topography, hydrodynamics, and geothermal gradient
pattern, there is another aspect of the model which
deserves consideration. Using the same basic assumptions,
it is clear that Alberta's uniform heat flow across the
Precambrian-Phanerozoic boundary is dissipated over a
Phanerozoic section of variable thickness, reaching a generally uniform temperature ( 5C, 41F) at the water
table. As a result, thick sections will have low geothermal
gradients and thin sections will have high geothermal gradients. A plot (not illustrated) of geothermal gradient vs.
depth, using only formation temperature data from strata
adjacent to the Precambrian surface, shows a fairly good
straight-hne relationship. This model is consistent with the
regional geothermal gradient trends, but it does not completely account for the fine detail of the geothermal gradient pattern, because it is clear that in local recharge and
discharge areas water flow either decreases or increases the
geothermal gradient, respectively, relative to what it would
be if based solely on thickness of the Phanerozoic
sequence. Furthermore, this model does not explain the
reverse paleogeothermal gradient described in a later section of this paper.

GEOTHERMAL GRADIENT PATTERN MODEL,


ALBERTA

HYDROCARBON ACCUMULATIONS,
HYDRODYNAMICS, AND
GEOTHERMAL GRADIENTS, ALBERTA

Introduction
From the associations found between topography (water
table), hydrodynamic regime, and geothermal gradient
Klemme (1975) has presented evidence from sedimenpattern in Alberta, at both the regional and local level, it is
possible to generalize these findings into a model, which is tary basins worldwide to show that present and past geoshown schematically in Figure 9. The basic assumptions thermal gradients probably had an influence on the
are: (1) uniform heat generation in the Precambrian base- relative magnitude of hydrocarbon occurrence, especially
ment, (2) uniform regional thermal conductivity of the many major (giant) accumulations. If we extend the find-

726

Geothermal Gradients, Hydrodynamics, and Hydrocarbons, Aiberta

Cretaceous
potentiometric
surface (m)
0

200km

Eocene-Oligocene
potentiometric
surface (m)
0

200 km

area of negative Ah

Figure 10Arabian Gulf petroleum province: topograplty; potentiometric surface and lateral flow directions in Cretaceous and
Eocene-Oligocene strata (simplified from Toth, 1980, Figures 37,38); geothermal gradient (from Klemme, 1975, Figure 26).

ings for Alberta to the evidence presented by Klemme, it


can be reasoned that a genetic link should exist between
hydrodynamics, geothermal gradient pattern, and hydrocarbon accumulation.
A paucity of published papers exists on both the hydrodynamics and geothermal gradients in sedimentary basins.
A striking example of available data, however, is from the
Arabian Gulf. The maps in Figure 10 show the topogra-

phy, potentiometric surface, and lateral flow directions in


the Cretaceous and Eocene-Oligocene. Based on Chiarelli's data, Toth (1980) considered the Arabian Gulf to be an
example of cross-formational ("leaky") discharge of
gravity-driven formation fluids. This is based on the
closed minimum of the Cretaceous potentiometric surface
combined with upward decreasing hydraulic heads which
are indicated by positive head differences between the Cre-

Brian Hitchon

727

Table l.Majoi' Hydrocarbon Occurrences, Alberta


Map No.
{Fig. 11)

Name

Stratigraphic Unit

Table 1. Continued

Initial
Volume
In place
I110* m^l

OIL SAND DEPOSITS:


1 Carbonate Trend Carboniferous and
200.
U. Devonian carbonates
beneath pre-Cretaceous
unconformity
138.1
L. Cretaceous,
2 Athabasca
McMurray Fm.
L. Cretaceous, Grand
43.
3 Cold Lake
Rapids, Clearwater and
McMurray Fms.
L. Cretaceous, Grand
18.9
4 Wabasca
Rapids and Clearwater Fms.
L. Cretaceous, Bluesky and
14.6
5 Peace River
Gething Fms.
HEAVY OILS:
6 Lloydminster

L. Cretaceous, Mannville Gp.

CONVENTIONAL CRUDE OILS:


7 Pembina
U. Cretaceous, U. Colorado
Gp., Cardium Fm.
U. Devonian, Beaverhill
8 Swan Hills
Lake Fm., Swan Hills Mbr.
U. Devonian, Woodbend
9 Redwater
Gp., Leduc Fm.
U. Devonian, Beaverhill
10 Judy Creek
Lake Fm., Swan Hills Mbr.
M. Devonian, Elk Point
11 Rainbow
Gp., Keg River Fm.
Carboniferous, Rundle Gp.,
12 Turner Valley
Turner Valley Fm.
U. Devonian, Beaverhill
13 Swan Hills
Lake Fm., Swan Hills Mbr.
South
M. Devonian, Elk Point
14 Mitsue
Gp., Watt Mountain Fm.,
Gilwood Ss.
U. Devonian, Woodbend
15 Bonnie Glen
Gp., Leduc Fm.
16 Willesden Green U. Cretaceous, U. Colorado
Gp., Cardium Fm.
M. Devonian, Elk Point
17 Nipisi
Gp., Watt Mountain Fm.,
Gilwood Ss.
Leduc-Woodbend
U.
Devonian, Woodbend
18
Gp., Leduc Fm.
SHALLOW NATURAL GAS:
19 SE Alberta Gas U. Cretaceous, Coterado
Gp., Medicine Hat Ss., and
System
A. Medicine Hat post-Colorado Supergroup,
Milk River Ss.
B. Suffield
C. Rest
ASSOCIATED NATURAL GAS:
20 Pembina
U. Cretaceous, U. Colorado
Gp., Cardium Fm.
Carboniferous, Rundle Gp.,
21 Turner Valley
Turner Valley Fm.
(continued)

0.178
1.180

Map]Mo.
(Fig. 11)

Name

Stratigraphic Unit

Initial
Volume
In place
(10* m^)

NONASSOCIATED NATURAL GAS:


U. Devonian, Wabamun Gp. 116.299
22 Waterton
and Carboniferous,
Rundle Gp.
U. Devonian, Beaverhill
23 Kaybob South
107.
Lake Fm., Swan Hills Mbr.
Carboniferous, Rundle Gp.
Carboniferous, Rundle Gp.

67.953

Carboniferous, Rundle Gp.


Carboniferous, Rundle Gp.
U. Devonian. Winterbum
Gp., Nisku Fm.
29 Westerose South U. Devonian, Woodbend
Gp., Leduc Fm.
30 Elmworth
L. Cretaceous, Fort
St. John Gp.

59.4
55.638
52.900

24 Crossfield
25 Jumping Pound
West
26 Edson
27 Brazeau River
28 Panther River

61.985

52.2
51.541

0.412

taceous and Eocene-Oligocene potentiometric surfaces.


The broad belt with geothermal gradients between 25 C/
0.171 km (LSS-'F/lOO ft) and 35C/km (1.9F/100 ft) is situated almost centrally over the area of ascending flow (+
0.169 Ah values of KAh), and it is almost surrounded by a region
of descending flow (-Ah values) with geothermal gradi0.159 ents less than 25''C/km (I.SST/IOO ft). It is probably not
a coincidence that the small "hot spot," with geothermal
0.144 gradients greater than SS^C/km (1.9F/100 ft), lies
directly over the giant Ghawar field. As observed by T6th
0.129 (IMO), "it is also noteworthy that the majority of accumulations are in areas of low values and positive differences
of hydraulic heads"; to which can now be added, "and
0.117
also high regional geothermal gradients." 'f
The model in Figure 9 was designed specifically for the
0.117
Alberta basin; however it is easy to visualize a contiguous
0.111 mirror image with marginal recharge regions having low
geothermal gradients, and a central discharge area having
high geothermal gradients, that would fit the description
0.102 of the Arabian Gulf. Having denwBrtrated the genetic
link between hydrodynamics, geotho-mal gradients, and
hydrocarbon accumiilations in the worlds most prolific
conventional oil province, it is now ^propriate to consider the generally more modest hydrocarbon accumulations in Alberta.
153.
0.207

57.3
179.069
113.
80.866

Alberta

The major hydrocarbon occurrences in Alberta are listed


in Table 1 and shown in Figure 11. The list includes only
conventional crude oil discoveries with more than 100 million m' of initial in-place crude oil, and natural gas discoveries with more than 50 billion m' of initial in-place natural
gas, as well as all major oil sand deposits and the most sig-

728

Geothermal Gradients, Hydrodynamics, and Hydrocarbons, Aiberta

oil sands deposit


I | > I conventlonil crude oil
[

] natural gas

Figure 11Distribution of ms|)or liydrocarbon occurrences.


Alberta. Numbers refer to details in Table 1.

major flow direction


Wi'^'M '' sands

'UCJ

__L'

-"

i^ure 12Schematic situation for Alberta's oil sand deposits.

nificant heavy oil accumulation. Volumes of in-place oil


sand deposits were obtained from Outtrim and Evans
(1978), and volumes of the remainder of the hydrocarbon
occurrences are from Energy Resources Conservation
Board (1980). Many fields have production from several
stratigraphic units, but only the most significant are listed
in Table 1. Conventional crude oils comprise 56% of the
total provincial in-place light and medium crude oil; the
Lloydminster field is 23% of the total provincial in-place
heavy crude oil; and natural gas occurrences represent
28% of the total provincial in-place natural gas (shallow
natural gas = 9%; other = 19%).
The oil sand deposits occur as biodegraded, waterwashed bitumen in rocks adjacent to the pre-Cretaceous
unconformity in the northeast part of the province, and
they account for 98.5% of the crude oil occurrences in
Alberta. Each deposit is a major (giant) hydrocarbon
accumulation by world standards. Deroo et al (1977) have
shown that the bitumen is not a very immature crude oil,
and that the organic matter in contiguous and contemporaneous shales has a low level of diagenesis, which correlates with a low yield of hydrocarbons. Thus, the less
deeply buried sedimentary rocks at the Alberta basin margin could not have contributed, by themselves, the enormous accumulations of hydrocarbons found in the oil
sand deposits. It is necessary to admit that there is a major
contribution from source rocks deeper in the basin. The
relative contribution of Upper Devonian, Lower Cretaceous, and Upper Cretaceous source rocks cannot be
ascertained at this time.
The special situation which results in such large accumulations is shown schematically in Figure 12. Potential
source sediments are the Upper Devonian shales in northwestern Alberta (> 300 m, 1,000 ft) and the Cretaceous
shales (> 1,500 m, 5,000 ft) situated deep in the western
portion of the Alberta basin. Hitchon (1969a) has shown
that there is a hydraulic divide from the Sweetgrass Arch
to the highest elevation in the eastern limit of the exposed
Frecambrian shield; its approximate position is visible on
the three hydraulic-head slice maps (Hitchon, l%9a. Figures 5, 6, 7). The major flow directions related to this
divide are shown in Figure 12. The oil sand deposits of
northeastern Alberta are all situated at the foci of a series
of converging flow paths that drain most of the Alberta
basin, including the two major potential sequences of
source sediments. The lack of oil sand deposits at the
major regional discharge in Manitoba may be attributed
mainly to the presence of the hydraulic divide, which cuts
off a major source rock sequence in the Cretaceous of
western Alberta. Thus, the combination of suitable
mature source sediments, a basin-wide hydraulic system
with a major hydraulic divide, widespread reservoir rocks,
and suitable timing, has resulted in these major hydrocarbon accumulations. The geothermal gradients associated
with these deposits reflect this combination of parameters.
With exception of the currently ill-defined carbonate
trend, all major oil sand deposits are associated with high
geothermal gradient anomalies. The Athabasca, Cold
Lake, and Wabasca deposits all occur within the regional
high geothermal gradient area with gradients generally
greater than 40C/km (2.2F/100 ft); several "hot spots"

Brian Hitchon

729

> 3000 feet


2000-3000
1000-2000

I < 1000
, Peace River
oil sand deposit

Geothermal
gradient {"Clkm)

i>33
I

I 30-33
25-30
<25

MontafKuse WHtTEMUD

1,000 r *'*''

'^"-'-^_<^

Peace
River

Cadotle
Lake

Lubicon
River

2,000
500 -

hydrauiic
head
--(meters
-"i^rth
above datum) "g
^ direction of ^ BOO
fluid flow
s
" ^

oil sand
deposit

-1,000
"

1,000

3,000
4,000

-1,500
-2,000

Figure 13Topography and geothermal gradient map, slcetched from Geothermal Gradient Map of North America (U.S. Geological
Survey and AAPG, 1976), with soft metric conversion, and hydrogeologic cross section (simplifled from Hitchon, 1974, Figures 12,
13), Peace River area. Alberta.

730

Geothermal Gradients, Hydrodynamics, and Hydrocarbons, Alberta

within this area have gradients greater than 50C/km


(2.75F/100 ft). The general T-shape of the Athabasca
and Wabasca oil sand deposits is reflected in the associated
40C/km (2.2F/100 ft) isogradient contour in that area.
The Primrose Air Weapons Range, situated between the
Athabasca and Cold Lake oil sand deposits, has geothermal gradients generally greater than 40C/km (2.2F/100
ft), and it is known to contain oil sand occurrences. The
Cold Lake oil sand deposit is centered over a 50C/km
(2.75F/10O ft) "hot spot." South of the Cold Lake
deposit, the heavy oil area around the Lloydminster field
is associated with gradients of about 40C/km (2.2F/100
ft). The Peace River area (Figures 1,2) is discussed by Hitchon (1974); Figure 13 is from that paper, together with the
geothermal gradient pattern sketched from the Geothermal Gradient Map of North America (U.S. Geological
Survey and AAPG, 1976). The hydrogeologic cross section shows that meteoric water infiltration over the Whitemud Hills and the col between the Buffalo Head Hills and
the Utikuma Hills has resulted in reduced formation water
sahnities and geothermal gradients less than 25C/km
(1.35F/100 ft) in closed low geothermal gradient anomalies.
Where the line of section crosses the deep valley of the
Peace River, hydraulic heads as low as 300 m (980 ft) above
datum and a closed high geothermal gradient anomaly
occur (gradients > 33C/km, I.S'F/IOO ft). A similar
anomaly overlies Loon Lake near the eastern end of the
cross section. The trilobate Peace River oil sand deposit is
situated within a pattern of three closed low geothermal
gradient anomalies with gradients less than 25C/km
(1.35 F/100 ft), and its trilobate form is outlined by a pattern of three closed high geothermal gradient anomalies
(gradients > 33C/km, 1.8F/100 ft), including one with
a gradient greater than 35C/km (1.9F/100 ft).
From this brief discussion of Alberta's oil sand deposits,
it is clear that their position in the basin is closely related to
the hydrodynamic and geothermal gradient regimes at
both the regional and local level.
In Alberta, lateral migration, under areas of medium elevation with medium geothermal gradients, is dominated
by a low fluid-potential drain. This drain developed in the
medium-depth part of the basin as a resuh of a thick
sequence of highly permeable Upper Devonian and Carboniferous carbonate rocks (Hitchon, 1969a). Six of the
major conventional crude oil occurrences and five of the
nonassociated natural gas occurrences are found in these
rocks. The low fluid-potential drain comprises drainage
channels of different fluid-potential subsystems; excellent
examples of these are the carbonate complexes of the
Upper Devonian Woodbend Group and Beaverhill Lake
Formation (Hitchon, 1969b). These hydrodynamic subsystems are associated with subtle temperature differences.
The general distribution of the Upper Devonian Woodbend Group and Beaverhill Lake Formation carbonate
comple^Jes is shown in Figure 14. The pressure profiles of
fluids in three of the complexes are illustrated in Figure 15.
The data for the Rimbey-Meadowbrook/Grosmont pressure profile extend from deep in the basin adjacent to the
disturbed belt, to the pre-Mesozoic subcrop of the Gros-

WoodlMfld Group
ctrbonata oomplax
BaavaiMII llta Formatlor
carbonata complax
100 km

Figure 14Genenitized distribution of Upper Devonian Woodbend Group and Beaverhill Lalie Formation carbonate complexes (from Alberta Society of Petroleum Geologists, 1964).
Numbersreferto fields listed In Table 1 and shown in Figure 11.

mont complex at about lat. 57N. Excellent hydraulic continuity exists along approximately 600 km (37S mi) of this
carbonate complex system with the elevation at zero pressure corresponding closely to the outcrop of Upper E)evonian rocks at Fort McMurray. The Bashaw carbonate
complex is hydraulically isolated within a shale basin, and
the pressure profile shows pressures 3,000-4,000 kPa (435580 psi) above that of the Rimbcy-Meadowbrook/
Grosmont complex, with a slightly steeper hydraulic
gradient. The Woodbend Group's Windfall carbonate
complex and the more extensive underlying Swan Hills
carbonate complex of the BeaverhiU Lake Formation are
lithologically and hydrologically continuous and are
treated as a unit in the pressure profile diagram. They
exhibit a pressure profile similar to that of the Bashaw
complex and are similarly situated with respect to shale
aquitards.
The temperature-depth profiles for these same carbon-

Brian Hitchon

731

I n i t i a l r e s e r v o i r p r e s s u r e (psi)
1,000

2,000

3,000
-1

4,000
1

sea
level

sea
evei

Bashaw carbonate complex


2,000
oil field

A gas field

subsea = 406 - ( 0 . 0 8 3 9 x pressure)


Leduc-Woodbend

correlation coefficient =

a -1000

0.982

-3,000

Swan Hills
Swan Hills South
Judy Cree(<

Westerose South
Rinnbey-I^^eadowbrool^/Grosmont
carbonate complex

6,000
c -2000
o

; 0 oil field

A gas field

;subsea = 351 - ( 0 . 0 9 4 9 x pressure)


; correlation coefficient =

-0.998

UJ

Kaybob South

8,000

-2500
Windfall-Swan Hills carbonate complex
D oil field

9,000

gas field

subsea = 588 - ( 0 . 0 9 0 0 x pressure)


correlation coefficient =

10,000

-0.996

-11,000
5,000

15,000

20,000

26,000

30,000

_L
35,000

I n i t i a l r e s e r v o i r p r e s s u r e (kPa)

Figure ISPressure profiles for three Upper Devonian carbonate complexes, Alberta.
ate complexes are given in Figure 16. The temperatureelevation (relative to sea level datum) profiles are shown in
Figure 17. Although there is more scatter to the data than
in the pressure profiles, both plots for the two high pressure systems show steeper temperature gradients than that
in the low pressure Rimbey-Meadowbrook/Grosmont
carbonate complex.
In the plot of reservoir temperature vs. depth, the regression lines are very close together, which might be construed
to indicate that the temperature within all three carbonate
complexes is solely a function of depth. When temperature is plotted against subsea elevation (by analogy to the
pressure profiles), however, the regression lines indicate
that the two high pressure systems can be distinguished
from the low pressure system. To some degree this would
suggest that the temperature profiles are related to the
pressure profiles. All three carbonate complexes exhibit a
good hydraulic connection with the warmer, deeper fluids

at their southwestern ends. In the Bashaw and WindfallSwan Hills carbonate complexes, the surrounding shales
are effective aquitards; therefore, the pressures and temperatures cannot be dissipated readily, resulting in relatively high pressure profiles and steeper pressure and
temperature gradients than in the .RimbeyMeadowbrook/Grosmont carbonate complex. Basically,
the Rimbey-Meadowbrook/Grosmont carbonate complex acts as a short circuit to the hydrauUc system. The
effect of this short circuit is probably not visible on the
geothermal gradient map. However, a narrow, long reentrant exists on the 30C/km (1.65F/100 ft) isogradient
contour over the middle portion of the RimbeyMeadowbrook carbonate complex and a wide reentrant
on the 33C/km (l.8F/100 ft) isogradient contour near
the southern end of the Grosmont carbonate complex.
Neither situation has a more plausible explanation at this
time.

732

Geothermal Gradients, Hydrodynamics, and Hydrocarbons, Alberta

Reservoir tennperature (F)


32
0

60
1

126

100
1

'

150

176

200

225

250

276

300

1,000
Rimbey-Meadowbrooi</Grosmonl carbonate complex
\

500

0 oil field

^A

A gas field

depth = - 3 1 3 +(32,8 x temperaturel


\ A

RedwaterQ

2,000

correlation coefficient = 0,966

- 3,000
-

\ / ^

1000

o\

4,000

1500

/X^^o
\ \ 0 0 ^
/
te*.
fV-Leduc-Woodbend
Bashaw carbonate complex
oil field

jT^ 0
\ A O

,-^

\^

depth = - 4 , 4 7 + (29,4 x
temperature)

2000

_E
x:
D.
O
TO
. 2500
O

A gas field

\\

correlation
coefficient = 0.962

5,000

.
_ 6,000
.
;
- 7,000 =

Bonnie Glen

,c
Westerose
South

^ ^
iR,
\ ,

>
CO

o
DC

a_

Q.
lU

8,000 Y.

Swan Hills
D n S w a n Hills South

O
"
>
9,000 [E

OTXx [] Judy Creek


V,^

3000

A
A

\ X
\\\

0
'-'

\ \

- 11,000
-

\ \ ^
D N\\

B
3500

10,000

Kaybob South

\
\ \

12,000

^ . \f
13,000
4000

14,000
Windfall-Swan Hills carbonate complex
4500

a oil field

'

gas field

depth = 49,1 +(28,4 X temperature)

\
\-

15,000

correlation coefficient = 0.886

5000
0

10

20

30

40

60

60

70

80

90

100

110

120

130

140

160

16,000

160

Reservoir t e m p e r a t u r e (C)

Figure 16Temperature-depth profiles for three Upper Devonian carbonate complexes, Alberta.
The largest natural gas occurrence in Alberta is designated the southeast Alberta Gas System by the Energy
Resources Conservation Board. It is located in southeastern Alberta between the Cypress Hills and the uplands
comprising the Wintering Hills and the Hand Hills. Production is from the shallow (350-500 m, 1,150-1,650 ft)
Upper Cretaceous (= Santonian) Milk River and Medi-

cine Hat sandstones. The sedimentary rocks are immature


at this depth (R^, = 0.43), and maximum maturity in strata
at the Precambrian basement ( 2,100 m, 6,900 ft) is only
RQ = 1.46 (Fuex, 1977). The chemical and isotopic composition of these natural gases (Fuex, 1977; Rice and Claypool, 1981), plus the immaturity of the adjacent rocks,
indicates a predomincuitly biogenic origin. Schwartz et al

Brian Hitchon

733

R e s e r v o i r t e m p e r a t u r e (F)
!2

50
1

500

75
i

100
1

125
1

150
1

176
1

200
1

226
1

250
1

275
1

300
1
1,000

^ ^
sea
level

&

O oil field

a \ A

-500

Rimbey-Meadowbrook/Grosmont carbonate complex <

\
RedwaterQ

A gas field

sea
level

subsea = 8 0 0 - ( 2 8 . 2 X temperature)
correlation coefficient =
\

- -1,000

-0.969

/ ^

o \^

- -2,000

*HS? O
1 -1000
n

V t

/ ^ * ^ 0
i /

Bashaw carbonate complex

>

a
<o
v>

oil field

0 -1500
<u

0
- B o n n i e Glen

A gas field

-4,000

wpttPmo^^iiiO
westerose
N ^
South
\ ^

a
<u
n
6,000
o

Swan Hills
D c w a n Hills South
C b - J u d y Creek

>

_>

< \

s u b s e a = 5 6 4 - { 2 4 . 9 X temperature) 1
s;
correlation c o e f f i c i e n t = - 0 . 9 6 1
|

- -3,000

EhLeduc-Woodbend

n
<D
c
O
O -2000

- 6,000

o
^
c
o

7,000

o
^

>

>

o
LU

DX^B

J8,000

^ ^ < ; ^ Kaybob South


-2500

\ *\ \ \

Windfall-Swan Hills carbonate c o m p l e x t'


D oil field

-3000

gas field

subsea = 510 - ( 2 4 . 9 x temperature)

-10,000

correlation c o e f f i c i e n t = - 0 . 8 4 4

3600
0

10

20

30

40

50

60

9,000

^ \ ,

70

80

90

100

110

120

.130

140

-11,000

160

16 0

R e s e r v o i r t e m p e r a t u r e (C)

Figure 17Temperature-elevation (relative to sea level datum) profiles for three Upper Devonian carbonate complexes, Alberta.

(1981) have made an extensive study of the regional variations in hydraulic conductivity and hydraulic head for
eight stratigraphic units in this area, as well as isotopic and
major-ion analyses of the formation waters. Their study
confirms the broad regional pattern of northward flow
found by Hitchon (1969a, b). Extensive cross-formational
flow was recognized by Schwartz et al (1981), but it was
not documented. The hydrogeologic cross section (Figure
18) illustrates the extent of the cross-formational flow in
the Mesozoic sequence only; it also occurs in Devonian
strata (cf., Hitchon, 1969b, Figures 1, 2). The section
south of the South Saskatchewan River is based on the
hydraulic-head maps from Schwartz et al (1981); to the
north, it is based on selected profiles compiled solely for

illustration purposes, and it should be regarded only as a


preliminary section.
Recharge occurs at the south end of the section, in the
Cypress Hills, in the surface-water divide between the
South Saskatchewan and Red Deer Rivers, and at the
Hand Hills. Discharge occurs at the major river valleys
and Pakowki Lake. Formation waters near the base of the
Lower Cretaceous strata have generally high hydraulic
heads, compared to those in the Bow Island Formation
near the top of the Lower Cretaceous section. Toward the
north end of the cross section, this relatively low
hydraulic-head zone moves up into the Second White
Speckled sandstone. Throughout the southeast Alberta
Gas System in the line of section, hydraulic heads in the

1,000 -

3"

<
3
.
O
a)

500

Q.

5"
3

0)

<

Q.
O

a
><

-500 -

3
0)

0)
3

a.
X

<
-6,000

Q.

8
o

-2,000

fl)

Lower Cretaceous

Tertiary
MR

Milk River Sandstone

Upper
_ - M H - Medicine Hat Sandstone
Cretaceous
2WS 2nd White Specified Shale
BFS- Base of Fish Scale Sandstone

equipotential line (m above datum)

Jurassic

flow line

Carboniferous

/////y/

Devonian

Precambrian

o individual hydraulic head data point

Pe

3
(A

approximate position of hydraulic low

approximate position of hydraulic high

Low/er Paleozoic

SE Alberta gas system

Figure 18Hydrogeologic cross section across southeast Alberta Gas System natural gas occurrence (Upper Cretaceous Milk River and Medicine Hat Sandstones).

0)

Brian Hitchon

Milk River and Medicine Hat sandstones are effectively


the same. The flow is upward from the Milk River Sandstone into the overlying Pakowki Formation, and downward from the Medicine Hat sandstone into the Lower
Cretaceous or Second White Speckled sandstone (depending on the position in the cross section). The southern limit
of the southeast Alberta Gas System occurs where the flow
direction changes abruptly in the recharging waters north
of the Cypress Hills. This change is illustrated in Figure 18
and in the hydraulic-head profile along the Milk River
sandstone (Figure 19).
Throughout almost the entire length of the major natural gas occurrence in these sandstones, the hydraulic gradient is low compared to the gradient at the southern and
northern margins. At these margins, the flow directions,
and consequently the hydraulic gradients, change abruptly
owing to recharge from the Cypress Hills and the Hand
Hills, respectively It is clear, therefore, that it is the present
hydrodynamic regime that controls the location for this
shallow gas occurrence. It is debatable whether it is neces-

N
1000[-/'<i*owt/
Lake
-^

I
I

CYPRESS
NLLS

S. Saskalcliewan
River

Red Deer
River

900

800

700

600
50

100

150

200

250

300

350

Distance along hydrogeologic


cross section (km)

Figure 19Hydraulic-head profile in Milk River sandstone


along hydrogeologic cross section in Figure 18.
sary to invoke exsolution of previously formed biogenic
gas following uplift and erosion during the Tertiary (Rice
and Claypool, 1981). The hydrodynamic regime correlates
well with the geothermal gradient anomaly map (Figure
8).
Of the remaining major hydrocarbon occurrences, Rainbow may be associated with a high geothermal gradient
anomaly which is related to discharge at the Zama and
Hay Lakes, immediately to the north. Crossfield is located
over a particularly strong high geothermal gradient anomaly. Four of the fields are situated within the Foothills Belt
and are structural traps in Carboniferous or Upper Devonian rocks. The hydrodynamic pattern in the Foothills
Beh is not well known, but the region is hydraulically continuous with the Alberta basin. None of the others appear
to be directly linked with these anomalies.
Minimal drilling has taken place in the Rocky Mountains; therefore, the hydrodynamic regime is essentially
unknown. Some indication may be obtained, however,
from study of thermal and mineral springs, most of which
are associated with various thrust faults (van Everdingen,
1972). Determination of the stable isotopes of hydrogen

735

and oxygen in some of the hot springs (Hitchon and Friedman, 1969) proves an origin from local precipitation.
Temperatures up to 53.9C (129F) have been recorded in
some of the hot springs (van Everdingen, 1972). This evidence suggests that the large differences in topography
between the highest elevations and the major valley bottoms have resulted in cell-type flow. This flow results in
meteoric water penetrating deeply into the geologic section, heating up, and emerging as hot springs where thrust
fauhs act as short circuits. If this suggestion is correct, the
hydrologic regime in the Rocky Mountain- may not be
connected with that in the Foothills Belt.
Summarizing the relation between hydrocarbon accumulations, hydrodynamics, and geothermal gradients in
Alberta, it is evident that moving formation water plays a
crucial role in hydrocarbon migration and accumulation,
and in the development of the geothermal gradient pattern. Where detailed study of the hydrologic regime is
available, the flow paths can commonly be related genetically to known hydrocarbon accumulations and to the
geothermal gradient pattern. One difficulty in developing
this thesis is that the level of knowledge and detail available on the temperature regime is far less than that for the
hydrologic regime. The perceptive reader will have realized that for some of the cases cited, the hydrologic regime
is presented in three dimensions whereas the geothermal
gradient pattern is generally known only in two dimensions; i.e., the gradient is the average from maximum
depth to the surface. Despite these difficulties, the logical
link between the hydrodynamic and geothermal regimes
and hydrocarbon accumulations is difficult to deny. Better
knowledge of the geothermal regime will allow a more
detailed analysis of the associations to be made, but is not
likely to alter the fundamental relationships.
PALEOGEOTHERMAL REGIMES IN ALBERTA
Introduction
Before discussing the paleogeothermal regimes in
Alberta, it is important to understand the Phanerozoic
evolution of the province. McCrossan and Porter (1973)
have summarized the grand-scale geology of North America, and the following comments are based on their review.
They divided the Phanerozoic into four megasequences:
lower Paleozoic, upper Paleozoic-Triassic, JuraCretaceous, and Tertiary. The regional distribution of
existing rocks, their thickness and gross lithology, and the
associated orogenic belts for each of these megasequences
in Alberta and adjacent areas, is illustrated in Figure 20.
During the time represented by the lower Paleozoic
megasequence (Cambrian to Devonian), the continent was
covered almost completely by marine sediments. The
gross hthologic characteristics of this interval are continuous, both laterally and vertically, and comprise the most
extensive development of marine carbonates of any megasequence. Sah pans are related to each period within this
megasequence. In western Canada, the Caribooan orogenic belt is part of a continuous chain of orogenic belts
around the craton. More than 3,000 m (9,800 ft) of sediments were deposited in the trough between this orogenic

736

Geothermal Gradients, Hydrodynamics, and Hydrocarbons, Alberta


Lower Paleozoic

Upper Paleozoic-Triassic

Jura-Cretaceous

Tertiary

erogenic belt ^ i s isopach (100m)

Figure 20Regional distribution, thickness, and gross lithologic characteristics of existing rocl^s, and associated orogenic belts for
four megasequences in Phanerozoic of Alberta and adjacent areas (from McCrossan and Porter, 1973).
twit and the shelf sediments on the craton in Alberta.
In the upper Paleozoic-Triassic megasequence (Carboniferous to Triassic), deposition of marine sediments is
much more restricted around the craton, but an almost
com{dete sequence is present in western Canada. The gross
lithologic assemblage is more variable than that in the preceding megasequence, and major interregional unconformities are more numerous and of shorter duration.
Some evidence exists for a westward tih of the craton.
More than 1,500 m (4,900 ft) of dominantly carbonate
sediments are found in the trough between the Inklinian
orogenic belt and the shelf sediments on the craton in
Alberta.
The Jura-Cretaceous megasequence is characterized by
the deposition of clastic sediments, mostly marine, but it
includes a large continental component. Off the Atlantic
and northern coastal margins, vast wedges of clastic sediments prograde seaward, whereas in western Canada
more than 4,500 m (14,700 ft) of clastic sediments were
deposited in a trough on the eastern side of the Columbian
orogenic belt. This contrast suggests a high-standing continent tilted toward the west.
During the time represented by the Tertiary megasequence, the positive movement of the continental plate
resulted in withdrawal of the epicontinental seas to the
coastal margins, and only continental clastic rocks are represented in western Canada.
Porter et al (1982) have revised their concepts of the
Phanerozoic history of the western Canada sedimentary
basin. They now suggest that strata deposited from the
late Proterozoic to the Middle Jurassic at the western
extremities of the "basin" formed a continental terrace
wedge. Therefore, the Caribooan and Inklinian orogenic
belts are only minor events; it is only the Columbian and
Laramide orogenies, which are related to the Middle
Jurassic to Paleocene foreland basin, that can be described
as important orogenic events. Regardless of this refinement, it is clear that immediately prior to the Laramide
orogeny the craton shelf in Alberta was covered by a sub-

stantial wedge of Paleozoic and Mesozoic sediments


which thickened gradually westward toward the position
of the present Foothills Belt. West of this position the sedimentary sequence thickens rapidly.
The Laramide orogeny in the Canadian Cordillera
occurred in two pulses (Taylor et al, 1964). An initial
minor pulse commenced in the late Late Cretaceous and
caused the first uplift of the Cordillera. A consequent
thick sequence of continental elastics was deposited during
the Paleocene in the Foothills area and across the Alberta
Plains; the depocenter was in the western Alberta Plains.
The second, a major pulse, began m the early Eocene, and
much of the Paleocene clastic sequence was subsequently
removed by erosion. The time associated with this second
Laramide orogenic pulse best represents the period of
maximum burial in Alberta; it is also the time when subaerial relief was probably at its maximum.
Early Eocene Paleogeothermal Regime
Because of the extensive post-Eocene erosion, two sets of
data are required to determine the geothermal regime at
the time of maximum subaerial relief. First, the thickness
of eroded strata needs to be calculated, and second, the
degree of organic metamorphism must be determined on a
regional basis for one specific time interval. Hacquebard
(1977) has made both determinations based on coal data in
Alberta. Coal rank is an excellent measure of the degree of
organic metamorphism. Once a specific rank has been
attained it cannot be reversed. Agencies that control rank
are pressure, temperature, and time of heating; all are
directly related to the depth of maximum biuial, and temperature is the most important factor.
In the Alberta Plains, outcrop and near-surface coals
range in age from Late Cretaceous to early Tertiary, and in
rank from lignites in the east to high volatile C bituminous
in the west. Lignites and subbituminous coals are ranked
on the basis of their calorific value, but because calorific
value and moisture content are closely related, they may

Brian ^ tchon
also be ranked on the basis of their moisture content,
which decreases in a regular pattern from 30% in the east
to 10% in the west.
Using correlations between moisture content and depth
of burial from Germany, Hacquebard (1977) determined
the maximum depth to which these near-surface coals had
been buried, which is, of course, a direct measure of the
thickness of eroded strata. He also determined the rank of
coals in the Lower Cretaceous Mannville Formation
(Aptian-Albian stage) across southern Alberta. Depth of
Mannville coals ranges from 600 m (1,950 ft) in the east to
3,400 m (11,000 ft) in the west, close to the Foothills Belt;
the corresponding ranks of the coals are R^, = 0.42 and
R^ = 1.58. Based on the relation of rank and maximum
paleotemperature, he calculated the paleogeothermal gradient across the province at the time of maximum burial.
Hacquebard found that at the time of maximum burial,
the geothermal gradient ranged from 28C/km (1.55F/
100 ft) in the east to less than 22C/km (1.2F/100 ft)
close to the present Foothills Belt. It should be noted that
this is a subdued replica of the present regional geothermal
gradient pattern. It existed immediately prior to the second pulse of the Laramide orogeny, before the erosion of
much of the Paleocene clastic sequence but after the initial
uplift of the Canadian Cordillera (e.g., when subaerial
reUef was probably at a maximum). Therefore, with
higher relief the hydraulic-head regime would be
enhanced, compared to the present, and as a consequence
the geothermal gradient pattern would be relatively subdued.
Hacquebard (1977) determined the maximum paleotemperature for 144 oil and gas fields in Alberta. From these
data and the calculated paleogeothermal gradient, it is
possible to calculate the thickness of eroded sediments and
the elevation of the land surface at the beginning of the
Eocene, before erosion of much of the Paleocene. The
sequence of calculations is as follows:
Thickness of eroded strata

= depth at maximum burial


- present depth of burial
= (paleotemperature-^ paleogeothermal
gradient)
- present depth of burial,

and
Elevation of early Eocene surface

= thickness of eroded strata


+ present ground elevation

Calculated values for Swan Hills field are shown graphically in Figure 21; Figure 22 is the early Eocene surface
determined by this method. Figure 23 shows the thickness
of eroded strata; and Figure 24 displays the regional geothermal gradient pattern that developed as a result of the
early Eocene surface in Figure 22. Only data in the Alberta
Plains area have been used.
For some geologists the most contentious aspect of this
exercise may be the thickness of eroded strata. Taylor et al
(1964) stated that
Erosion on a grand scale accompanied and followed the uplift of the
mountains. Only an insignificant portion of the enormous amount of
debris that moved eastward along drainage channels in Eocene and Oligocene time remains on the Canadian Plain:
An enormous amount of

737

Surface
Meters
temperature
3000
16C
-I12611

Depth of eroded
^ strata since
Surface t maximum burial
temperature

sea
level

PaleoPresent
temperature temperature
101C
104C

Swan Hills
1429 -(Beaverhill
Lake A)

Figure 21Illustration of method of calculation of early Eocene


surface elevation and thickness of eroded strata, Swan Hills
field.

debris was eroded from the rising mountains and their flanking plateaux
and plains during the Eocene and Oligocene epochs. These clastic materials must have formed aggradation plains and basin deposits reaching far
from the mountains, yet Uttle remains of them.

Hacquebard (1977) estimates maximum thickness of


eroded sediments at about 3,000 m (9,800 ft) in the west
and about 800 m (2,600 ft) in the east, based on his coalification diagram (his Figures 3-8).
In a study of the area almost immediately north of that
examined by Hacquebard, Magara (1976; see also Kumar,
1979) determined the thickness of sedimentary rocks
removed by erosion by the use of shale-compaction data,
such as the shale transit-time values recorded on sonic
logs. He concluded that as much as 1,400 m (4,600 ft) of
sediments had been removed. Figure 23 indicates approximately 2,500 m (8,200 ft) of erosion in the area studied by
Magara.
Using recently acquired data on coal moisture from a
deep-coal exploration program carried out by the Alberta
Research Council, Nurkowski (personal communication)
has also determined the thickness of eroded sediments
using a technique similar to that of Hacquebard (1977).
Nurkowski's data were contoured using 9-township averages, and his isopachs have a regional trend similar to Figure 23, although the values are about 500 m (1,600 ft)
thinner. It should be noted that no one attempting to determine the thickness of eroded sediments, including the
writer, has considered isostatic uplift in their calculations.
Neglect of this factor, in part, may account for the different values obtained by different methods.
Although the exact values of the thickness of eroded
sediments may be debatable, the regional trends are not,
and it is only these regional trends which will now be discussed.
About early Eocene time, when topographic relief was
at a maximum, the land surface (Figure 22) sloped generally northeastward in the central part of the province and
in a more easterly direction in southern Alberta. Presum-

738

Geothermal Gradients, Hydrodynamics, and Hydrocarbons, Alberta

ably, this change in slope direction reflects a change in


trend of the ancestral Rocky Mountains, which is evident
in the present Rocky Mountains. The inferred drainage on
this surface is illustrated in the inset map in Figure 22. Sufficient detail exists to indicate the major drainage divides
and river basins; the small area of southwestward drainage
is problematic and likely is an artifact of the method of
constructing the surface topography. The inset map shows
one minor and two major drainage basins trending northeastward in the central part of the province. Southern
Alberta is drained by two river systems which flow northeastward in their upper reaches but finally leave the province flowing in an easterly direction. Barton et al (1964)
have reconstructed the preglacial topography and drainage of western Canada, which they considered to be Late
Tertiary or early Pleistocene in age; it is shown in the main
map in Figure 22, taken directly from their paper. A
remarkable similarity exists between these two drainage
systemsgeneral drainage directions and the number of

significant drainage basins. This implies that the Late Tertiary or early Pleistocene drainage of Alberta, as constructed by Barton et al (1964), was in existence in the early
Eocene, and had not changed in any fundamental feature
for more than 50 m.>. During this time a tremendous
wedge of Tertiary and Upper Cretaceous rocks was
removed by erosion (Figure 23).
The regional paleogeothermal gradient pattern during
the early Eocene (Figure 24) is based on the work of Hacquebard (1977), and it shows gradients ranging from
21C/km (1.15F/100 ft) in the western uplands to 27C/
km (1.5F/100 ft) in the lowlands to the east. This trend
suggests control of the paleogeothermal gradient by the
paleohydrodynamics. Overlaid on the paleogeothermal
gradient pattern is the inferred drainage from the inset
map in Figure 22. Most major river valleys exhibit reentrants of the isogeothermal gradient contours along their
valleys, with the converse true for the drainage divides.
This pattern fits perfectly the schematic model of the asso-

Figure 22Map of early Eocene surface and ancestral drainage


pattern as determined by text calculation method. Contours in Figure 23Isopach map of calculated thickness of eroded sedimeters.
ments (m).

Brian Hitchon
elation of topography (water table), hydrodynamics, and
temperature regime (Figure 9). It is concluded, therefore,
that the early Eocene paleogeothermal regime was controlled by the paleohydrodynamic regime at that time, and
that the enhanced topographic relief, compared to the
present, resulted in a relatively subdued paleogeothermal
gradient pattern.
Late Paleocene Paleogeothermal Gradient
Continental sedimentation continued without interruption from the Late Cretaceous into the Paleocene in most
of Alberta. Clastic debris from the residual highlands of
the Cordilleran geanticline spread eastward to feather out
in Manitoba. Minor amounts of coal accumulated in
swamps associated with lakes and deltas. This description
(from Tiylor et al, 1964) suggests that the surface on which
the Paleocene sediments were deposited was one of low
reUef, and that the minor Laramide orogenic pulse in the

739

Canadian Cordillera was accompanied by epeirogenic


downwarping of Alberta to accommodate the thick Paleocene sequence. Therefore, the situation at the end of the
Paleocene, but before the second Laramide orogenic
pulse, must have been similar to that in the cross sections
in Figure 25 (the figure has been adjusted palinspastically
for 130 km, 80 mi, of tectonic compression caused by the
major Laramide thrusting). The depocenters for Paleozoic and Mesozoic sequences are west of that for the Tertiary, in accordance with the isopach maps in Figure 20.
Both the coal rank and coalification gradient (Hacquebard, 1977; Hacquebard and Donaldson, 1974)
increase gradually from the eastern Alberta Plains to the
edge of the Foothills Beh; westward of the belt both the
coal rank and gradient increase much more rapidly. TWo
situations may give rise to this trend (Figure 26). In a uniformly dipping stratigraphic unit (coal seam), the trend
may be produced by a logarithmic increase of R^ with
depth (section A-B). However, a linear increase of R^ with
depth (section A' -B') also will produce such a trend for a
gradually steepening unit. The palinspastic reconstruction
of the cross sections in Figure 25 suggests that the former
situation is applicable to Alberta (the isoreflectance lines
across southern Alberta at late Paleocene time are shown
in the bottom cross section). In the eastern Alberta Plains
the maximum R at the base of the Phanerozoic section is
about 0.5, whereas in the western Alberta Plains it is over
3.0 in a section approximately twice as thick. It appears
evident, therefore, that the geothermal gradient during the
late Paleocene was opposite to that in the early Eocene,
although the range of values was similar, based on the relationship of coal rank and paleotemperature (Hacquebard,
1977, Figures 3-7). For the eastern Alberta Plains the gradient was about 23''C/km (1.25F/100 ft), and for the
western Alberta Plains it was about 30C/km (I.65F/
100 ft). Thus the main effect of the hydrodynamic regime
initiated by the second Laramide orogenic pulse was to
decrease the geothermal gradient in western Alberta by
about 9C/km (0.5F/100 ft) and increase it in eastern
Alberta by about 4C/km (0.2''F/100 ft). The 5C/km
(0.25 "F/100 ft) difference between these values probably
reflects delayed readjustment of the hydrodynamic system. As erosion proceeded during the Tertiary, the entire
Alberta basin warmed up until the present temperature
regime was attained.
HYDROCARBON GENERATION HISTORY, ALBERTA

Palog60thermal
griKllcnt {'CAm)

As previously noted, early in the compaction history of


any sedimentary basin, before subaerial topographic relief
has developed, fluid flow will be upward and updip in the
geologic section. This suggests that throughout most of
21
Alberta for much of the Phanerozoic, flow has been
Limit of data

Inferred drainage divide


updip, from west to east. It was also noted that if the ther^ ^ Inferred drainage system
mal maturation conditions are appropriate for hydrocarbon generation, hydrocarbon accumulations associated
too km
with the ascending limbs and quasi-stagnant zones of flow
Figniw 24Regional paleogeothermal gradient pattern in early systems should be observable in both compacted and
Eocene al time of maximum sediment burial and probably maxi- undercompacted sedimentary basins. In both situations
the relative timing of hydrocarbon generation, trap formum topographic relief.
27

26

26
24
23
22

740

Geothermal Gradients, Hydrodynamics, and Hydrocarbons, Alberta

* Mannville coals
studied by
Hacquebard (1977)
5* Kootenay coal
studlad by
Hacquebard and ^
Donaldson (1974) ^

C
Initial
'
Laramlde
uplift

B'B
I
ROCKY
MTNS.

FOOTHLLS WESTERN
BELT
ALBERTA
POMS

EASTBM
ALBERTA
PLANS

100 km
H
60 mi

coal rank [ t t g j ^o
coallflcation gradient iSM % Rg/IOOm

.130 km pallnspastlc^
adjustment

FOOTHLLS
BB.T

WESTERN
ALjerrA
PLANS

EASTERN
ALBERTA
PLANS

100 km

Figure 25Cross section through southern Alberta at late Paieocene time, showing coal rank and coallflcation gradients for early
Mesozok coals, and reconstructed Isoreflcctancc contours (bottom cross section). Coal rank and coaHfleatton gradient data refer to
study by Hacquebard (1977) on Lower Cretaceous Mamiville Formation coals, and by Hacquebard and Donaldson (1974) on Upper
Jurassic Kootenay Formation (Purbeckian-Portlandian stages) coals from Foothills Belt.

Brian Hitchon

mation, and development of an appropriate flow system is


crucial.
By means of burial history curves, Deroo et al (1977)
have shown that the initial shallow burial of Devonian
strata in Alberta was followed by a long period with little
increase in depth of burial. It was not until about the middle Cretaceous that a significant increase in depth of burial
occurred, with maximum depth of burial in the early
Eocene. These authors (p. 110-111) suggest that the principal phase of crude oil genesis did not occur until the Cretaceous and probably Late Cretaceous, with oil genesis
taking place in source rocks of Cretaceous age at approximately the same time or slightly later. In Figure 25 the position of the isoreflectance lines in relation to the
Tertiary-Mesozoic boundary would support their suggestion. Crude oil generation conditions are limited to the
deepest Tertiary rocks, the Mesozoic of western and central Alberta, and the Paleozoic of central Alberta. The
catagenic natural gas regime is confined to the deepest
Mesozoic strata and the Paleozoic of western Alberta;
deep Paleozoic rocks are beyond the deep dry gas regime.
Biogenic natural gas may be found in Tertiary rocks, and
the shallower Mesozoic and Paleozoic sections. Because
some crude oils (e.g., the oil sand deposits) are found
beyond these limits, long distance migration must be
invoked.
Deroo et al (1977) also observed a major consequence of
the Late Cretaceous timing of crude oil genesis: a group of
favorable traps were in place at the time of generation.
Because fluid flow has nearly always been updip, most of
the Alberta basin would have been in the regime of ascending flow systems or quasi-stagnant zones from the time of
the principal phase of crude oil genesis (Late Cretaceous)
until the early Eocene (perhaps 10 m.y.); at this time,
hydrodynamic readjustment to the newly uplifted Rocky
Mountains and Foothills Belt occurred. Also at this time,
descending limbs (unfavorable for hydrocarbon accumulation) developed in the region of the Foothills Belt,
though most of the Alberta basin remained in a favorable
hydrocarbon accumulation regime.
According to T6th (1978), breaching of the preCretaceous unconformity in northeastern Alberta, which
resulted in development of the low fluid-potential drain
identified by the writer, was a relatively recent event (< 4
m.yB.P.). Depending on the exact time of breaching (0.3
to 1.3 m.y) 85 to 21 % of the previously existing potential
differences still remain. If this is correct, the genetic association of many Devonian and Carboniferous hydrocarbon occurrences with the low fluid-potential drain may be
fortuitous. Their hydrocarbons have had from Late Cretaceous to perhaps Pliocene to move to their present positions, and not the mere 4 m.y. since breaching occurred. At
present it is not possible to determine the importance of
the time of breaching in relation to hydrocarbon accumulation within the low fluid-potential drain. However, for
at least the southeast Alberta Gas System, hydrodynamic
conditions appear to have been favorable since the early
Eocene.
Figure 27 is a cross section along the same line as that in
Figure 18, showing the early Eocene surface from Figure
22 and the present relative position of the Milk River sand-

741

A'

B'

Arbitrary
datum

rO.4
CD

'
IS
su

_c

'5

0.8
in
1.2 ^
^Az^^
1.6

-^
/

s^O-.

D.

Figure 26Schematic relation of attitude of siratigraphic unit


(coal seam) and increase of R with depth Oeft, logarithmic
increase; right, linear increase).

Early Eocene
topography (m)

n.

Recharge
Discharge

s
3.0
2.5
2.0
1.5
1.0 - Milk River SR
0.5

gas field

-0.5
-1.0
-1.5

Pe

-2.0

Figure 27Eariy Eocene hydrologic features across southeast


Alberta Gas System. Cross section along same line as in Figure
18.

742

Geothermal Gradients, Hydrodynamics, and Hydrocarbons, Alberta

stone and Precambrian basement from Figure 18. It is significant that the present position of this gas field is in a
major early Eocene discharge region, and therefore its
position may already have been determined as early as the
beginning of the Eocene.
The relation of the timing of hydrocarbon generation
and the development of the hydrodynamic system responsible for conditions favoring hydrocarbon accumulation
in Alberta appear to be almost coincident in the context of
geologic time. The former was probably Late Cretaceous,
and the latter may have extended from Late Cretaceous
(compaction flow regime) into at least the early Tertiary
when subaerial gravity-induced formation flow was operative. The extent to which the present topographic surface
is responsible for hydrocarbon accumulation through
gravity-induced cross-formational flow must remain
unknown because this process was operative as long ago as
early Eocene; it appears to have had a similar effect on the
geothermal gradient pattern at that time as does the
present topography on the present geothermal gradient
pattern. It is certain, however, that gravity-induced crossformational flow does control the hydrodynamic regime,
which in turn influences the geothermal gradient pattern
and the accumulation of hydrocarbons.
CONCLUSIONS
1. A genetic link exists between the topography (watertable elevation), hydrodynamic regime, and geothermal
gradient pattern in Alberta.
2. On a regional scale, high topographic areas, such as
the Foothills Belt, have high water-table elevations with
correspondingly high potentiometric surfaces. These areas
control the regional recharge of cold meteoric water and
therefore have low geothermal gradients (< 20C/km,
I. r F/100 ft). Areas of medium elevation exhibit regional
lateral flow and intermediate geothermal gradients (2040''C/km, l.r-2.2F/I00 ft). Regional topographic
lows, such as northeastern Alberta, correspond with low
potentiometric surfaces, and the regional discharge of
warm formation waters from deep in the Alberta basin
results in high geothermal gradients (40 to > 50C/km,
2.2to>2.75F/I00ft).
3. Local topographic features cause perturbations in
this picture. River valleys result in drawndown potentiometric surfaces and reentrants on the regional geothermal
isogradient contours, sometimes to the extent that high
geothermal gradient anomalies are created along the
length of the valley. Contiguous local uplands may resuh
in closed low geothermal gradient anomalies along their
length, analogous to the effects of river valleys. Closed
drainage basins have closed relatively low potentiometric
surfaces and closed high geothermal gradient contours.
Isolated local uplands show local recharge features and
locally lower geothermal gradients.
4. From the above relations a model is developed that is
applicable to any compacted sedimentary basin with subaerial relief.
5. Using coal rank data, an early Eocene surface is
reconstructed and compared to the paleogeothermal gradient pattern at that time. The western upland areas had

paleogeothermal gradients as low as 21C/km (1.15F/


100 ft), and the eastern lowlands had gradients of 27C/
km (1.5 F/100 ft). The inferred drainage on this surface is
similar to the late Tertiary or early Pleistocene drainage of
Alberta. This similarity suggests that the ancestral drainage of Alberta was in existence in the early Eocene and the
drainage pattern did not change in any fundamental feature for more than 50 m.y. The fine detail of the inferred
early Eocene drainage pattern and the paleogeothermal
gradient pattern conform to the model, and it is concluded
that the early Eocene paleogeothermal regime was controlled by the paleohydrodynamic regime at that time. The
enhanced topographic relief, compared to the present,
resulted in a relatively subdued paleogeothermal gradient
pattern.
6. Using the same coal rank data and making palinspastic adjustment for the tectonic compression caused by
the major Laramide thrusting, it is shown that the predeformation isoreflectance lines show a logarithmic increase
with depth. Therefore, in the late Paleocene, the paleogeothermal gradient in the eastern Alberta Plains was about
23C/km (1.25F/100 ft), compared to about 30C/km
(1.65F/100 ft) in the western Alberta Plains; this is a
regional trend opposite to the post-Laramide trends.
Thus, the second Laramide erogenic pulse initiated a
topographically controlled hydrodynamic regime which
decreased the geothermal gradient in western Alberta by
about 9C/km (0.5F/100 ft) and increased it in eastern
Alberta by about 4C/km (0.2F/100 ft). As erosion proceeded during the Tertiary, the entire Alberta basin
warmed up until the present temperature regime was
attained.
7. With respect to the hydrodynamic and temperature
regimes, detailed analysis of the position of the oil sand
deposits, the Upper Devonian hydrocarbon accumulations in the low fluid potential drain of central Alberta,
and the southeast Alberta Gas System, shows that in every
case there is a genetic link between the hydrocarbon occurrence and the present fluid regime. The extent to which the
present topographic surface is responsible for hydrocarbon accumulation through gravity-induced crossformational flow remains unknown, because this process
was operative as long ago as early Eocene. Furthermore, it
appears to have had a similar effect on the geothermal gradient pattern at that time as does the present topography
on the present geothermal gradient pattern. For at least
the southeast Alberta Gas System, its position may
already have been determined as early as the beginning of
the Eocene.
REFERENCES
Barton, R. H., E. A. Christiansen, W. O. Kupsch, W. H. Mathews, C. P.
Gravenor, and L. A. Bayrock, 1964, Quaternary, in R. G. McCrossan
and R. P. Claister, eds., Geological history of western Canada:
Alberta Sodety of Petroleum Geologists, p. 195-200.
Canadian Society of Petroleum Geologists, 1975, Geological highway
map of Alberta. Scale 7 in. = 2.5 mi.
Deroo, G., T. G. Powell, B. Tissot, and R. G. McCrossan, 1977, The origin and migration of petroleum in the western Canadian sedimentary
basin. Albertaa geochemical and thermal maturation study: Geological Survey of Canada Bulletin 262,136 p.
Energy Resources Conservation Board, 1980, Alberta's reserves of crude

Brian Hitchon
oil, gas, natural gas liquids, and sulphur at 31 December 1979: Calgary, Alberta, Canada, Energy Resources Conservation Board, 504 p.
Freeze, R. A., 1966, Theoretical analysis of regional groundwater flow:
PhD thesis. University of California, Berkeley, 304 p.
1%9, Theoretical analysis of regional groundwater flow: Canada
Department of Energy, Mines, and Resources, Inland Waters Branch,
Science Series 3,147 p.
and P. A. Witherspoon, 1966, Theoretical analysis of regional
groundwater flow1, analytical and numerical solutions to the
mathematical model: Water Resources Research, v. 2, p. 641-656.
1967, Theoretical analysis of regional groundwater
flow2, effect of water-table configuration on subsurface permeability variation: Water Resources Research, v. 3, p. 623-634.
1968, Theoretical analysis of regional groundwater
flow3, quantitative interpretations: Water Resources Research, v.
4, p. 581-590.
Fuex, A. N., 1977, The use of stable carbon isotopes in hydrocarbon
exploration: Journal of Geochemical Exploration, v. 7, p. 155-188.
Hackbarth, D. A., 1978a, Regional hydrogeology of the Athabasca oil
sands area, Alberta, in D. A. Redford and A. G, Winestock, eds.,
The oil sands of Canada-Venezuela 1977: Canadian Institute of Mining and Metallurgy Special Volume 17, p. 87-102.
1978b, Groundwater temperature in the Athabasca oil sands
area. Alberta: Canadian Journal of Earth Sciences, v. 15, p. 16891700.
Hacquebard, P. A,, 1977, Rank of coal as an index of organic metamorphism for oil and gas in Alberta, in G. Deroo et al. The origin and
migration of petroleum in the western Canadian sedimentary basin.
Alberta: Geological Survey of Canada Bulletin 262, p. 11-22.
and J. R. Donaldson, 1974, Rank studies of coals in the Rocky
Mountains and inner Foothills Belt Canada, in R. R. Dutcher, P. A.
Hacquebard, J. M. Schopf, and J. A. Simon, eds.. Carbonaceous
Materials as Indicators of Metamorphism: GSA Special Paper 153, p.
75-94.
Hitchon, B., 1969a, Fluid flow in the western Canada sedimentary basin,
I. effect of topography: Water Resources Research, v. 5, p. 186-195.
1969b, Fluid flow in the western Canada sedimentary basin, 2.
effect of geology: Water Resources Research, v. 5, p. 460-469.
1971, Origin of oil: geological and geochemical constraints, in R.
F. Gould, ed.. Origin and refining of petroleum: American Chemical
Society Advances in Chemistry Series, v. 103, p. 30-66.
1974, Application of geochemistry to the search for crude oil and
natural gas, in A. A. Levinson, ed.. Introduction to exploration geochemistry: Calgary, Canada, Applied Publishing Ltd., p. 509-545.
and I. Friedman, 1969, Geochemistry and origin of formation
waters in the western Canada sedimentary basin, 1. Stable isotopes of
hydrogen and oxygen: Geochimica et Cosmochimica Acta, v. 33, p.
1321-1349.
A. A. Levinson and S. W. Reeder, 1%9, Regional variations of
river water composition resulting from halite solution, Mackenzie
River drainage basin, Canada: Water Resources Research, v. 5, p.
1395-1403.
Hubbert, M. K., 1940, The theory of groundwater motion: Journal of
Geology, v. 48, p. 785-944.
Klemme, H. D., 1975, Geothermal gradients, heat flow and hydrocarbon
recovery, in A. G. Fischer and S. Judson, eds., Petroleum and global
tectonics: Princeton, New Jersey, Princeton University Press, p. 251304.
Kumar, N., 1979, Thickness of removed sedimentary rocks, paleopore

743

pressure, and paleotemperature, southwestern part of western Canada Basin: discussion: AAPG Bulletin, v. 63, p. 812-815.
Magara, K., 1976, Thickness of removed sedimentary rocks, paleopore
pressure, and paleotemperature, southwestern part of western Canada basin: AAPG Bulletin, v. 60, p. 554-565.
Majorowicz, J. A., and A. M. Jessop, 1981, Regional heat flow patterns
in the western Canadian sedimentary basin: Tectonophysics, v. 74, p.
209-238.
McCrossan, R. G., and J. W. Porter, 1973, The geology and petroleum
potential of the Canadian sedimentary basinsa synthesis, in R. G.
McCrossan, ed.. The future petroleum provinces of Canadatheir
geology and potential: Canadian Society of Petroleum Geologists
Memoir 1, p. 589-720.
Outtrim, C. P., and R. G. Evans, 1978, Alberta's oil sands reserves and
their evaluation, in D. A. Redford and A. G. Winestock, eds.. The oil
sands of Canada-Venezuela 1977: Canadian Institute of Mining and
Metallurgy Special Volume 17, p. 36-66.
Porter, J. W., R. A. Price, andR. G. McCrossan, 1982, The western Canada sedimentary basin: Philosophical Transitions of the Royal Society of London, Series A, v. 305, p. 169-192.
Rice, D. D., and G. E. Claypool, 1981, Generation, accumulation, and
resource potential of biogenic gas: AAPG Bulletin, v. 65, p. 5-25.
Roberts, W. H., Ill, 1979, Some uses of temperature data in petroleum
exploration: paper presented to Symposium II, Unconventional
Methods in Exploration for Petroleum and Natural Gas, Dallas,
Texas, September 13-14,1979.
Schwartz, F W., K. Muehlenbachs, and D. W. Chorley, 1981, Flowsystem controls of the chemical evolution of groundwater; Journal of
Hydrology, v. 54, p. 225-243.
Stevenson, D. R., and D. M. Borneuf, 1977, Hydrogeology of the Medicine Hat area. Alberta: Alberta Research Council Report 75-2,11 p.
Taylor, R. S., W. H. Mathews, and W. O. Kupsch, 1964, Tertiary, in R. G.
McCrossan and R. P. Glaister, eds.. Geological history of western
Canada: Alberta Society of Petroleum Geologists, p. 190-194.
T6th, J., 1%2, A theory of groundwater motion in small basins in central
Alberta, Canada: Journal of Geophysical Research, v. 67, p. 43754387.
1963, A theoretical analysis of groundwater flow in small drainage basins: Journal of Geophysical Research, v. 68,4795-4812.
1970, Relations between electric analogue patterns of groimdwater flow and accumulation of hydrocarbons: Canadian Journal of
Earth Sciences, v. 7, p. 988-1007.
1978, Gravity-induced cross-formational flow of formation fluids. Red Earth region. Alberta, Canada: Analysis, patterns, evolution: Water Resources Research, v. 14, no. 5, p. 805-843.
1979, Patterns of dynamic pressure increment of formation-fluid
flow in large drainage basins, exemplified by the Red Earth region.
Alberta, Canada: Bulletin of Canadian Petroleum Geology, v. 27, p.
63-83.
1980, Cross-formational gravity-flow of groundwater: a mechanism of the transport and accumulation of petroleum (the generalized
hydraulic theory of petroleum migration), in W. H. Roberts, III, and
R. J. Cordell, eds.. Problems of Petroleum Migration: AAPG Studies in Geology 10, p. 121-167.
U.S. Geological Survey and AAPG, 1976, Geothermal gradient map of
North America: U. S. Geological Survey, Scale 1:500,000.2 sheets.
van Everdingen, R. 0., 1972, Thermal and mineral springs in the southern Rocky Mountains of Canada: Ottawa, Canada, Environment
Canada, 151 p.

Anda mungkin juga menyukai