You are on page 1of 250

Geolhermics in Basin Analysis

COMPUTER APPLICATIONS IN THE EARTH SCIENCES


Aseries edited by Daniel F. Merriam

1969 - Computer Applications in the Earth Sciences


1970 - Geostatistics
1972 - Mathematical Models of Sedimentary Processes
1981 - Computer Applications in the Earth Sciences: An Update of the 70s
1988 - Current Trends in Geomathematics
1992 - Use of Microcomputers in Geology
1993 - Computerized Basin Analysis: The Prognosis of Energy and Mineral Resources
1996 - Geologic Modeling and Mapping
1999 - Geothermics in Basin Analysis
Geothermics in Basin Analysis

Edited by
Andrea Forster
GeoForsehungslentrum Potsdom
Potsdom, Germony

and
Daniel F. Merriam
Konsos Geologicol Survey
University of Konsos
Lowrenee, Konsos

Springer Science+Business Media, LLC


Proceedings ol the American Association ol Petroleum Geologists/SEPM (Society ol
Sedimentary Geologists), held May 20, 1996, and the Canadian Society ol Petroleum
Geologists/SEPM (Society ol Sedimentary Geologists), he Id June 4,1997
ISBN 978-1-4613-7154-0 ISBN 978-1-4615-4751-8 (eBook)
DOI 10.1007/978-1-4615-4751-8
© 1999 Springer Science+Business Media New York
Originally published by Kluwer Academic I Plenum Publishers in 1999
Softeover reprint of the hardcover 1st edition 1999

All rights reserved


No part ol this book may be reproduced, stored in a retrieval system, or transmitted in any lorm
or by any means, electronic, mechanical, photocopying, microlilming, recording, or otherwise,
without written permission Irom the Publisher
PREFACE

Most of the papers presented here are an outgrowth of sessions on geothermics at two
meetings - the American Association of Petroleum Geologists/SEPM (Society of Sedimentary
Geologists) in San Diego, California (May 1996) and the Canadian Society of Petroleum
Geologists/SEPM (Society of Sedimentary Geologists) in Calgary, Canada (June 1997). In
the geological community there is an increased interest in geothermics as applied in
sedimentary basins. The focus is to analyze the thermal state of the sedimentary sequence and
the impact of heat flow from the basement as well as the processes that lead to the deposition
and alteration of the sediments and their mineral resources. Thus, increased interest also
centers in the sophistication in modeling organic matter maturation and petroleum generation.
Two good examples of this interest are an earlier publication by the Norwegian Petroleum
Society of the proceedings of a conference on Basin Modelling: Advances and Applications
(1993), and the recently released book by Colin Barker, Thermal Modeling of Petroleum
Generation: Theory and Applications (1996). With this book we offer now another
contribution to the study of sedimentary basins stressing essential parts of problems in which
geothermics is engaged.
The papers cover a wide variety of topics and show just how varied and diverse the
subject of geothermics is and how many aspects of the geosciences it permeates. Subject
matter includes the measuring oftemperature logs and capturing of industrial temperature data
and their interpretation to delineate subsurface conditions and processes, the importance of
porosity and pore filling for modeling thermal fields, the thermal insulation of shales,
geothermal anomalies associated with mud diapirs, basin hydrodynamic regimes, temperatures
related to magmatic underplating and plate tectonics. Geographically the basins covered
include the Taranaki Basin in New Zealand, Alberta Basin in Canada, South Caspian Basin
in Azerbaijan, and in the U.S., the Michigan Basin in Michigan and adjacent areas, and
Cherokee Basin in Kansas. We believe that the wide and diverse coverage will give the
beginner a good place to start with geothermics, the practitioner many things to think about
and use, and the casual reader a good idea of the subject.
We would like to thank those who reviewed the papers and helped improve this
contribution. We have relied heavily on our respective colleagues: Phil Armstrong (University
of Utah), Stefen Bachu (Alberta Geological Survey), Ulf Bayer (GeoForschungsZentrum
Potsdam), Graeme Beardsmore (Southem Methodist University), David Blackwell (Southern
Methodist University), David Deming (University of Oklahoma), John Doveton (Kansas
Geological Survey), Peter Gretener (University of Alberta), Peer Hoth
(GeoForschungsZentrum Potsdam), Ernst Huenges (GeoForschungsZentrum Potsdam), Al
Macfarlane (Kansas Geological Survey), David Newell (Kansas Geological Survey), JeffNunn
(Louisiana State University), Henry Pollack (University of Michigan), John Sass (U.S.

v
VI PREFACE

Geological Survey), Marios Sophocleous (Kansas Geological Survey), Jom Springer


(GeoForschungsZentrum Potsdam), and Lynn Watney (Kansas Geological Survey).
We would like to thank LeaAnn Davidson of the Kansas Geological Survey for help
in the final preparation of manuscripts for publication. Her expertise in manuscript processing
was much appreciated. Janice Sorensen ofthe Kansas Geological Survey helped locate hard-
to-find references; Cora Cowan, also of the Geological Survey assisted with preparation of the
Index. All of the authors are to be thanked for their papers and in helping to make this a
notable contribution to the subject.

Nice, France Andrea Forster


Dan Merriam
CONTENTS

Introduction, by D.D. Blackwell .......................................................................................... .ix

High-resolution temperature logs in a petroleum setting:


examples and applications, by D.D. Blackwell,
G.R Beardsmore, RK. Nishimori, and RJ. McMullen, Jr ....................................................l

Problems and potential of industrial temperature data from


a cratonic basin environment, by A. Forster and D.F. Merriam.......................................... .35

Present heat flow along a profile across the Western Canada


Sedimentary Basin: the extent of hydrodynamic influence,
by J.A. Majorowicz, G. Garven, A. Jessop, and C. Jessop ................................................... 61

Regional-scale geothermal and hydrodynamic regimes in the


Alberta Basin: a synthesis, by S. Bachu............................................................................... 81

Basin-scale groundwater flow and advective heat flow: an


example from the northern Great Plains, by W.D. Gosnold, Jr................ ,.......................... 99

Thermal insulation by low thermal conductivity shales:


implications for basin-scale fluid flow and heat transport
by J .A. Nunn, G. Lin, and L. Zhang................................................................................... 117

Temperature and maturity effects of magmatic underplating


in the Gjallar Ridge, Norwegian Sea, by W. Fjeldskaar,
H. Johansen, T. A. Dodd, andM. Thompson..................................................................... 131

Combining tectonics and thermal fields in Taranaki Basin,


New Zealand, by P.A. Armstrong and D.S. Chapman ..................................................... 151

Thermal history of a deep well in the Michigan Basin:


implications for a complex burial history, by W.D. Everham
and J.E. Huntoon............................................................................................................... .177

Rising mud diapirs and their thermal anomalies, by


E. Bagirov and I. Lerche ................................................................................................... .203

vii
Vlll CONTENTS

Effect of oil and gas saturation on simulation of temperature


history and maturation, by H.S. Poelchau, C. Zwach,
Th. Hantschel, and D.H. Welte ........................................................................................... 219

Contributors ........................................................................................................................ 237

Index ................................................................................................................................... 239


INTRODUCTION

David D. Blackwell

Department of Geological Sciences


Southern Methodist University, Dallas, Texas

The motivation for the technical sessions that led to this book is that geothermal
conditions play an important role in basin-scale processes, including subsidence,
sedimentation, compaction, and diagenesis, and affect the alteration of organic matter and the
generation ofhydrocaroons. In all of the integrated basin modeling computer programs used
for hydrocarbon assessment, heat flow is one of the major input parameters and constraints.
Research in the past few years, however, suggests that our views on measuring and interpreting
terrestrial heat-flow density in sedimentary basin settings should be reevaluated. Special
efforts need to be made to evaluate the errors in conventional approaches to the measurement
of thermal conditions and to the study of the heat transfer in the subsurface particularly from
the point of view of separating conductive heat flow from an advective component. Clarifying
the nature of the observed heat flow is essential in characterizing sedimentary basins in terms
of their tectonophysical state and their reservoir properties.
There are many comprehensive and elegant numerical simulators available for use in
the analysis of basin thermal structure and organic maturation history. All of these programs,
however sophisticated, depend on the quality of the input data for useful output calculations.
At the present time the weak link in the process is the quality of this input and realistic
evaluations of error. There is much to be done in-the form of well constrained case histories,
laboratory and field measurements of inputs, and error analysis before the input is up to the
standards of the modeling capabilities. Thus as stated in the call for papers the discussions in
this volume particularly address two objectives: (1) basic studies of the evaluation and
interpretation of heat flow and thermal structure of sedimentary basins; and (2) examples of
modeling of modern and paleogeothermal conditions.
There are not many volumes that focus on the thermal aspects of sedimentary basins.
Some of the most basic facts about the thermal field in many sedimentary basins remain
unknown or are in question. Even the dominant mode of heat transfer, conduction or
convection, is uncertain in many areas. In part this lack of understanding is related to the
limited applicability in the exploration setting of the methods of classical heat-flow
determination (Haenel, Rybach, and Stegena, 1988) because of the nature of the thermal data
that the explorationist has to work with, that is BHT's and cuttings, with their lack of detail,

lX
x BLACKWELL

resolving power, and error evaluation capability. Some of these limitations are discussed and
ways around some of them are suggested in this volume.
So aside from the introductory material contained in the text books by Gretener (1981),
Lerche (1990), and Jessop (1990) those interested in the subject of geothermics of sedimentary
basins have to build, in general, a background from the literature. Earlier volumes that contain
papers with a similar theme were edited by Durand (1984), by Burris (1986), and by N aeser
and McCulloh (1989). As is the situation in most areas of hydrocarbon exploration the
advances in the field have been hampered by the financial state ofthe hydrocarbon exploration
industry and there is a gap in time in books concerned with this subject. At the same time the
importance of understanding the history of hydrocarbon maturation and movement is
becoming even more clear (Shirley, 1998). Thus this volume is timely.
The papers in this volume focus on the temperature and heat-flow field but none of the
papers address the paleothermal indicators such as apatite fission track dating, diagenetic
reactions, details of organic maturation and so forth. However, this aspect ofthe field has been
summarized recently in detail by Barker (1996). The papers included in this book can be
divided into several groups. The first chapter deals with the collection of detailed temperature
data in sedimentary basins. Several of the papers point out limitations of the BHT data sets
that must be used at this time. Blackwell, Beardsmore, Nishimori, and McMullen discuss why
developments in technology allow the routine collection of a new class of thermal data logs
in petroleum exploration situations, that high-resolution, precision equilibrium temperature
logs. They give several examples of sets of equilibrium logs from hydrocarbon field studies
an demonstrate the detailed information that is available from these types oflogs.
The largest group of papers are concerned with the partitioning of heat transfer into
convective and conductive components. The amount of heat that can be carried in long
sedimentary aquifers have been the subject of much discussion in the literature. In particular
the Prairies Basin in Alberta, Canada has been the site of vastly different interpretations of the
thermal effect of regional fluid flow. The papers in this volume by Bachu, and by Majorowicz,
Garven, Jessop, and Jessop seem to be approaching a consensus that the scale of the aquifer
is too great flow for regional heat transfer from one end to the other. On the other hand
Gosnold shows an example of the length scale where advection can become important, the
Dakota aquifer system in the north-central plains region of the US. The differences in
interpretation have been driven in part by the chronic problems in the BHT data interpretation
and lack of good information on in situ thermal conductivity values.
The paper by Fjeldskaar, Johansen, Dood, and Thompson is concerned with a basin in
the Norwegian Sea near the continental margin. They use crustal structure determinations to
constrain stretching ratios and estimate the amount of underplating. The paper represents
examples of the type of thought that goes into preliminary exploration decisions along
continental margins where tectonics and the details of continental extension can be critical in
determining the thermal history of a package of sediments.
Particularly complicated but lacking in detailed case studies are thermal regimes in
basins in areas of active tectonism. The paper by Armstrong and Chapman is concerned with
an such an area, the Taranaki Basin in New Zealand. In this active volcanic arclback arc
setting the heat flow probably is varying as a function of time and space so careful
determination of the basic thermal parameters is necessary to work out the present thermal
regime so that the paleothermal conditions can be more closely measured.
Evenham and Huntoon describe a detailed thermal history study of a well in the
cratonic Michigan Basin of the United States. The simple cratonic basins in the United States
INTRODUCTION Xl

present difficulties in the application of thermal history models and this paper dissects the case
for one well in that basin with a lot of thermal maturity data.
The paper by Nunn and others falls into several of the categories. It is a theoretical
study that investigates the effect oflarge-scale fluid flow using the Arkoma foreland basin as
an example. So the paper is a theoretical example of the problem of conductive versus
advective heat transfer. The thermal effects of this fluid flow depend significantly on the
thermal conductivity of the sediment section. Thus even in this theoretical example the authors
must point out that an accurate knowledge of the thermal parameters is necessary.
The thermal effect of salt domes has received some attention but that affect on the
thermal regime is discussed only briefly in the literature; the effect of mud diapirs is one of
these. Although occurring in the Gulf Coast for example, they are little described in the
literature. The paper by Bagirov and Lerche describes the Abikh dome in the South Caspian
Basin and presents a theoretical model for the effect of the growth of the mud dome on the
thermal field around it.
The direct detection of hydrocarbons with thermal techniques has been discussed in the
literature although no consensus has been reached. This subject is related to the use of
geothermics in basin analysis in a couple of different ways that are discussed in this volume.
The fact that this issue has not been resolved yet indicates the distance we have to go in fully
understanding the thermal field in sedimentary rocks. Forster and Merriam show results from
an area in eastern Kansas where significant temperature anomalies seem to be associated with
some anticlines. Forster, Merriam, and Davis (1997) illustrate details of several such
anomalies. These types of anomalies might be the result of advection effects or lateral changes
in thermal conductivity. Lateral changes in thermal conductivity could be the result of the
effect of hydrocarbon in the pore space in and above a reservoir. This effect of the change in
thermal conductivity because of the saturation with hydrocarbons is discussed by Poe1chau,
Zwach, Hantschel, and Welte as are the implications of this effect. As do the authors of
several other papers in this volume Poe1chau, Zwach, Hantschel, and Welte emphasize that
simply using default values for thermal conductivity in thermal history numerical modeling
programs can be a mistake.
Forster and Merriam also discuss some problems with using BHT data, particularly in
a sedimentary environment of thin cover over a Precambrian basement. This discussion
corroborates the conclusions of Majorowicz, Garven, Jessop, and Jessop that shallow BHT
values (from wells less than 500 m deep) are particularly prone to large errors. Forster and
Merriam also point out the impossibility of measuring formation gradients with BHT data (as
opposed to averages for a large depth range, 100's of meters) which may be close to the actual
mean gradient. Thus interval thermal conductivity cannot be measured and errors in thermal
conductivity association with lithology cannot be recognized.
Thus a wide variety oftopics related to Geothermics in Basin Analysis are covered by
the papers in this volume. The unified conclusion of all of the papers seems to be that, to reach
the type of precision needed to use thermal data for maturation modeling and exploration
decisions, fundamental understanding of thermal regimes, and the factors controlling them is
vital. And in spite of the progress during the last 20 years there are some basic questions that
need resolution and additional study. Furthermore there are large areas ofthe globe for which
accurate thermal analyses do not exist. Thus the continuation of studies of the sort described
in this book are vital to the increasing accuracy of hydrocarbon exploration.
Xll INTRODUCTION

REFERENCES

Barker, C., 1996, Thermal modeling of petroleum generation: theory and application: Developments in
Petroleum Science 45, Elsevter, Amsterdam, 512 p.
Burris, J., ed., 1986, Thermal modeling in sedimentary basins: Colloques et Seminaires 44, Editions Technip,
Paris, 600 p.
Durand, B., ed., 1984, Thermal phenomena in sedimentary basins: Colloques et Seminaires 41, Editions
Technip, Paris, 325 p.
Forster, A., Merriam, D. F. and Davis, J. c., 1997, Spatial analysis of temperature (BHTIDST) data and
consequences for heat flow determination in sedimentary basins: Geol. Rundschau, v. 86, no. 2, p.
252-261.
Gretener, P. E., 1981, Geothermics, using temperature in hydrocarbon exploration: Continuing Education
Course Note Series 17, Am. Assoc. Petroleum Geologists, Tulsa, Oklahoma, 156 p.
Haenel, R., Rybach, 1., and Stegna, 1., eds., 1988, Handbook of terrestrial heat-flow density determination: D.
Reidel Publishing, Dordrecht, Holland, xx pp.
Jessop, A. M, 1990, Thermal geophysics: Developments in Solid Earth Geophysics 17: Elsevier, Amsterdam,
306p.
Lerche, I., 1990, Basin analysis, quantitative methods: Academic Press, San Diego, v. 1,510 p. and v. 2, 570
p.
Naeser, N. D., and McCulloh, T. H., eds., 1989, Thermal history of sedimentary basins: methods and case
histories: Springer-Verlag, New York, 319 p.
Shirley, K, 1998, Egypt desert an exploration oasis: new concepts expand productive area: Am. Assoc. Petroleum
Geologists Explorer, v. 19, no 8, 32-34.
HIGH-RESOLUTION TEMPERATURE LOGS IN A
PETROLEUM SETTING: EXAMPLES AND APPLICATIONS

David D. Blackwell l and Graeme R. Beardsmore l


Richard K. Nishimori2, and Richard J. McMullen, Jr.2

IDepartment of Geological Sciences


Southern Methodist University, Dallas, Texas
2Mobil Technology Company, Dallas, Texas

ABSTRACT

Examples of high-resolution temperature logs measured in oil and gas fields in the United
States are presented and the pertinent features useful in basin analysis are discussed. We point
out that wells suitable for equilibrium or near equilibrium temperature logs usually are
available, and we describe by examples criteria for the evaluation of the quality of a high-
resolution temperature log. Examples of temperature gradient logs from two fields in the
Paleozoic-age Anadarko Basin in Oklahoma, one field in the Cenozoic Gulf Coast Basin, and
one field in the Mesozoic/Cenozoic Sacramento Basin in northern California are described and
their application to the analysis of basin thermal structure discussed. The major criteria that
can be used to evaluate the quality of the log are level of (temperature) noise, presence/absence
of negative/zero gradient sections, degree of correlation with other geophysical logs, and well
to well comparisons. Even logs that are not in complete equilibrium contain significant
information compared to a typical set ofBHT points. The development of memory PIT tools
and their deployment in the field for production logging indicates that the potential now exists
for routine collection of high-resolution temperature data in hydrocarbon settings worldwide.
Because the thermal regime of many boreholes in producing fields may be closer to
equilibrium than has been thought in the past, the new temperature capability can be used in
a practical way. High-resolution logs can furnish detailed information on the gradient, the
ratios of the thermal conductivity values in hard-to-sample lithologies to those iIi lithologies
easier to characterize, and ultimately more precise understanding of the thermal regime
(whether conductive, convective, etc.) in individual wells and sedimentary basins. So the use
of these types of high-resolution logs can be added to the petroleum explorationist's data sets
to be used in basin thermal analysis as an important new source of information that can be used
to increase the precision or estimate the errors of the models developed based on conventional
BHT's and cuttings measurements.
2 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

INTRODUCTION

The generation of petroleum products from organic material is driven primarily by


temperature (e.g. Connan, 1974). Modeling thermal conditions in a sedimentary basin through
time therefore is a critical step in assessing the maturity of petroleum source beds, the timing
of petroleum generation, likely fluid migration paths, and reservoir locations. A crucial
constraint on past thermal conditions is the present-day temperature distribution within the
basin. An accurate determination of present-day thermal regime thus is a vital step in assessing
a hydrocarbon prospect.
Heat transport within a sedimentary setting may be by conduction (direct heat transfer
between adjacent grains) or advection (heat carried by moving fluid). In a thermally
equilibrated, purely conductive setting, where heat transfer by fluid movement within or
between sedimentary layers is negligible, vertical heat flow generally can be assumed to be
constant with depth and, together with thermal conductivity, will define the temperature
distribution. Therefore, present-day vertical heat flow is a vital parameter in assessing organic
maturity in prospective petroleum locations.
Theoretically, heat flow is a relatively simple value to calculate because one need only
measure the thermal gradient and thermal conductivity within the conducting strata and the
conductive heat flow is determined fully because Q = - KxdT/dz, where K is thermal
conductivity, dT/dz is the vertical thermal gradient, and Q is heat flow. Because both gradient
and thermal conductivity have equal importance to the final heat flow and in calculating basin
thermal regimes as a function of time, accurate thermal history determination requires accurate
measurements of both parameters. Historically, however, accurate measurements of these two
parameters have been difficult to obtain using typical exploration data sets. To this day, there
is controversy about the accuracy and precision of the typical techniques used in the petroleum
setting to measure both subsurface gradients and the in situ thermal conductivity of earth
materials.
In situ thermal conductivity, inherently, is a difficult quantity to determine. It can not be
measured directly, but has to be deduced from indirect thermal measurements of covarying
properties obtained from well logs or core and (more usual) cuttings samples. Coupled with
a limited understanding of the relationship between conductivity values measured in the
laboratory and in situ, it is inevitable that the uncertainty in many heat-flow measurements
might be expected to be associated primarily with the thermal conductivity component because
there are no such intrinsic difficulties in measuring thermal gradient. Relatively inexpensive
and simple technology allows temperature measurement to a precision of 0.001 °C, so thermal
gradients can be calculated to a ±0.5°CIkm precision at I-meter resolution. Such precision
effectively reduces uncertainty in the thermal gradient component of heat flow to I to 5% in
most instances. However, the dominant source of temperature data used in thermal analysis
of sedimentary basins is well-log header Bottom Hole Temperatures (BHT). With these BHT
data, thermal gradients, even on km scales, rarely can be determined to better than ±5-15%
(see Jessop, 1990 and the recent error discussion by Lee, Deming, and Chen, 1996), and thus
there is not sufficient resolution to discriminate significant lateral and vertical thermal
conductivity variations from heat-flow variations. Combined errors imply that the heat-flow
values determined using BHT and cuttings measurements have typical errors of at least ±20%.
Jessop (1990) has described one of the few comparison studies using a tightly constrained and
comparable sets of the two types of data. In his study multiple BHT points from wells in an
area 3Ox30 km2 surrounding sites of four wells with detailed temperature logs were compared.
He concluded that the average of the BHT was within 10°C of the actual temperature but
HIGH-RESOLUTION TEMPERATURE LOGS 3

showed that the BHT data had no vertical resolution of gradient because of the depth clustering
and scatter of the BHT points. He emphasize that because of the completely different
information content, both types of data are essential for understanding the thermal state of
sedimentary basins.
Given the ease of acquiring quality temperature data and the great advantages in
application compared to BHT data, with the added bonus that in situ thermal conductivity
ratios also can be obtained in most situations, it is difficult to understand why they are so rarely
sought. Use of such data should reduce heat-flow errors to on the order of 5-10% and thus
result in at least a 100% decrease in the error of present-day heat-flow measurement. Some
of the reasons include the lack of readily available commercial in-the-field temperature logging
equipment, the unsuitability of open-hole temperature logs made immediately. following
completion of a well, and a belief that equilibrium temperature logs can not be made in typical
petroleum well settings. The objective of this paper is to discuss all of these arguments in the
light of modern technological advances in equipment and in the understanding of the thermal
regime of wells. As part of this discussion the equipment, methodology, processing, and
interpretation of downhole precision temperature logs will be reviewed.
Many thermal problems in sedimentary basins have remained unsolved because of the
poor resolving power ofBHT data and the limited number of detailed equilibrium temperature
logs described in the petroleum exploration literature. For example, the only book dedicated
to using temperature in hydrocarbon exporation (Gretener, 1981) can only point to one
example temperature log from a well in sediments that illustrates the inverse relationship of
thermal conductivity to thermal gradient (Gretener's fig. 4.3-3 and fig. 4.6-4, and copies of
paper field log prints). In the heat-flow literature there are numerous examples of the behavior
of detailed temperature logs in sediments, but these typically are not available or known in the
basin analysis field. Thus questions such as the thermal effects of salt, refraction of basement
uplifts, direct detection of hydrocarbons (Forster, Merriam, and Davis, 1998; McGee, Meyer,
and Pringle, 1989; etc.), and the effect of fluid flow on the temperature field in sedimentary
basins (Deming and others, 1992; Majorowitz and others, 1998; Bodner and Sharp, 1988; etc.)
remain open to experimental study. However, because of recent changes in technology the
methods are at hand to make temperature logs in sedimentary basins everywhere on a routine
basis as was not possible in the past (e.g. Wisian and others, 1998). These recent tool
developments are part of the motivation for this discussion focused on the use of detailed,
accurate temperature logs.
The application and interpretation of high-precision thermal gradient logs is best
illustrated using real examples. For this purpose, a number of examples are described from
several different sedimentary basin settings show how precision temperature and gradient data
enhance our knowledge of the thermal regime in a well. It is important to understand how
thermal gradient logs relate to lithology and other well logs, how the geologist can be
confident that a temperature or gradient log represents equilibrium conditions, how we can
ultimately use a gradient log to make a best estimate of heat flow in a well, and how we can
extrapolate our knowledge to other nearby wells.
Several other examples of precision temperature logs and their correlation to lithology
and other log information were presented by Blackwell and Steele (1989a, 1989b). There are
other examples in the literature as well, although on a more limited basis (Demongodin and
others, 1991, for example). Other examples specific to Kansas (Blackwell and Steele, 1989b),
Nebraska (Gosnold, 1990), and the Anadarko Basin were described by Carter and others
(1998). Gallardo and Blackwell (1999) illustrated how the addition of a few accurate, detailed
temperature logs may be used to calibrate in situ thermal conductivity values in a sedimentary
4 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

basin. They showed that in a conductive setting those calibrations, together with lithological
analyses, allow predictions of basin temperatures as accurately as, but independent ofBHT
analyses. Thus the possibility of real error analysis of the conventional BHT-cuttings
techniques (e.g. Lee, Deming, and Chen, 1996) may be possible. Brigaud, Chapman, and
LeDouaran (1990) and Griffiths and others (1992) have described detailed systems for
calculating thermal conductivity from well logs, but they did not have any detailed in situ
thermal conductivity distributions to compare with the results of their predictions. Thus the
combination of the two data sets should allow a level understanding of present-day basin
thermal structure not heretofore obtained.
The focus in these cited papers was not on the thermal regime in individual wells vis a
vis evaluation of equilibrium conditions so the details of the well settings were not discussed
except that Carter and others (1998) do have brief discussions of individual well thermal
conditions. The particular objective of this paper is to illustrate how, in real hydrocarbon
settings, useful temperature data can be obtained, some ofthe characteristics of how thermal
data quality may be recognized, and how high-quality thermal data may be utilized in basin
thermal studies. In addition the examples we present here extend and further illustrate the
relation between thermal gradient and rock type. A location map of the wells referred to in this
paper and those described in detail by Blackwell and Steele (l989a, 1989b), Gosnold (1990),
and Carter and others (1998) is shown in Figure 1. The location of the Anadarko Basin,
referred to in two of the examples is shown also. Detailed temperature-depth logs have been
described for the different setting of sedimentation in an active tectonic region, coastal
California, that is for the Ventura Basin (De Rito and others, 1989) and the Santa Maria Basin
(Williams and others, 1994).

CANADA

MEXICO

Figure 1. Location map of wells with high-resolution temperature logs: fields with wells described in this paper
(crosses); locations from Blackwell and Steele (1989a, squares); locations from Blackwell and Steele (1989b,
dots) ; locations from Gosnold (1991, stars), and locations from Carter and others (1998, pluses).
HIGH-RESOLUTION TEMPERATURE LOGS 5

APPARATUS

Temperature can be measured only by direct methods. That is, to measure the
temperature of the Earth we must physically lower instruments down available holes. The
temperature at points away from the holes must be interpolated from known data. High-
precision instruments for measuring downhole temperature have been available for many
years. Most are electronic in nature, utilizing thermistor or platinum resistance sensors as tl).e
temperature sensitive component (Gretener, 1981; Blackwell and Spafford, 1987, for example).
Once calibrated, a simple resistance measurement is sufficient to determine temperature.
Platinum resistance thermometers are superior to thermistors in that they are accurate and
stable and have a nearly linear resistance-temperature response for a large temperature range.
Unfortunately, their resistance is small (25-50Q), so if analog wireline techniques
(measurements of voltage or current using a multiconductor electrical cable for connection to
the probe) are used heavy, low resistance cable is required to maintain accuracy.
The simple analog downhole tool contains a thermistor or platinum sensor in a probe that
descends the hole. Electrical contact is maintained with the surface and real-time thermistor
resistance is monitored using a digital multimeter. Such. systems are simple to design and
operate, but are limited by a need for four leads and high cablehead leakage resistance.
Temperature resolution of 0.001 °C precision is possible with careful design (Blackwell and
Spafford, 1987).
Commercial temperature logging tools typically convert resistance to frequency
downhole so that a single wire (with steel sheath) is sufficient for logging and cable head
leakage resistance can be lower with good results. A problem with these tools is that the
frequency cOtinting typically has been for too short a time interval to give the O.OOI°C
resolution needed for high-quality logs and 0.1 °C usually is the accepted resolution. Modem
electronics now allow a 0.001 °C temperature resolution with frequency tools if so designed.
A classical production tool is the Kuster or Amarada bomb mechanical
pressure/temperature tool run on a slick-line (a solid wire used for mechanical strength only).
At the present time downhole-computer slick-line pressure/temperature (PIT) tools are
replacing this production tool technology (Larimore, Goiggon, and Bayhn, 1997). Real-time
surface monitoring of downhole pressure and temperature is not possible with these computer
tools because they are self-contained with onboard battery, memory, and processing chips.
The electronics may be housed in a sealed Dewar flask for operation at high temperatures.
Computer tools generally use a platinum temperature sensor for greater temperature stability.
Recording is initialized at the surface and the tool and the surface computer are time
synchronized, then the tool is simply lowered down the hole and the depth recorded as a
function of time by the computer connected to a digital depth encoder at the surface. Probe
resistance is recorded automatically at preset time intervals, and when the tool is returned to
the surface, data are downloaded onto a PC where the temperature (and pressure) are correlated
by time with depth to generate a conventional property-depth log. Computer tools are more
expensive than wire-line tools, but are more versatile. Slick-line is less expensive than wire-
line, and can be pressure isolated more easily for logging high-pressure or flowing wells.
Also, computer tools generally are designed to withstand high temperature and pressure, so are
more suited to logging in deep, hot, pressurized, producing or other hostile environment wells.
The temperature sensor should be mounted as near to the leading tip of the probe as
possible, so as to minimize disturbance to the well fluids prior to temperature measurement.
The probe should be a rugged construction of brass (or similar high thermal conductivity
material) and sealed to keep borehole fluids away from electrical connections. One operational
6 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

problem is that a cage usually is used to protect the sensor tube during logging. During
openhole logging the cage typically gets plugged with mud and drill cuttings effectively
increasing the time constant of the tool from seconds to minutes and seriously degrading log
qUality.
An accurate depth log must accompany any temperature log. To this end, some sort of
odometer must be included with the logging system to record accurately the length ofline that
has been fed from the winch. For some deep or hot wells, corrections may have to be applied
to compensate for elastic extension and thennal expansion within the wire. With wireline
equipment depth can be recorded simultaneously with real-time temperature data and stored
on a PC or other memory device. Computer tools include an internal clock and record data as
time-temperature(-pressure-etc.) pairs. A time-depth log must be collected independently at
the surface so that subsequent processing can merge the two data sets to produce the desired
depth-temperature pairs. These tools are capable of collection of research quality temperature
logs if the tools are calibrated (Wisian and others, 1998). The range of equipment now
available from service companies (Larimore, Goiggon, and Bayhn, 1997) makes it possible to
collect precision temperature gradient data under any conditions and in any locations where
hydrocarbons are located.
A completely different type of temperature logging system has become available recently
for well logging. It is referred to as a Distributed optical fiber Temperature Sensing system
(DTS) and is based on the Raman effect of back-scattered laser light in an optical fiber. It has
several major advantages over other types of logging systems. The DTS system is able to
provide repeated, near instantaneous measurements of temperature along the full length of the
fiber without disturbing the surrounding bore fluid. This makes it ideal for studying transient
events (e.g. GroJ3wig, Hurtig, and KUhn, 1996; Sakaguchi and Matsushima, 1995). It currently
is of limited precision (0.1 0c) and depth resolution (0.25-1.Om) compared to other systems
(Forster and others, 1997; Wisian and others, 1998) but provides data unattainable by other
methods.

METHODOLOGY AND PROCESSING

Precision temperature logging will yield only highest resolution of true formation
temperature (and gradient) if survey procedures are planned carefully and followed. When
planning a temperature logging survey, several factors need to be considered. Is the hole in
thennal equilibrium? What temperature and depth resolution are required? Could there be
convection or production disturbances within the hole and how will these degrade the quality
of the log? Many of these questions cannot be answered with certainty and one of the
objectives of this paper is to present examples oflogging in actual field environments that
illustrate some of the effects, and lack thereof, that can degrade temperature log quality and
thus develop empirical data on the conditions necessary for obtaining high-quality temperature
data.

Ensuring Equilibration

Precision-temperature logging can not be conducted directly after drilling. This puts it
at odds with openhole logging techniques. In order to obtain meaningful temperature results
the well fluid must be in thermal equilibrium with the surrounding rock strata. For this to hold
true, the fluid must be allowed time to achieve thermal equilibrium. Any event that disturbs
HIGH-RESOLUTION TEMPERATURE LOGS 7

the well fluid column also causes a thermal disturbance. Such events include drilling,
production, and logging. The amount of time required for equilibration depends on the
magnitude of the disturbance and is difficult to quantify so empirical examples will be
presented. Thermal equilibrium is more likely to be at least approximately approached in pre-
or post-production logging situations where the well environment is more conducive to
collection of good temperature logs in any event.
Drilling always causes a great thermal disturbance. Continuous circulation of large
volumes of fluid through the well during the drilling process disturbs the equilibrium
temperature of the surrounding strata by an amount from which it can take months to recover
completely. The longer the drilling time, the greater the recovery time. Ideally, at least three
times the drilling time has been cited as the minimum time that should be allowed to pass
before logging a newly drilled well (Jaeger, 1961). However, Carter and others (1998)
reported temperature logs from wells in the Anadarko Basin, some of which had been logged
at a rest time approximately equal to the drilling duration with acceptable results. Of course
the drilling time is shorter in the bottom of the well so the temperatures in the deeper part of
a well will approach equilibrium faster in an absolute sense than the shallower part.
Production, or removal, of fluids from a well, also causes a thermal disturbance, but the
magnitude is not as great as for drilling. The amount of time required to reequilibrate depends
strongly on the construction and production history of the well. A typical production well is
cased with 15 cm steel pipe, cemented, and produces through 5-7 cm (2-3 in) diameter steel
tubing. This configuration acts similar to a heat exchanger (Ramey, 1962) and so gradients
are less disturbed than temperatures. If production rates are moderate, as they may be toward
the end of the life of a well, the thermal disturbance around the well will be small and not vary
much with depth, and thermal gradient equilibrium should be attained in the tube, except in
the immediate vicinity of the production zone(s) a relatively short time after production is
halted. If flow is through a larger tube, or at high rates, the disturbance will be greater and a
longer recovery time will be necessary. In general, though, logging can be carried out several
weeks to months after production has ceased. The well construction of cemented casing and
tubing also contributes to lower gradient noise by removing hole size variations and reducing
the effective hole diameter and thus convection noise (see next).
In most fields there are wells that have been shut in for considerable lengths of time for
various reasons and require no further equilibration period. To ensure completely static
conditions we may install a packer above the perforations in pressured wells. If the cable is
packed off at the surface during logging it is not clear that the in-hole plugs are helpful in
improving the resulting log quality.
The act oflogging a hole, itself, will disturb the well fluids with the motion of the probe.
Logging should be conducted ideally DOWN the hole to ensure that the temperature of
undisturbed fluid is measured. Most other logging procedures run UP from the bottom of the
hole. The disturbance by the probe is considerably less than that caused by drilling or
production, and generally a day should be sufficient for reequilibration. Logging upwards, or
immediately relogging a hole may give satisfactory results if fine detail is not required in the
log. Such logs generally are noisier than first-run down logs, but medium- and broad-scale
temperature trends are retained.

Optimal Logging Speed

Efficiency dictates that logging should be conducted at the maximum rate that willretum
the quality of data required. Logging speed is limited by two factors; the spatial resolution
8 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

required on the log, and the thermal lag effect of the probe. Logging speed translates directly
into data point separation if sensor output is recorded at specified time intervals, the usual
situation for computer tools. It then is simply a matter oflogging at a rate to return data at the
required depth interval. Electric line tools are designed to trigger at a regular depth intervals
(e.g.O.1m). From a practical point of view recording temperatures at 0.2 to 0.1 m (3 to 6 in)
is sufficient to obtain maximum thermal gradient resolution in a typical well situation.
With each finite distance the probe descends, the temperature changes by a finite amount.
It takes a finite time for this temperature change to propagate through the body of the probe
to the thermistor or platinum sensor. The time lag may be only on the order of seconds, but
if the descent rate is rapid, this lag translates into an effective depth offset on the final
temperature log and a loss of high-frequency variations. The exact value of the time lag
depends on the thermal bulk, or time constant, of the probe. In general, more robust, thick or
stainless steel probes have a higher thermal bulk and longer time lags than flimsy, thin or brass
probes. The environment within the hole also may conspire to increase the thermal bulk ofthe
probe as well. An uncased well typically will be muddy, and mud can cake in the protective
cage over the tip of the probe, dramatically increasing the time lag of the instrument. Figure
2 illustrates the depth offset and loss of detail that can result from such a situation. This hole
was logged prior to setting a shallow casing string by a commercial logging company and
immediately thereafter by our electric line equipment. There was a lost circulation zone at 60
to 70 m that was heated by the loss of fluid. The commercial log with the plugged probe
locates the zone at about 90 to 110 m because of the lag effect and broadens the apparent zone
of fluid loss. There has been some study of the use of deconvolution to obtain the hole
response from a temperature log made at a speed above that at which equilibrium is maintained
(see Nielsen and Balling, 1984, for a discussion of the topic).

-------------
-
-
....":::.--..::.---
_----- .......
Ci::·;;;;;~;;eJ l.!W
.......

--~----
100 f )
I /'
\ /'
E
J\
:5 200 1\ \
\
0.
a.>
o ~
l \
Old Maid Flat 7 -A, Oregon ,,\
)\

\}\
300 8/17/80 Commercial

8/17180 SMU

400L-L-~-L-L~--L-~~~-L~~--~~

10 20 30 40
Temperature, °C

Figure 2. Lag effect of mudcaked sensor. Example is from Old Maid Flat #7a near Mt. Hood, Oregon
(Blackwell, Murphey, and Steele, 1982).
HIGH-RESOLUTION TEMPERATURE LOGS 9

Generally, in clean, cased holes, logging speeds of 0.1-0.3mJs are optimal with probe
time constants of 4 to 10 seconds, assuming that the cage does not get plugged with mud,
cuttings, etc. In our logging we have preferred not to log at speeds fast enough to require
deconvolution for increased depth resolution, and because of the problem of plugging of the
probe at which point the time constant becomes so long that the quality of the log is severely
degraded. Even a relatively fast log, in a clean hole, with a heavy commercial probe returns
useful information, however. Given reasonable probe time constants and logging speeds the
resolution of formation temperature and gradient is limited by the amount of thermal
convection in the fluid column as described next, so that present equipment is capable of
returning maximum information on formation gradients in almost all logging situations.

Processing

Little processing, other than prior probe calibration, is required to extract temperature
data from the raw sensor resistance data. A simple way to calibrate is to submerse the probe
in a well-mixed, thermally insulated, temperature variable water bath and note the resistance
for a wide temperature range. The water temperature can be determined accurately using a
commercially produced, precalibrated, platinum or mercury standard temperature probe with
accuracy specifications from the NBS, submersed in the same bath. The probe should be
calibrated over the entire temperature range of possible logging situations. Commercial tools
generally will be precalibrated and supplied with their own processing software. Field checks
on calibration with an ice bath are easy and important. The ice bath should be prepared with
a slush of ice and water with the ice just floating in the water. The temperature will be
between 0.01 and O.OOI°C even iftapwater is used to prepare the bath.

Convection

Some effort has been made to develop a logging tool that measures the temperature of
the surrounding formation, and not that of the bore fluid (for example, the nuclear logging tool
evaluated by Ross and others, 1982) but instruments in use at this time measure the
temperature of the fluid. Theory and empirical evidence show that in general the equilibrium
temperature of the well fluid is that of the surrounding strata with no correction necessary even
for a cased and cemented well (see Diment, 1967). However, a vertical column of fluid with
temperature increasing with depth may experience a convective disturbance to the equilibrium
formation temperature. The earliest examination of this possibility within boreholes was by
Hales (1937), in relation to geyser eruptions. His results suggest that for any borehole with
temperatures above 4°C there is a critical thermal gradient, above which convection may be
expected in the bore fluid. He determined the critical gradient to be inversely proportional to
the fourth power of the hole radius. For a water-filled borehole at 95°C (203°F), the critical
gradient is (Jeffreys, 1937):

or 0.0014
az =-r-4-

where r and z are measured in centimeters. The coefficient on the right is viscosity dependent,
decreasing for higher temperatures and increasing to 0.014 for water at 20°C (68°F). The
theory was tested subsequently and defended by Auld (1948). It is apparent that, even at
surface temperatures, normal geothermal gradients (25°C/km; 1.37 °FIlOO ft) should induce
10 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

convection in water-filled wells with radii greater than about 2.5cm (1 "). In wells with other
types of fluids conditions may be different; for example air and oil may be more stable
(Sammel, 1968), but require slower logging speeds because they are less efficient at dissipating
heat from the probe.
Interest in the problem reappeared some years later, when Diment (1967) reexamined the
model and added an extra term for the adiabatic temperature gradient, which Hales (1937) had
discarded as insignificant. Results were virtually identical for most realistic situations.
Gretener (1967) tested and confirmed Diment's theory. Sammel (1968) also investigated
convection in wells with similar results. All of these results implied that in regions of average
thermal gradient, the fluid column in wells of radius greater than 2.5cm (1 ") filled with water
probably is convecting.
These findings could be interpreted to cast doubt on the validity oftemperature data from
virtually every borehole ever logged. In most situations, however, although convection may
be present within a fluid column, the magnitude of the gradient disturbance caused by
convection cells is determined to be small. The detailed correlation of lithology (other
geophysical logs) to thermal gradient logs at the I-meter level empirically proves that in
practical situations natural convection is a minor problem. Several studies (e.g. Wisian and
others, 1998; Gretener, 1967; Diment, 1967) have noted that even for boreholes larger than
critical radius, convection cells do not extend more than several well diameters in height, and
have little effect on the overall logged temperature profile on a meter scale. Empirical
measurements in a large diameter geothermal well (18 cm, 7 in) with high and variable thermal
gradient by Diment and Urban (1982) showed that the amplitude of the induced temperature
fluctuation at a particular depth is proportional to the thermal gradient, but that even at
gradients as high as 278°CIkrn the meter scale gradient was not changed (see Fig. 3).
Thus, we conclude that generally convection within a water-filled wellbore will increase
noise without significantly disrupting broader temperature gradient patterns. Regions of higher
gradient will yield noisier logs, as will sections of open holes that may have been washed out,
thus increasing the effective radius. The convection induced in large diameter washout areas
in open holes is another reason for the general low quality of open-hole temperature logs
(Blackwell, Murphey, and Steele, 1982).

Precision Thermal Gradient Logs

Examples of precision thermal logs that illustrate a variety of the points discussed here
for petroleum settings are illustrated in the following sections. The temperature gradient log
is of more interest than the absolute temperature because gradient is the parameter required for
determining heat flow. There are a number of finite difference methods for estimating gradient
from discrete depth-temperature data. The simplest is to take the temperature difference
between two successive data points and divide it by the depth difference, assigning the
resultant gradient to the depth of either datum. This is known as a forward or backward
difference, depending on whether the gradient is equated with the upper or lower depth.
Discretization error is reduced if we use the average of the forward and backward differences
at each temperature datum. This is known as a centered difference because each gradient
estimate is centered upon a specific depth datum. Once the depth-gradient log has been
produced, it may be necessary to filter high-frequency noise from the record. Generally, for
the logs described here, a 5-11 point mean or median filter is sufficient to clean up the record
and remove spurious spikes from the gradient log.
HIGH-RESOLUTION TEMPERATURE LOGS 11

TIME (MINUTES)

0 10 20 30 50 60

.10

.05

-.05

-.10
~
w
""
:::>
!;t .05
III
a..
~
t-
0

-.05

.05

-.05
.02
0
.02
·.02
0
5
·.02

Figure 3. Amplitude of temperature oscillation with time as function of mean thermal gradient in large diameter
geothermal well (East Mesa, California #31-1, Diment and Urban, 1982). Recordings are arranged in order of
decreasing thermal gradient from 278 °CIkm (1), to 178 °CIkm (2), to 103 °CIkm (3), to 37 °CIkm (4), to 8
°CIkm (5). Probe was centered in well and time constant was about 2 seconds.

WEST RANCH FIELD, TEXAS

In 1983, precision temperature logs were recorded for two wells in the West Ranch field,
near Vanderbilt, Texas (Fig_ 4, Table 1). The wells, #493 and #496, are separated by a
distance of about 1.0 km (0.6 mile). The data were recorded from the surface to a depth of
approximately 1830 m (6000 ft) in each well, with a temperature resolution of 0_001 °C, depth
increment of 1.Om (3.3 ft), and a logging speed ofO.06ms- 1 (12ft/min). Both wells were drilled
in the fall of 1981 (drilling took about 1 month for each well), cased, cemented, perforated in
the Frio, and tested. They had remained undisturbed since 1981 so there was more than
sufficient time to achieve thermal equilibrium, and therefore these wells represent ideal
conditions for precision temperature logging.
The geology of the field has been described partially by Galloway and Cheng (1985)_
The trap for the field is a simple domal anticline with a closure of over 50 m and dimensions
of about 6x4 km. The producing section is in a series of transgressive Frio (Oliogene)
barrier/strandplain sands and regressive shales starting at a depth of about 1550 m_ The two
wells logged are near the top of the structure.
12 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

96.55°

Figure 4. Location of West Ranch Field and #493 and #496 wells, Texas Gulf Coast.

A comparison of the temperature and gradient logs from the two wells (Fig. 5) shows a
high degree of correlation. The thermal profiles are almost identical down to about 1500 m
(5000 ft). The log from well #493 has a higher noise level than the log from #496 so the
gradient data were subjected to smoothing using a seven-point (6m; 20 ft) moving average.
The data are otherwise as recorded. The large oscillations and some of the differences between
the two logs in the bottom portion of the wells reflect the fact that these holes are in an old and
productive oil field, with production zones in the sands between 1500-2000 m (5000-6500 ft).
The extreme gradient excursions represent remnant production disturbances in the formations
(not associated with these wells which were not produced) caused by moving fluids and
expanding gas.
The lithologies penetrated by the two wells are dominantly sand and shale. The higher
gradient sections correspond to zones that are higher in shale content, whereas lower gradients
occur in sand-rich sections. The lowest gradients in both holes, above the potentially disturbed
section, are about 18-20°CIkm (1-1.1 °F/100 ft) and occur in intervals of the well where the
natural gamma-ray values are 40 to 60 API (see Figs. 6 and 7). Except within the depth range
of production disturbances, minimum gradients apparently correspond to clean sands.
Gradient highs show more variation with a range of 33-45°CIkm (1.8-2.5 °F/100 ft), tending
to increase with depth down to about 1500 m (5000 ft). The corresponding natural gamma-ray
values are 70 to 90 API units. The highest average gradients in both wells (outside the depth
HIGH-RESOLUTION TEMPERATURE LOGS 13

Table 1. Location of precision temperature logs described by this paper, Blackwell and Steele,
and Carter and others.

Location Longitude Latitude Town/Range Depth Ft. Depth M. Date Logged


This paper
West Ranch -493 -96.606 28.782 NA 6258 1908 5123/82
West Ranch -496 -96.599 28.789 NA 6241 1903 5123182
Postle-Hough 69 -101.606 36.851 5N113E1-36 6320 1927 10120/81
Postle-Hough 101 -101.650 36.837 4N113E1-3 6055 1846 5/1/90
Postle-Hough 103 -101.656 36.854 5N/13E1-33 6085 1855 4129/90
Postle-Hough 132 -101.659 36.913 5N/13E1-9 6007 1832 4130/90
Spiers -97.845 34.871 5N/6W-28 10950 3339 12111/90
Donehy#4 -122.069 39.565 20NI2W-14 4787 1460 8126/91
Miner Jones -122.069 39.526 20NI2W-33 3109 948 8127/91
Sprague Lewis 49-60 -122.086 39.541 20NI2W-33 5768 1759 8127/91
Sprague Lewis #1 -122.079 39.538 20Nl2W-34 5639 1719 8128/91
Sprague Lewis #3" -122.080 39.540 20NI2W-34 NA NA NA

Blackwell and Steele (1989a)


Watson #1 -94.905 38.477 18S123E-18 1915 584 6/9/81
GElS #1/Smokeyhill -97.575 38.872 13SI2W-32 3427 1045 11117/80
SMUWELL -96.782 32.844 NA 2673 815 10/5/82
MWX-1 -107.870 39.233 6S/94W-28 8477 2585 8/9/82
Chapman #1 -96.091 30.184 NA 8199 2500 2114/84
Parker 13-9 -101.500 40.133 2N/37W-9 3608 1100 7130/82
C. Hovland #1 -102.433 48.922 163N/90W-29 5887 1795 9/13164

Blackwell and Steele (1989b)


Rooks Co. -99.543 39.245 9S120W-27 3427 1045 11/15/80
Big Springs -95.478 39.013 12S/17E-13 2886 880 11125/81
LK-1 -98.167 38.383 19S/8W-23 751 229 11117n0
LK-2 -98.167 38.367 19S/8W-26 1076 328 11/17n0
Butler Co. -99.972 37.830 25S/4E-34 2417 737 11119/80
Sallyard#9 -96.477 37.833 25S/8E-36 1259 384 11119/80
T.E.Bird -95.923 37.860 25S/13E-24 1446 441 11118/80
Frontenac -94.742 37.457 30S/24E-2 1115 340 1110/80
USGS-BST -95.207 37.330 31S120E-22 1804 550 614180

Carter et al. (1998)


Garner -98.473 36.189 21N/15W-10 7071 2156 1131/88
Leforce -97.554 36.435 26NI7W-14 5332 1626 10m87
Frances -97.450 36.130 20N/5W-17 6009 1832 1015/87
Mackey -99.443 35.278 11N124W-4 7166 2185 712187
Cavitt -97.564 35.012 6NI7W-4 7051 2150 1/4188
Ferris -98.090 35.023 7N/9W-28 55485 16918 3/4/88
Perdasofy -98.262 34.498 4N112W-11 2952 900 12126/62

Gosnold et al. (1990)


Parker -101.510 40.150 2N137W-9CC 4182 1275 7130/82
Hardy -101.146 40.229 3N/34W-15BD 2493 760 7129/82
Burton -99.579 42.939 34N119W-8AB 2444 745 7122182
" used for conductivity only
14 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

Temperature,OC Gradient, ·C/km


o 20 40 60 80 o 20 40 60
Or-~~-.~~-r.

West Ranch
#493 & 496

500

E
~ 1000
o

1500

Frio
Formation

2000 '----A---L..---L_'----A-......

Figure 5. Comparison of temperature and thermal gradient logs from West Ranch wells #493 (dashed line) and
#496 (solid line).

range of production disturbance) are about 40°CIkm (2.2 °FIlOO ft) in the 1420-1500 m
(4650-5000 ft) range and this gradient is shown as the "shale" line in Figures 6 and 7. Deeper
than 1500 m (5000 ft) it is difficult to discriminate high gradients because of shale content
from those because of production disturbances.
The two thermal logs clearly illustrate the reproducibility of high-quality precision
temperature data and the one-dimensional thermal regime in this field. Despite a separation
of 1.0 kIn (0.6 mile), the measured thermal gradients can be correlated almost point for point
on a scale of3 m (10 ft) and the temperatures at the same depth in the two wells do not differ
by more than 0.35°C outside the depths of production disturbance. Such close correlation
indicates that individual variations in gradient are significant on a fine scale. This important
conclusion gives us confidence in inferring relative in situ thermal conductivity values for
depth intervals of meters to lOs of meters from precision thermal gradient logs. This result
contrasts with the kilometer-scale resolving power ofBHT data. Blanchard and Sharp (1985)
have postulated large-scale natural convection in the sands in this field on the basis of an
apparent cellular pattern to the BHT's. Based on these two temperature logs, such variations
probably are noise rather than signal because the logs in these wells show no evidence of
departure from conductive conditions except in the immediate producing zones.
Total gamma-ray logs usually are used to estimate the proportion of shale within a
sequence. A high gamma-ray count corresponds to a relatively high proportion of uranium,
thorium, and potassium-bearing minerals, which generally implies a clay-rich lithology.
HIGH-RESOLUTION TEMPERATURE LOGS 15

Gradient, °C/km Gamma Count, API Units Sonic velocity, IJs/m


o 40 80 o
50 100 150 200 400 600
o r----r-~===r-___,

Caliper

500

E
~ 1000
Q)
Cl

1500

Sonic
Production
/sturbance

2000 1--....1--1--....1-----1
20 40 60
Caliper, cm
Figure 6. Caliper, gradient, natural gamma ray, and travel-time logs for West Ranch #493.

Conversely, low gamma-ray activity generally implies a 'clean,' or clay-free lithology. There
is a close correlation between the behavior of the thermal gradient and gamma-ray logs for the
two wells as illustrated for well #493 in Figure 6. On a fine scale, low gamma-ray activity
generally is associated with low gradients within the drillhole. These sections can be
interpreted as sands with a low content of uranium, thorium, and potassium and a high quartz
content and thermal conductivity (thus a low gradient). Broader trends in gradient with depth,
16 BLACKWELL, BEARDSMORE, NISHIMORl, AND McMULLEN

Gradient,OC/km Gamma Count, API Units Sonic velocity, iJs/m


o 40 80 o 50 100 150 o 400 800
o

Caliper
500

E
~ 1000
Q)
o

1500

Sonic

20 40 60
Caliper, cm
Figure 7. Caliper, gradient, natural gamma-ray, and travel-time logs for West Ranch #496.

particularly within the predominantly shale sections, are mirrored on the gamma-ray log. The
thick shale unit between 1340-152Om (4400-5000ft) is defined clearly on both logs, although
the subsection of particularly high gradient does not have a distinctive gamma-ray signature.
The only significant zones of noncorrelation, above the disturbed region, are between
760-820 m (2500-2700 ft) and 1500-1570 m (5000-5150 ft). In the shallower section, the
gamma log shows sand whereas the gradient log shows shale, whereas in the deeper section
the gamma log shows a massive shale layer whereas the gradient log shows a more sandy
HIGH-RESOLUTION TEMPERATURE LOGS 17

lithology. Deeper than 1600 m (5250 ft), the gradient log is too noisy because of production
disturbances to identify any clear correlation or noncorrelation between the logs.
The results are similar for well #496 as illustrated in Figure 7, except that the quieter
gradient log allows a more detailed comparison. The fine-scale correlation between the
gradient and gamma-ray logs is particularly clear in the upper part of the well where rapid
variation in lithology has point-for-point correspondence on the two logs. Again, though, there
are zones of noncorrelation between 760-820 m (2500-2700 ft) and 1500-1570 m (5000-5150
ft). The gamma-ray log shows that the main shale unit is thicker than the gradient log shows.
These results show that gradient logs are as sensitive at least as gamma-ray logs in
distinguishing between sand and shale units. In regions where the two logs do not correlate,
there is no way of determining which is the more reliable indicator without confirming
lithology by some other methods.
One may expect a similar correlation between sonic velocity and temperature gradient,
because shale has the longest travel times and the highest gradients. However, this relationship
generally is valid only at shallow depths. Compaction effects cause velocity to increase with
depth faster in shale than in sand, so that travel times in the two lithologies converge and
eventually coincide. It is interesting to note that the compaction effect is the opposite for
temperature gradient because compaction tends to enhance the thermal conductivity contrast
between lithologies by reducing the low-conductivity water content in the sand.
It is immediately obvious from the logs (Figs. 6 and 7) that the velocity distinction
between lithologies is diminished markedly below about 1200 m (4000 ft). This is particularly
obvious between 1200-1650 m (4000-5400 ft) in well #496. The sand between 1260-1360
m (4150-4450 ft) has a low gradient, but much of it has high travel times. Furthermore,
neither the upper nor the lower contact of the thick shale unit has any expression on the travel
time log.
The quality of sonic logs is dependent on the condition of the holes, and there are several
depths in these wells at which the hole diameters are abnormal, particularly in #493 (Fig. 6).
Unfortunately, these depths tend to coincide with sand units as interpreted from the gamma-ray
and gradient logs. As there are only a few significant sand units in the holes, and many of
these are washed out, it is almost impossible to establish a value for the sonic velocity of the
sands from the logs.
These examples illustrate an important point. Below an arbitrary depth, in a sand/shale
environment such as the Gulf Coast it is not possible to deduce thermal conductivity from
velocity information alone. With independent lithological data (for example, a gamma-ray
log) it may be possible to relate thermal conductivity to lithology and sonic velocity, although
in different geological settings the velocity-depth curves for sand and shale will differ. This
represents a major limitation for techniques which attempt to calculate subsurface temperature
using only seismic data (e.g. Houbolt and Wells, 1980). However, techniques that use multiple
logs to determine both lithology and porosity (Brigaud, Chapman, and LeDouaran, 1990;
Griffiths and others, 1992) offer real promise if the bulk thermal conductivity values of the
various lithologies are known, a problem with the shale lithology.
Another interesting point arises from a comparison of the sonic and gradient logs. The
sonic log indicates a gradual increase in the velocity of shale with depth, as would be expected
during dewatering and compaction. However, the temperature gradient within the shale
remains constant, or perhaps even increases over the same depth range, contrary to the usual
assumption in basin thermal analysis (e.g. Funnell and others, 1996), based on the assumption
of a constant shale rock-component thermal conductivity and loss of a low thermal
18 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

conductivity water component. There are two possible explanations for this observation.
Either heat flow increases with depth, so that the expected higher conductivity at depth does
not result in a decrease in gradient, or else the thermal conductivity of shale does not increase
with compaction (Blackwell and Steele, 1989a). The latter possibility has implications for all
shale conductivity models, and is discussed briefly in the conclusion.
Thus gradient logs can be used with other well logs to define zones of similar lithology
and contacts between different lithological units. If heat flow is constant, then thermal
gradient is inversely proportional to thermal conductivity and the consistent gradients in
similar lithologic units indicates constant vertical heat flow in the well. Thermal conductivity
is a rock property that primarily is a function of rock composition and porosity, so gradient
logs should be sensitive to lithological change, as are other logs such as total gamma-ray count
and sonic velocity.
Thermal conductivity measuerments were made on 6 core samples (three shale and three
sand) from well #493 in the depth interval from 1307 m to 1867.5 m. The average value for
the three sands was 2.62 WImK and the average value for the three shales was 1.30 WImK for
a ratio of2: 1. Additional thermal conductivity information is available from McKenna, Sharp,
and Lynch (1996) who measured thermal conductivity on a nwnber of Frio core samples in the
SMU Geothermal Laboratory. They determined an average value for the clean quartzose
(>35% quartz) Frio sands with a porosity of 19 to 23% of2.73 W/mK. Clean quartzose sands
with a quartz content of <35% had an average value of 2.31 W/mK. A reasonable in situ
thermal conductivity of the cleanest sands should be about 2.5±0.1 W/mK with a 5 to 10%
temperature effect on the sand. With this value the heat flow for a gradient of 20° CIkm would
be about 50 m Wm-2 whereas for the gradient of 40° CIkm in the shales and a thermal
conductivity of 1.3 W/mK the heat flow would be about 52 mWm-2 • The agreement is close
considering the possible errors in the values of thermal conductivity.

ANADARKO BASIN, OKLAHOMA

The Anadarko Basin (Fig. 1) is a foreland basin in Oklahoma, formed primarily in the
Pennsylvanian on top of a Cambrian rift basin (see Johnson and others, 1988). The thermal
regime in the basin has been explored using detailed lithologic analysis, thermal conductivity
measurements, and detailed temperature logs by Carter and others (1998) and Gallardo and
Blackwell (1999). Details of some of the temperature logs run there have been described by
Carter and others (1998). Two additional areas, one field with four sites and one field with one
site, are described here, to illustrate the information they provide on the use of temperature
logging in the petrolewn setting.

Postle-Hougb Field

In 1991, precision temperature logs were collected from three wells in the Postle-Hough
field in Texas County, near Guymon in the Oklahoma Panhandle. The field is located in the
western edge of the Anadarko Basin (Fig. 8; Table 1). All three wells had produced oil and/or
gas from Pennsylvanian Morrow sands and were temporarily off production awaiting
recompletion or further examination. A fourth well in the field was logged previously in 1981.
High-precision temperature data at a depth interval of 0.1 m were collected from all holes to
a depth of about 1,800 m (6000 ft). A temperature-depth plot comparing all four wells (Fig.
9) shows the temperature logs to be essentially linear, increasing from a mean of about 15°C
HIGH-RESOLUTION TEMPERATURE LOGS 19

13E
I
CRISS-fO km /f

13-2 ~ COLO;;oo-.{ KANSAS


5N 5N
I OKLAHO/ ~o,tI.-Ho",h
field

103 36°52.5 /TEXAS I


.
69

I I I 11~1 I II 14y
13E
~-=-
1 mi 1 km
100° 37.5'

Figure 8. Detailed location map of wells in Postle-Hough field, Anadarko Basin, Oklahoma Panhandle.

Temperature,OC
20 30 40 50 60 70

400

#103

#13·2
800
E
.t:
Q.
GI

C 1200

1600
Postle-Hough Field

'.
2000L-----~----~----~----~----~-----L----~----~

Figure 9. Temperature versus depth curves for four wells in Postle-Hough field.

(59°F) at the surface to bottom-hole temperatures of about 60°C (140°F). However, the
different logs show detailed structures that are examples of how the quality of a temperature
log can be evaluated from internal evidence, because the theoretical effects of the many and
different disturbances are impossible to anticipate.
The temperature and gradient logs for well #101 (Figs. 9 and 10) have characteristics that
strongly suggest the well was not in thermal equilibrium when logged. The high amplitude
oscillations in the gradient log (especially above 600 m; 2000 ft), including several regions of
negative gradient, imply that the temperature in this well is not equilibrated to the temperature
of the surrounding strata. The section below 1600 m (5250 ft) seems to be the only part of the
20 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

Gradient Gradient Gradient Gradient Gamma Ray


°C/km °C/km °C/km °C/km API Units
40 0 40 0 40 0 40 o 100 200

200

400t--~---

600

800

E
~1000
Q)
o

Figure 10. Gradient logs for four wells in Postle-Hough field and natural gamma-ray log for well #13-2.
Expected gradient for heat flow of 55 mWm-2 for thermal conductivity values of (1) evaporites, (2)
sandstones/limestone, (3) Pennsylvanian and Permian red shales, and (4) pre-Permian marine shales from
Gallardo and Blackwell (1999) are shown by vertical lines for #103 and #13-2 wells.

well not dominated by noise on a 10 to 20 m scale. Although the average thermal gradient is
similar to that in the other wells, the perturbations in the log render it inadequate for detailed
analysis. The parallelism of the temperature curve to the equilibrium ones except in the
bottom 100 m of the well demonstrates the heat exchange nature of production disturbances
and suggests that the well actually had been producing more recently than thought at the time
HIGH-RESOLUTION TEMPERATURE LOGS 21

of logging. The characteristics that would allow the departure from equilibrium to be
recognized if only the one log was available are the oscillatory temperatures and negative
gradients, and the lack of detailed correlation (at the scale of 1 to 10m units) to the lithology.
The temperature and gradient logs from well #69 (Figs. 9 and 10) indicate that this well was
closer to equilibrium, but showed a number of disturbed regions. Between 400-850 m
(1300-2800 ft) temperature is elevated with respect to wells #13-2 and #103 (Fig. 9). This type
of thermal disturbance typically is seen when fluid flows between two or more zones in an
open well. Well #69, however, is cased so the explanation is not so simple. There may be
fluid flow either behind the casing or between perforated sections. There are temperature steps
and gradient spikes at 1100, 1535, and 1857 m again indicating some nonconductive
disturbance in the wellbore possibly the result oflarge-scale, but slow, vertical fluid flow. The
overly smooth gradient curve indicates that there was some small flow in the borehole at the
time of logging, that the probe sensor cage was plugged, or that there was oil in the well and
the time constant of the probe was affected by the low conductivity of oil. This example
illustrates how detailed knowledge of the well construction is necessary to explain departures
from thermal equilibrium and conductive conditions. In spite of these problems, the larger
scale gradient features correlate with wells #13-2 and #103 and the well could be used for
lithologic identification in most sections at a 5 to 10 m window with some degraded precision
compared to the best logs. For example, zones oflow gradient are evident around 380m (1250
ft), 520 m (1700 ft) and 800 m (2600 ft), as are high gradients in the shale dominated section
below 1700 m (5550 ft).
The gradient log for well #103 (Fig. 10) is of visibly higher quality than those of the two
previous wells. Although two high amplitude spikes occur (at 880 m and 1290 m; 2900 ft and
4250 ft), noise levels are low. These properties are characteristic of a well close to thermal
equilibrium. The only section where equilibrium remains questionable is in the depth range
of 850-1 020 m (2800 to 3350 ft) where there are moderate amplitude oscillations that do not
occur in the other wells for the same section. The lithology for this well correlates closely with
changes in the character of the gradient log based on the correlation of the log character to that
of well #13-2 and the changes in gradient near the top of the Pennsylvanian and the Cherokee.
Well #13-2 seems to be close to equilibrium. The only questionable section is between
1000-1130 m (3280-3700 ft) where the gradient log is oscillatory, whereas on the #103 log
the same interval exhibits an essentially uniform gradient. The formation picks for this well
correspond closely with changes in the nature of the gradient log. Correlation with the
gamma-ray log available for the well below 550 m is good on both a broad and fine scale (Fig.
10), except in the questionable section. The areas of higher gradient also have higher gamma-
ray activity. Lower temperature gradients are associated with sections dominated by
limestone, dolomite, and evaporites which have lower natural gamma-ray activity. Many of
the units are so thin, however, that their thermal signal may be aliased. Near the top of the
Pennsylvanian System both wells #103 and #13-2 Show changes in gradient character that are
typical of the regional thermal character of the Anadarko Basin (Carter and others, 1998;
Gallardo and Blackwell, 1999). Higher gradient (lower thermal conductivity) marine shales
occur in the Pennsylvanian (dashed line #4) and lowest gradient (highest thermal conductivity)
evaporites disappear below the base on the Permian (dashed line # 1). The highest gradients,
associated with in the Cherokee Group shales, can be recognized in all four wells. This section
has the highest natural gamma-ray activity as shown on the log for well #13-2.
These four wells are all in the same region and all penetrate the same section. However,
there are obvious differences in the character of the temperature gradient logs. The differences
22 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

can be attributed to different degrees of thennal equilibration. Some typical features of


disequilibrium have been illustrated using these gradient logs. All four wells show at least
minor wellbore effects resulting from production and other disturbances. These effects include
production heating and a high noise level on log # 10 1, spiky gradient zones in #69, and regions
with unexplained gradient oscillations in the 850-1130 m (2800-3700 ft) depth range in wells
#103 and #13-2. Induced movement of fluid because of production undoubtedly has caused
some thennal disturbance to the gradients in the deepest parts of the wells (>1800 m; 5900 ft).
The wells each have intervals with smoother or more noisy gradient logs than other
intervals in the well. These variations may relate to the type and stability of the fluid in the
wellbore. These particular wells were filled with fluid only days before the logs were run, but
this time was sufficient for local equilibrium as indicated by wells #103 and #13-2. So the high
noise in #101 indicates that the well had been producing more recently and the wall
temperatures were farther from equilibration than in the situation of the other three wells.
Possibly the well records were incorrect in this instance or the wrong well was logged by
mistake. The close correlation between gamma-ray count and temperature gradient in the two
#103 and #13-2 wells gives a strong indication that heat flow is dominated by conductive
processes. Thus, the gradient logs give a good indication of the contrast in thennal
conductivity between fonnations in this set of wells and the temperature regime in this area.
In this example no samples were available from the field for thennal conductivity
measurement. So measurements were made on 8 cuttings samples from a well (Criss) about
10 kIn (6 mi) northeast of the field. The samples range from 1237.5 to 1813.6 m in depth,
were predominately limestone in lithology, and have an average thennal conductivity of2.5
W/mK corrected for porosity and temperature. The gradient that corresponds with the lowest
natural gamma-ray activity is about 22 °CIkm so the heat flow may be on the order of
55mWm-2 •
Typical thennal conductivity values from Midcontinent lithologies from several
companion studies are shown in Table 2 (after Gallardo and Blackwell, 1999). The expected
gradient for an assumed heat flow of 55 mWm-2 for the typical thennal conductivity values of
evaporites, sandstones/limestone, Pennsylvanian and Pennian red shales, and, pre-Pennian
marine shales from Gallardo and Blackwell (1999) are shown by vertical lines on Figure 10
for the #103 and #13-2 wells (see also Carter and others, 1998, figs. 7 and 8).
The top of the Pennsylvanian section occurs at about 1061 m. Gallardo and Blackwell
(1999) determined that the average thennal conductivity of the Permian "red shale" section was
about 2.0 W/mK and that the average Pennsylvanian marine "shale" had a thennal conductivity
of about 1.5 W/mK. The log for the #13-2 well shows an increase in the amplitude of the
highest gradient sections above and below about 1061 m in almost exact proportion to the
expected difference between the two lithologies. The Chase Fonnation at 796 m is an
evaporitic layer as can be seen from the low gradient there. There also is an overall
consistency of the predicted and observed gradients and their ratios with the lithology as
generally seen on the gamma-ray log verifying the conductive nature of the thennal setting in
the field. The gradients in the #103 well are similar except that the gradients in the
Pennsylvanian shales are significantly lower than those in the #13-2 well and the gradient log
thus is more subdued. There may be a small thennal disturbance present in the well, or some
other effect to cause some of the shale sections to have a reduced response on the gradient log.
The gradients and gamma activity corresponding to shales in the Pennsylvanian are about
36 °CIkm and 100 API except at the bottom of the well where the gradients (and the gamma-
ray counts) jump significantly so that the average gradient is about 55 °CIkm in the shales
above the producing horizons and the average gamma count is 200 API. If the heat flow
HIGH-RESOLUTION TEMPERATURE LOGS 23

Table 2. Estimated typical thermal conductivity values (W/mK) for Anadarko Basin
lithologies.
Lithology Laboratory Inversion Kansas
(Carter et al.. 1998) (Gallardo & (Blackwell &
N Avg. Std. Err. Blackwell. 1999) Steele, 1989b)
Pennian redbed 0 1.712.0 2.00
Other shale 57 1.47 0.06 1.50 1.26
Sandstone 141 4.01" 0.10 4.21 2.47
Granite wash 5 4.13 0.23 3.35
Limestone 52 2.96 0.05 2.90 2.94
Dolomite 5 4.50 0.24 4.38 4.34
Anhydrite 3 6.68 0.32 4.65 4.65
* Average of 125 samples over 2.8 W/mK.
** 1.7 for Ferris #1-28. 2.0 for most other sections of the basin.

is 55 mWm-2 and the mean shale thennal gradient is 36 °CIkm, then the inferred thennal
conductivity would be 1.53 W/mK, identical to the average shale value that Gallardo and
Blackwell (1999) detennined from well calibration, for the shales in the Anadarko Basin.
Thus the high gradients in the deepest shales imply that those shales have a lower thennal
conductivity (about 1.0 W/mK). The apparent thennal conductivity effect may be the result
of oil saturation of the shale above the reservoir, variations in shale lithology, or heat-flow
changes because of disturbance associated with production. Variations in shale lithology is
the preferred explanation because a similar abrupt increase in thermal gradient is seen in wells
in the Texas Panhandle that were drilled as test wells and not for hydrocarbon purposes and
because the gamma-ray signature changes. This change in the gamnia-ray character suggests
a lithologic change. In Kansas, the shale conductivity was detennined to be as low as 1.0
W/mK in darker, more organic rich, radioactive shales such as the Devonian Woodford shale.

Spiers #1

In October, 1990, a precision temperature log with a sample interval of 0.1 m was run to
a depth of 3338 m (10950 ft) in the well Spiers #1 in the Chitwood field, Grady County,
Oklahoma. The well was completed in 1946 to a total depth of3616 m (I I ,865ft) and plugged
back to 3606 m (1 I ,830ft). Last production was in 1984, when the well was producing only
two barrels of liquid per day. Within the hole, the base of the Pontotoc Fonnation, which
approximately correlates to the base of the Pennianitop of the Pennsylvanian, is at 1583 m
(5195 ft).
A temperature-depth log (Fig. II) shows that the temperature increases from about
15.5°C (60°F) near the surface to a temperature of about 80°C (176°F) at 3340 m. The average
thermal gradient increases near the Pennian-Pennsylvanian boundary from about 16.5°CIkm,
to about 22.9°CIkm in the deeper part of the well. The only section of the log which does not
seem to demonstrate a basically conductive thermal pattern is between 2880-3080 m
(9450-10,100 ft) where cooling of the well by gas expansion in obvious. Gas enters the well
at 2885m (9465 ft), presumably behind the casing, which causes a cooling disturbance as it
expands into the annulus. The gas apparently moves down the hole and enters (or has entered
in the recent past) the tubing at a depth of about 3075 m (10,088 ft). The interval is evident
on the temperature-depth log as a negative spike where the gas enters, and temperatures at
each depth in the interval of flow are slightly lower than temperatures interpolated from above
and below. On the gradient log (Fig. 11) large-scale oscillations are evident near the gas
inflow zone. Presumably, these are the result of convection associated with the high
temperature gradients.
24 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

The predominantly conductive nature ofthe heat flow in Spiers #1 is demonstrated by a


good correlation between the gradient log and conventional gamma-ray and electric logs (Fig.
11). Unfortunately, no single conventional log covers the entire range of the gradient log and
the logs are old and uncalibrated. A natural gamma log was run from the surface to around
1830 m (6000 ft), and a resistivity log from 1830 m (6000 ft) to total depth. There is good
correlation at the 3-15 m (10-50 ft) scale between the gradient log and the respective wireline
logs. In general, shale sections associated with high-temperature gradients have high gamma
activity and low electrical resistance. However, it is difficult to make a one to one correlation
in many ofthe thin-bedded Permian units (above 1583m; 5195ft). The clearest correlation can
be seen in relatively thick shale units whereas many of the sand and limestone units are
relatively thin and the thermal signal of individual beds may be below the spacial resolution
of the gradient log. For example, about 1275 m (4180 ft) there is a fairly uniform shale bed
on the order of 60 m (200 ft) thick which is clearly apparent on both the gradient and gamma-
ray logs. The thermal gradient in this unit is about 20.0°CIkm. The lowest gradients are
harder to characterize but generally lie between Il-l4°CIkm. These may be assumed to
represent typical sandstone, limestone, and dolomite/evaporite units.
In the rocks of Pennsylvanian age, the shales have average gradients significantly higher
than those in the Permian section. Typical shale gradients in the older rocks peak around
30°CIkm. Sands in the Pennsylvanian section, with the exception of the thick Springer-age
Primrose Sand are similar to those in the Permian, thin and difficult to distinguish with
certainty. However, the undisturbed zones with minimum gradients average about
11-14°CIkm within the same range as for the high conductivity units in the Permian. These
zones are predominantly sandstone and the consistancy in thermal gradient implies that the
thermal conductivity of sand remains essentially unchanged across the Permian-Pennsylvanian
boundary whereas the shales increase in gradient.
The Primrose Sand is a 60 m (200 ft) thick sandbody of Springer age near the base of the
well and shows clearly on the gradient log. There is a clear correlation between the resistivity
and gradient logs in this section, with the sand having a relatively high resistivity and low
gradient. The mean gradient in the Primrose Sand is 11.5±0.2°CIkm.
No thermal conductivity data are available from Spiers #1 or from any well in close
proximity. However, the Ferris # 1 well that was logged to 5 km by Carter and others (1998)
and used as a thermal conductivity test well by Gallardo and Blackwell (1999) is only 32 km
(20 miles) NNW of the Spiers #1 well. In order to produce an estimate of heat flow for the
Spiers #1 well, we can use the regional average values of thermal conductivity for Anadarko
Basin lithologies from that well. Ferris #1-28 had been shut in for several months when a
high-precision temperature log was run (Carter and others, 1998; Gallardo and Blackwell,
1999). The well bottomed at 5601 m (18,376 ft) in the Pennsylvanian Springer Group. A best
estimate for heat flow in Ferris #1-28 was calculated to be 39±3 mW/m2 (value range is
standard error) from the gradients and thermal conductivity of two sandstone intervals.
Conductivity measurements were made on core samples of Marchand Sandstone (Skiatook
Group) and a Morrow sandstone (Morrow Group) taken from wells in close proximity to Ferris
#1-28. As in the Postle-Hough field the four predicted gradients for the typical Anadarko
Basin lithologies are shown on the gradient plot for the Ferris # 1 well in panel 2 of Figure 11
corresponding to the measured heat flow of39 mW/m2 •
Shale is the most difficult lithology to characterize in situ because of sampling problems
(Blackwell and Steele, 1989a). For example, the laboratory measurements of thermal
conductivity published by Carter and others (1998) were made on dry core samples because
the shale disaggregated upon saturation. Obviously, these values cannot represent in situ
HIGH-RESOLUTION TEMPERATURE LOGS 25

Tern peratu re Gradient Gradient Gamma Ray


°C °C/km °C/km Total counts
o 50 0 20 o 20 o 40 80
o

500

1000
,
\
\ Ferris
\ #1-28

1500
\I,
\
,,
E ,
\,
:5
a. ,,
<I> ,,
Cl
,,
2000 ,,
,,
\

2500

3000
,,
,
\,
,,
3500
Ferris #1-28 Spiers #1 0 20 40
Resistivity, Ohm-m

Figure 11. Temperature, gradient natural gamma-ray, and resistivity for Spiers #1 and gradient log for Ferris #1
(Carter and others, 1998). Expected gradient for heat flow of 40 mWm-2 for thermal conductivity values of(1)
evaporites, (2) sandstones/limestone, (3) Pennsylvanian and Permian red shales, and (4) pre-Permian marine
shales from Gallardo and Blackwell (1999) are shown by vertical lines for two wells.

thennal conductivity as accurately as measurements made on saturated samples. The inversion


technique, using interval thennal gradients to deduce the conductivity contrast between
different lithologies, should give a more accurate estimate of in situ conditions. The process
26 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

confirmed that typical Permian redbed shales have a higher thermal conductivity than the
marine shales in the older section as already seen in the Postle-Hough field and for the
Anadarko Basin in general (Gallardo and Blackwell, 1999).
As was done for the Postle-Hough and the Ferris #1 wells, the gradients for the four
lithologic groups for a particular heat flow, in this situation 40 mWm·2 , are plotted on the
gradient log for Spiers # 1 in panel 3 in Figure 11. Three categories of lithologies, excepting
evaporites, correspond to the three categories of gradients in the well and the "red shale"
marine shale thermal conductivity contrast is seen clearly near the PermianlPennsylvanian
contact. The value given in Table 2 for evaporite represents the in situ thermal conductivity
for a composition of 70% anhydrite and 30% redbed. The mud log for Spiers #1 notes only
minor amounts of evaporite in the Permian section, which is typical for the southeastern part
ofthe basin and there are no gradients that match curve 1, the theoretical evaporite gradient,
in the well.
The relative values ofthe temperature gradient in different lithologies should be inversely
proportional to the thermal conductivity values. Thus, the gradients in the Permian redbeds
should be about 0.75 ofthose in the Pennsylvanian shales, and the gradients within the sands
should be 2-3 times those in the shales. These ratios are close to those actually observed on
the Spiers #1 gradient log. Heat flow in Spiers #1 is apparently about 40 mW/m2, consistent
with the results for Ferris #1-28 of39±3 mW/m2.

WILLOWS-BEEHIVE BEND GAS FIELD, CALIFORNIA

The Willows-Beehive Bend field produces gas from Cretaceous sands and is located in
the northern part of the Sacramento Basin in central California (Figs. 1 and 12). The geologic
section consists mostly of shale and clay-rich arkose and graywacke. All rocks shallower than
about 450 m (1500 ft) are continental and post-Eocene in age. A thick Eocene valley fill, the
Princeton Gorge, crosses the marine Eocene shale dominated section in the area of the field,
with a base at about 1080 m (3600 ft) in Sprague Lewis #1. Below this unit are the gas-
hosting, intertonguing marine, upper Cretaceous deposits of the shale dominated Kione and
coarse clastic Forbes Formations. The contact is about 1280 m (4200 ft) in Sprague Lewis #1
and the base of the Forbes is just below the maximum depth ofthe log. Temperature logs were
made in August 1991 in four wells in the field, all roughly along strike and spanning a distance
of about 5 km (3 mi, see Table 2 and Fig. 12). All had been in production but were
temporarily off-line awaiting recompletion or further examination. There was not much log
information in this field with only a SP log available for the Sprague Lewis #1 well.
Conditions differed between wells at the time ofthe survey. Doheny #4, about 5 km north
of the other three wells, was logged openhole with no surface pressure control, but its
producing section was isolated with a plug. The other wells were logged using surface
pressure control equipment but the wells otherwise were open to the bottom. The two
Sprague-Lewis wells had significant shut-in pressure and the gradient logs were noisy. All
of the wells had been topped with water several days before the survey.
Doheny #4 exhibits a strong thermal disturbance within the top 150 m (500 ft). It seems
that the fluid within this zone may have been disturbed just prior to logging. No explanation
for the disturbance was put forward by the logging party, and it remains unexplained. The
gradient log is unusable within this top section but the remainder of the log seems stable and
reliable.
HIGH-RESOLUTION TEMPERATURE LOGS 27

NORTHERN
SACRAMENTO
BASIN,
CALIFORNIA

~ ..,.....
.
:z
a:

x
OUTCROPS OF uPPER
CRETACEOUS STRATA

FORBES GAS FIELDS

EDGE OF FOOTHilLS

OTHER GAS FIELDS

# CITIES

Figure 12. Location map of Willows-Beehive Bend field, northern Sacramento Basin, California.

The temperature and gradient logs for Sprague Lewis #49-60 (Figs. 13 and 14) indicate
two zones about 1550 m (5100 ft) with reduced temperature and negative gradient. The nature
of these thermal disturbances is typical of the cooling effect of gas expanding into the
wellbore. Although this well was capped the producing zones were open to the formation.
Either the gas-bearing formations produced some gas during, or just prior to, the logging run
or the expansion during the production of the well has cooled the formation well away from
the well. The upper half of the gradient log is noisy and a 31 point averaging filter was used
for the curve in Figure 14.
The temperature logs show nearly linear temperature-depth curves for all wells (Fig. 13),
increasing from a mean annual surface temperature of 18°C (65°F) to about 50°C (122°F) at
1700 m (5600 ft). Unlike the previous examples the temperature logs indicate some lateral
thermal changes in this field as the Miner Jones #1 and Sprague Lewis #1 wells are about 2°C
hotter below 1000 m than the other two wells. The gradient logs (Fig. 14) are all noisy,
probably because of the recent disturbance ofthe well fluids, and required more filtering prior
to display. A 3-meter median filter followed by a 3-meter average filter was sufficient to
reduce noise in all situations except Sprague-Lewis #1 as mentioned next. Within individual
formations, the sand units typically have only 30-40% quartz with the remainder of the coarse
clastic component composed of feldspar and rock fragments of metamorphic and volcanic
provenance. This lack of high thermal conductivity quartz in the coarsest units explains the
lack of gradient variation because of the lack of major conductivity variation. The gradients
in the wells generally are between only 15 and 20 °CIkm.
28 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

10 50 60
0

200 Miner Jones #1

Doheny #4
400
Sprague Lewis #1
600 Sprague Lewis #49-60

E 800
.c
ii
GI 1000
C
,
1200 ,,
Willows-Beehive Bend
,
,,
1400
" '-
' ..., ......
1600
"""',

1800
Figure 13. Temperature versus depth for four wells in Willows-Beehive Bend field with detailed temperature
logs.

There are zones within Sprague Lewis #1 that stand out on the SP log (Fig. 14), in spite
of the homogenous nature of the formations. The SP log in the interval between 800-1250 m
(2600-4100 ft) shows significant variations that correlate with the general character of the
gradient log, that is zones with lower SP, probably the sandier zones, have lower thermal
gradients. That depth range, along with a similar zone between 1500-1650 m (4900-5400 ft),
coincides with packets of coarse clastic deposits that invade a dominantly shale environment.
The gradient log in this well was noisy and required an unusual amount of smoothing (45 point
average of 0.3 m data) so the gradient log in Figure 14 is averaged more than the other logs
making detailed correlation with the SP log more difficult. A minor production disturbance
is present in the vicinity of 1500 m.
The mean thermal gradients for all four wells and a number of smaller intervals from
Sprague Lewis #1 are presented in Table 3. The mean gradients for the four wells range from
only 18.0±0.2 °CIkm for Doheny #4 to 19.7±0.2 °CIkm for Sprague Lewis #49-69 (about LO-
U OF/100ft). The gradients are low and show little vertical variation, in contrast to the wells
in the Midcontinent and Gulf Coast illustrated previously. The lithologies in these wells are
"impure" and difficult to characterize in terms of sand, shale, etc. This type of lithology
variation probably is characteristic of many continental margin settings, particularly active
ones, and the in situ thermal conductivity will be difficult to characterize on cuttings samples
alone.
Thermal conductivity measurements were made on 20 core samples from both Sprague
Lewis #1 and nearby Sprague Lewis #3. The divided bar technique (e.g. Blackwell and
Spafford, 1987) was used to measure thermal conductivity of the cores at 21°C (70°F). Most
samples were from clastic rocks with sand-size grains, but a small number of shale samples
allowed representative measurements in all major lithology groups. Many of the samples
proved unsuitable for the divided bar technique. The shallow samples were poorly consolidated
HIGH-RESOLUTION TEMPERATURE LOGS 29
Gradient, ·C/km Gradient, C/km Gradient, ·C/km Gradient, ·C/km SP, mV
10 20 30 10 20 30 50 100
Or-'-~--~'-~--r-~-r~~.--r-.r-~-r~--T"~~

200

400

600

800
E
.c
C.
CD
o
1000

1200

1400

1600

10 20 30 o 5 10
Figure 14. Thennal gradient logs for four Willows-Beehive Bend wells and SP log for Sprague LeWis #1.
Dashed lines at 14.5 and 20 0C/km correspond to predicted gradients for heat flow of 28 mW/m·2 and thennal
conductivities of2.0 and 1.4 W/mK.

and some samples with pebbles disintegrated because the strength contrast between the grains
and matrix was too great.
Sandstone porosity shows a weak depth dependence. Shallower than about 1200 m
(4000 ft) the pore volume is consistently around 28-30%. Deeper than about 1600 m (5300
ft), however, the average sandstone porosity drops to about 23%. Overall, the shale is less
porous than the sandstone. The average thenna! conductivity for six saturated sandstones
above 1200 m (4000 ft) is 1.92±O.08 W/rriK. For the five saturated sandstones below 1500 m
30 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

Table 3. Interval gradients for wells and calculated heat flow for intervals within Sprague
Lewis #1.
Well name Depth interval Gradient Conductivity. Heat flow
(m (ft]) (OCJkm) {WInU9' (mW/m2)
..._..................QQ~.!:!.11..!~~L........._..__._._¥.±-li~2!!!.~m.:L....._!~~Q..!.l1.L..___...____..._._._.__._.___.____._.._.._..
......_._..._.M.i!!.!:r.lp.!!~.!~L._.._.._.___l?.~2.!!ill.OO-ll9!tL......._.12.:1..:t: 0.2 _._.._._._._..___.._____._.._._.___
..........~p..!"!!&!1..!:•.~~!~_~?..:§Q...._._._.J.~~~737!!!.~5l.Qg.J..._ .......!2.1_.:t:.9:.L__._.._.._. __.___._____._.__.._. __
Sprague Lewis #1 488-724m (1600-2375'] 18.8 ± 0.2 1.4 26
808-847m (2650-2779'] 19.9 ± 0.2 1.4 28
847-914m [2780-3000'] 14.6 ± 0.2 1.73 25
884-917m [2900-3010'] 14.2 ± 0.2 1.79 25
1067-1103m [3500-3620'] 18.8 ± 0.2 1.82 34
1104-1182m [3621-3879'] 18.4 ± 0.2 1.82 33
1280-1539m [4200-5050'] 20.2 ± 0.2 1.4 28
1618-1661m [5307-5450'] 21.0 ± 0.2 1.4 29
___________________ 1~~UQL~~2~~~J ____ J2J________ _________ ___ _
~~ 1~

152-1219m [500-4000'] 17.7 ± 0.2 [1.6] [28]


1402-1448m [4600-4750'] 16.6 ± 0.2 [1.7] [28]
1463-1478m [4800-4850'] 29.5 ± 0.7 [0.95] [28]
1485-1494m [4871-4900'] 19.7 ± 1.3 (1.4] [28]
1494-1509m [4900-4950'] 27.3 ± 0.5 [1.0] [28]
1219-1719m (4000-5641') 21.1 ± 0.2 (1.3) [28]
Conductivity values in (bIllCkels] are inferred values for a heat flow of 28 mW/m'.

(5000 ft), the value is 2.10±0.10 W/rnK.. The average of three shale cores from 844 to 1745
m in depth and with porosities of 2.13 to 2.41 gmlcc is 1.40 W/rnK.. Thus the shale
conductivity values are almost identical for samples with widely different densities. These
figures also are consistent with the only moderately higher thermal gradients observed in the
shale sections. Based on the thermal logs and the limited measurements, shale conductivity
in the Willows-Beehive Bend Field seems little affected by porosity (depth). The relative
magnitude of the thermal gradient in different lithologies should be inversely proportional to
the thermal conductivity values. Thus, according to these limited lab measurements the
thermal gradient in the shale should be about 1.4 times that in the sandstone.
Coarse clastic units may represent the most reliable intervals for determining heat flow
because of the difficulty in evaluating the in situ thermal conductivity of the shale. However,
the thickness and geometry of these units in this field is not ideal. Sandstone generally has
thermal conductivity about 3-4 WIrnK. as a result of its high quartz content. The conductivity
values of about 2 W/rnK. for the Willows-Beehive Bend sandstones are the result of both high
porosities and a lack of quartz. The slightly higher thermal conductivity of the deeper sands
can be attributed to lower porosity values.
In many sedimentary basins, shale is the most abundant rock type, and because it has low
~hermal conductivity it tends to control the thermal structure of the basin more than any other
lithology. In this part of the section in the Sacramento Basin, however, there is not a major
thermal conductivity contrast between sandstone and shale (at least within the logged depth
range) so the shale effect is not as dominant. The gradient logs show only small variations
with lithology when compared to the other settings described here.
Average thermal gradients for intervals around each thermal conductivity sample that
seemed to represent the lithology of that thermal conductivity sample were combined with the
measured thermal conductivity to determine interval heat flow within Sprague Lewis #1 (Table
3). Some of the intervals were selected to correlate with the depths sampled in Sprague Lewis
#3. Sandstone conductivity in the bottom part of the well was decreased by 5% to allow for
the temperature effect. The heat-flow values are calculated from the product of the thermal
conductivity and the least-squares fit to the temperature-depth curve in the intervals listed.
The statistical uncertainty can not be calculated for individual intervals because there is no way
to assess the error in the thermal conductivity. However, the uncertainty in the heat flow value
for the entire well can be evaluated based on the spread in the interval heat-flow values. The
HIGH-RESOLUTION TEMPERATURE LOGS 31

best estimate of heat-flow for the well is a depth weighted average of all values, 28±3 mW/m-2•
The mean gradient for Sprague Lewis #1 between 500-1720 m (1650-5640 ft) is 19.0°CIkm.
If the section is considered to be 25% sandstone, then the mean thermal conductivity is about
1.5 W/mK. The similarity of the gradients in the other three wells, suggests that heat flow
should be similar in each.
It is possible to use the mean heat flow to estimate thermal conductivity in zones where
samples are not available. This inverse calculation is particUlarly useful for characterizing
intervals where sampling is difficult or problematical, such as shales, and to identify zones in
the well of possible nonconductive behavior. The thermal gradient across the shale interval
can be estimated and the heat flow can be used to solve for the shale conductivity. The results
can be cross checked against lithologic or log information if available or used to characterize
the thermal conductivity for whole formations where sampling is difficult or nonexistent. This
procedure was used in several sections of Sprague Lewis #1 and the results are the
conductivity values in square brackets in the bottom part of Table 3. Of particular note are the
low values in two zones between 1463 m (4800 ft) and 1494 m (4900 ft). The low values
might be the result of shale lithology or perhaps areas where the effects of gas in the pores
reduce the conductivity. Dashed lines at 14.5 and 20 °CIkm are shown on Figure 14 for each
gradient log. These correspond to a heat flow of 28 mWm-2 and thermal conductivities of 1.4
and 2.0 W/mK.
The heat flow in this field is low compared to the global average. In a nontectonic setting
such a low heat flow might suggest a large-scale deep aquifer transporting heat away from the
region. However, the low heat flow in the Sierra Nevada Mountains and the Great Valley of
California is a crustal effect resulting from the cooling effects of subduction beneath this area
as recently as 5 Ma (Blackwell, 1971). The cooled crust is heating up at this time and the rapid
rise of the Sierra Nevada Mountains is because of the associated thermal expansion. A thermal
model for the region was discussed by Saltus and Lachenbruch (1991).

CONCLUSIONS

The general points and conclusions of this paper are:


(1) Temperature gradient logs make useful lithology indicators. Of the abundant
lithologies, shale exhibits the highest thermal gradients, whereas sands and carbonates
generally have low gradients. Because most sedimentary basin settings are one-dimensional
and, at least, locally conductive, a good correlation should exist between an equilibrium
thermal gradient log and other lithological indicator logs such as gamma ray and sonic
velocity.
(2) A poor correlation between sonic velocity and temperature gradient indicates that it
is not possible to calculate thermal conductivity based purely on seismic velocity. Some
independent lithological indicator (such as a gamma-ray log) is necessary to draw any
relationship.
(3) The information described in this paper supports the findings of Blackwell and Steele
(1989a, 1989b) that the conductivity of shale does not increase with depth-increase/porosity-
decrease, as would be expected from a simple compaction model. The explanation they
proposed for this lack relates to the anisotropic thermal conductivity of sheet silicates. Clay
platelets have higher thermal conductivity parallel to the plates compared to perpendicular.
At shallow depths, clay platelets are oriented randomly and heat can be transmitted along the
plates of many clay particles. However, the high porosity also lowers the overall thermal
32 BLACKWELL, BEARDSMORE, NISHIMORI, AND McMULLEN

conductivity of the mud because of the high water content. With increasing depth 90mes
compaction. Clay particles progressively rotate towards the horizontal, so that vertical heat
flow must increasingly traverse through the low conductivity axis of the platelets. This effect
offsets the conductivity increase resulting from water loss. Thus, the conductivity variation
with depth for shale may be relatively small even after large porosity changes.
(4) A thermal gradient log from a well not in complete thermal equilibrium will exhibit
such features as high noise level, spikes, and regions of high-amplitude oscillation. A gradient
log that closely correlates with the gamma-ray log (or other lithology sensitive type oflog) for
the same depth range, has a low noise level, and no spikes or steps is likely to represent
equilibrium thermal conditions in a conductive setting.
(5) Precision temperature and gradient logs can yield fine-scale information about fluid
movement within a well, including the location of entry and exit points, and regions of gas
production and expansion. These data may be unavailable by any other method.
(6) Continuous precision gradient logs allow thermal conductivity measurements from
specific depths to be associated with specific thermal gradients. This allows a number of heat-
flow estimates to be made in anyone well, and a weighted mean can be used as a best estimate
for the entire well. Also, once a heat-flow value is know for a well, it can be used in
intermediate zones to obtain a best estimate of thermal conductivity. This is one way of
investigating in situ conductivity of formations that are difficult to sample.
(7) Technology changes indicate that the equipment for making precision temperature
logs in a petroleum setting is more available than ever before, making it possible to add the
sort of information described here to the thermal aspect of basin analysis and to understand the
details of the thermal structure of sedimentary basins in a way not possible in the past.
(8) A low heat-flow value of 28±3 mW/m-2 was determined for the Willows-Beehive
Bend gas field in the Sacramento Valley of California. This value is consistent with the low
heat flow in the Sierra Nevada Mountains to the east and is the result of the residual thermal
effects of the subduction that ended in this area about 5 Ma.

ACKNOWLEDGMENTS

The logging and thermal conductivity measurements were supported by contracts from
Mobil Research Laboratories to the SMU Geothermal Laboratory. Robert E. Spafford did
most of the logging although some was done by John L. Steele. We thank Mobil Technology
Company for permission to publish these results.

REFERENCES

Auld, M.J., 1948, Temperature gradients for convection in well models: Jour. Applied Physics, v. 19, no. 2, p.
218.
Blackwel~ D. D., 1971, The thermal structure of the continental crust, in Heacock, 1. G., ed., The Structure and
Physical Properties of the Earth's Crust: Am. Geophys. Union Mon. 14, p. 169-184.
Blackwell, D.O., and Spafford, R.E., 1987, Experimental methods in continental heat flow, in Sammis, C.G., and
Henyey, T.L., eds., Methods of Experimental Physics: Geophysics, v. 24(B), Academic Press, New York,
p.189-226. .
Blackwell, D.O., and Steele, J.L., I 989a, Thermal conductivity of sedimentary rocks: measurement and
significance, in Naeser, N.D., and McColloh, T.H., eds., Thermal History of Sedimentary Basins: Springer-
Verlag, New York, p.13-36.
HIGH-RESOLUTION TEMPERATURE LOGS 33
Blackwell, D.D., and Steele, J.L., 1989b, Heat flow and geothermal potential of Kansas: Kansas Geol. Survey
Bull. 226, p. 267-295.
Blackwell, D. D., Murphey, C. F., and Steele, J. L., 1982, Heat flow and geophysical log analysis for OMF-7A
geothermal test well, Mount Hood, Oregon, in Priest, G.R., and Vogt, B.F., eds., Geology and Geothermal
Resources of the Mount Hood Area, Oregon: Oregon Dept. Geol. Mineral Indust. Spec. Paper. 14, p.
47-56.
Blanchard, P. E., and Sharp, Jr., J. M., 1985, Possible free convection in thick Gulf Coast sandstone sequences:
Am. Assoc. Petroleum Geologists, South West Section Trans, p. 6-12.
Bodner, D. P., and Sharp, J. M., Jr., 1988, Temperature variations in the South Texas subsurface: Am. Assoc.
Petroleum Geologists Bull., v. 72, no. 1, p. 21-32.
Brigaud, F. G., Chapman, D. S., and Le Douaran, S., 1990, Thermal conductivity in sedimentary basins
predicted from lithologic data and geophysical logs: Am. Assoc. Petroleum Geologists Bull., v. 74, no. 9,
p. 1459-1477.
Bristow, Q., and Conaway, J. G., 1984, Temperature gradient measurements in boreholes using low noise high
resolution techniques: Geol. Survey Canada Current Research, Pt. B, Paper 84-lB, p. 101-108.
Carter, L.S., Kelley, S.A., Blackwell, D.D., and Naeser, N.D., 1998, Heat flow and thermal history of the
Anadarko Basin, Oklahoma: Am. Assoc. Petroleum Geologists Bull., v. 82, no. 2, p. 291-316.
Counan, J., 1974, Time-temperature relation in oil genesis: Am. Assoc. Petroleum Geologists Bull.• v. 58, no.
12, p. 2516-2521.
Deming, D., Sass, J. H., Lachenbruch, and R. F. DeRito, 1992, Heat flow and subsurface temperature as evidence
for basin scale groundwater flow, North Slope of Alaska: Geol. Soc. America Bull, v. 104, no. 5, p.
528-542.
Demongodin, L., Pinoteau, B., Vasseur, G., and Gable, R., 1991, Thermal conductivity and well logs: a case
study in the Paris basin: Geophysical Jour., v. 105, no. 3, p. 675-691.
Diment, W.H., 1967, Thermal regime of a large diameter borehole: instability of the water column and
comparison of air- and water-filled conditions: Geophysics, v. 32, no. 4, p. 720-726.
Diment, W. H., and Urban, T. C., 1982, Temperature changes with time in the slotted interval of a deep, shut-in
geothermal well near thermal equilibrium: East Mesa well 31-1, Imperial Valley, California, 1977-1982:
Geothermal Resources Council Trans., v. 6, p. 249-252.
De Rito, R. F., Lachenbruch, A. H., Moses, T. H., Jr., and Munroe, R. J., 1989, Heat flow and thermotectonic
problems of the central Ventura basin, southern California: Jour. Geophys. Res., v. 94, no. Bl, p. 681-699.
Forster, A., Merriam, D. F., and Davis, J. C., 1998, Spatial analysis of temperature (BHTIDST) data and
consequences for heat-flow determination in sedimentary basins: Geol. Rundschau, v. 86, no. 2, p. 252-
261.
Forster, A., Schrotter, J., Merriam, D. F., and Blackwell, D. D., 1997, Application of optical-fiber temperature
logging--an example in a sedimentary environment: Geophysics, v. 62, no. 4, p. 1107-1113.
Funnell, R., Chapman, D.S., Allis, R., and Armstrong, P., 1996, Thermal state of the Taranaki Basin, New
Zealand: Jour. Geophys. Res., v. BIOI, no. 11, p. 25,197-25,215.
Gallardo, J.D., and Blackwell, D. D., 1999, Thermal model of the Anadarko Basin, Oklahoma: Am. Assoc.
Petroleum Geologists Bull., v. 83, no. 2, in press.
Galloway, W. E., and Cheng, E. S., 1985, Reservior facies architecture in a microtidal barrier system-Frio
formation, Texas Gulf Coast: Texas Bur. Econ. Geo!. Rept. Invest. No. 144, 36 p.
Gosnold, W. D., Jr., 1990, Heat flow in the Great Plains of the United States: Jour. Geophys. Res., v. B95, no.
1, p. 353-374.
Gretener, P.E., 1967, On the thermal instability oflarge diameter well--an observational report: Geophysics, v.
32,no.4,p.727-738.
Gretener, P. E., 1981, Using temperature in hydrocarbon exploration: Am. Assoc. Petroleum Geologists,
Education Course Note Series No. 17, 170 p.
Griffiths, C. M., Brereton, N. R., Beausillon, R., and Castillo, D., 1992, Thermal conductivity prediction from
petrophysical data: a case study, in Hurst, A., Griffiths, C.M., and Worthington, P.F., eds., Geological
Applications of Wireline Logs II: Geol. Soc. Spec. Pub!. No. 65, p. 299-315.
Gro/3wig, S., Hurtig, E., and KUhn, K., 1996, Fibre optic temperature sensing: a new tool for temperature
measurements in boreholes: Geophysics, v. 61, no. 4, p. 1065-1067.
Hales, A.L., 1937, Convection currents in geysers: Monthly Notices Roy. Astr. Soc., Geophysics Suppl., v. 4,
p.122-131.
34 BLACKWELL, BEARDSMORE, NISlllMORI, AND McMULLEN

Houbolt, J. J. H. c., and Wells, P.R.A., 1980, Estimation of heat flow in oil wells based on a relation between
heat conductivity and sound velocity: Geologie en Minjnbouw, v. 59, no. 3, p. 215-224.
Jaeger, 1. C., 1961, The effect of drilling fluid on temperatures measured in boreholes: Jour. Geophys. Res., v.
66,no.2,p.563-569.
Jeffreys, H., 1937, Notes on Mr. Hales's paper: Monthly Notices Roy. Astr. Soc., Geophysics Suppl., v. 4, p. 131.
Jessop, A. M., 1990, Comparison of industrial and high resolution thermal data in a sedimentary basin: Pageoph.,
v.133,no.2,p.251-267.
Johnson, K. S., Amsden, T. W, Denison, R. E., Dutton, S. P., Goldstein, A. G., Rascoe, B., Jr., Sutherland, P.
K., and Thompson, D. M., 1988, Southern mid-continent region, in Sloss, L. L., ed., Sedimentary Cover-
North American Craton, U. S.: Geol. Soc. America, The Geology of North America, D-2, p. 307-359.
Larimore, D. R., Goiggon, J. J., and Bayhn, R.I., III, 1997, Low-cost solutions for well interventions through
advanced slickline service: SPE 35236, 1996, quoted in Jour. Petrol. Tech., 118-122.
Lee, Y., Deming, D., and Chen, K. F., 1996, Heat flow and heat production in the Arkoma basin and Oklahoma
platfonn, southeastern Oklahoma: Jour. Geophys. Res., v. BIOI, no.ll, p. 25,387-25,401.
Majorowitz, J. A., Jessop, A. M., Jessop, c., and Deuma, M., 1998, Heat flow and subsurface temperature along
a SW-NE profile across the Western Canada Sedimentary Basin, this volume.
McGee, H. W., Meyer, H. J., and Pringle, T. R., 1989, Shallow geothermal anomalies overlying deeper oil and
gas deposits in the Rocky Mountain region: Am. Assoc. Petroleum Geologists Bull., v. 73, no. 5, p.
576-597.
McKenna, T.E., Sharp, 1. M., Jr., and Lynch, F. L., 1996, Thermal conductivity of Wilcox and Frio sandstones
in south Texas (Gulf of Mexico basin): Am. Assoc. Petroleum Geologists Bull., v. 80, no. 8, p. 1203-1215.
Nielsen, S. B., and N. Balling, 1984, Accuracy and resolution in continuous temperature logging:
Tectonophysics, v. 103, no. 1, p. 1-10.
Ponzini, G., Crosta, G., and Guidici, M., 1989, Identification ofthennal conductivities by temperature gradient
profiles: one-dimensional steady flow: Geophysics, v. 54, no. 5, p. 643-653.
Ramey, H. J., 1962, Wellbore heat transmission: Jour. Petrol. Tech., v. 14, no. 4, p. 427-435.
Robertson, E.C., 1988, Thennal properties of rocks: U.S. Geol. Survey Open-file Rept. 88-441, 106 p.
Ross, E.W., Vagelatos, N., Dickerson, J.M., and Nguyen, V., 1982, Nuclear logging and geothermal log
interpretation: fonnation temperature sonde evaluation, in Hallenburg, J.K., ed., Geothennal log
interpretation handbook: Soc. Prof. Well Log Analysts, Tulsa, Oklahoma, p. V7-V52.
Sakaguchi, K., and Matsushima, N., 1995, Temperature profile monitoring in geothermal wells by distributed
temperature sensing technique: Geothennal Resource Council Trans., v. 19, p. 355-358.
Saltus, R.W., and Lachenbruch, A.H., 1991, Thermal evolution of the Sierra Nevada: tectonic implications of
new heat flow data: Tectonics, v. 10, no.2, p. 10,325-10,344.
Sammel, E. A., 1968, Convective flow and its effect in temperature logging in small diameter wells: Geophysics,
v.33,no.6,p.l004-1012.
Williams, C. F., Galanis, S. P., Jr., F. V. Grubb, and T. H. Moses, Jr., 1994, The thermal regime of Santa Maria
Province, California: U. S. Geol. Survey Bull. 1995, Chapt. F, 25 p.
Wisian, K.W., Blackwell, D.D., Bellani, S., Henf1ing, J.A., Normann, R.A., Lysne, P.C., Forster, A. and
Schrotter, J., 1998, Field comparison of conventional and new technology temperature logging systems:
Geothermics, v. 27, no. 2, p. 131-141.
PROBLEMS AND POTENTIAL OF INDUSTRIAL TEMPERATURE
DATA FROM A CRATONIC BASIN ENVIRONMENT

Andrea FBrsterl and Daniel F. Merriam2


lGeoF orschungsZentrum Potsdam
Potsdam, Germany
2Kansas Geological Survey
University of Kansas
Lawrence, Kansas USA

ABSTRACT

A thin veneer of clastic and carbonate sediments 460-1,375 m (1,500-4,500 ft) thick
overlies a Precambrian crystalline basement complex in southeastern Kansas (Midcontinent,
USA). Well-log temperatures were analyzed and related to modeled temperatures for
different-size areas in a relatively simple structural setting of the shallow, cratonic Cherokee
Basin. A statistical analysis of bottom-hole temperatures (BHTs) confirmed that (1) there is
no significant change of temperature with season or through the 40 years of drilling and
logging the wells in the area; and (2) the distribution of values is normal indicating they
were recorded correctly on the rig within the instrumental precision of about 1 K. It was
obvious from the large data set of nonequilibrium BHTs analyzed by depth and stratigraphic
unit that they differed from drillstem test temperatures (DSTs) and modeled temperature-
depth distributions based on heat conduction. At shallow depth (less than 500 m), BHT
values as read from the logs are higher than 'true' formation temperature; in a depth range of
500-700 m, values scatter around a 'true' formation temperature; and at greater depths (up to
1,100 m), the uncorrected BHTs slightly underestimate the formation temperature.
Different amounts of data scatter in the composite BHT-depth plots occur in different size
areas, which also impacts the calculated geothermal gradients and empirical correction
methods developed on the basis of these data. Although the BHTs in the 60x75 km area of
Elk and Chautauqua counties scatter ±5-9°C around some mean value, seemingly the scatter
can be reduced slightly when working with smaller areas, for example 10xl0 km. A
mapping approach made to investigate the variability of the 60x75-km BHT cluster in more
detail showed part of the BHT scatter to be related to regional geology. Temperature
residuals of the trend are on the order of ±2.5°C with some values as large as ±7.5°C. This
provides an indication of variability of BHTs resulting from other influences, which might
be different perturbations as a result of drilling practices and different shut-in times of the

35
36 FORSTER AND MERRIAM

wells. In addition, some of the temperature highs are shown to be related to local, subtle
anticlines developed in the sedimentary cover over faulted basement blocks and therefore
contain signal. The separation of different effects on BHTs has implications on the
reliability of geothermal gradient and heat-flow density estimations.

INTRODUCTION

Thermal conditions in sedimentary basins are related to heat flow from the
underlying basement and the lithology and petrophysical parameters of the different
sedimentary units. The accurate assessment of both the thermal structure and heat-flow
distribution in basins is necessary to resolve processes, such as fluid migration, hydrocarbon
maturation, and diagenesis. One of the goals in heat-flow studies is to identify and quantify
the different modes of heat transfer within the sedimentary units.
Heat flow normally is determined from an average temperature gradient or a series of
formation or interval temperature gradients obtained from a high-resolution temperature log
made in a borehole under equilibrium in conjunction with an estimate of thermal
conductivity from rock samples of the sedimentary sequence. However, the only
temperature data usually available for basin analysis are bottom-hole temperatures (BHTs)
or temperatures from drill stem tests (DSTs) measured in commercial boreholes drilled in the
search for hydrocarbons. BHTs are measured in most hydrocarbon exploration wells, thus
constituting large databases, but the data represent only one measurement from one
particular depth.
Although the limitations of the data are understood, it is assumed that the lack of
high-resolution continuous temperature logs can be compensated adequately by good spatial
coverage of well logs with BHTs and a range of depths measured. Several studies
concerned with the methodology of handling and interpreting large BHT data sets have been
published in the last two decades (e.g. Majorowicz and Jessop, 1981a, 1981b; Majorowicz
and others, 1984, 1985; Lam and Jones, 1984; Drury, Jessop, and Lewis, 1984; Bachu,
1985; Chapman and others, 1984; Speece and others, 1985; Deming and Chapman, 1988;
Ben Dhia, 1988; Lucazeau and Ben Dhia, 1989). The general understanding is that BHTs
have to be corrected to equilibrium ('true') formation temperature if they are used for heat-
flow density estimates. But uncorrected BHT data have been used successfully under certain
conditions for regional investigations (Bodner and Sharp, 1988).
Several methods to correct BHTs to 'true' formation temperature are available. The
emphasis on these correction procedures has been either on the development of numerical
methods to correct BHTs or by modeling the temperature disturbance effects in the
borehole. Deming (1989) summarized the available correction methods and concluded that
none are completely satisfactory because in most instances the information needed to handle
the correction algorithm is not available. Hermanrud, Cao, and Lerche (1990) evaluated the
error involved in correcting BHTs collected in boreholes from an oil field by using different
numerical correction methods under known conditions. They concluded that advanced
models gave values closest to DST ('true') temperatures, but with large standard deviations,
whereas less advanced models gave temperatures too low also with large standard
deviations. The error was attributed to poor quality data and physical effects not accounted
for in the calculations.
Other studies were concerned with the use of inversion techniques to resolve the
spatial variability of corrected BHT data (Speece and others, 1985; Willett and Chapman,
PROBLEMS AND POTENTIAL OF INDUSTRIAL TEMPERATURE 37

1986, 1987; Deming and Chapman, 1988; Willet, 1990). Although the models seem to
mitigate the effects of BHT data noise, the lack of available high-resolution temperature
data does not allow judgement on what model best estimates actual temperature fields
(McPherson and Chapman, 1996).
Some of the variability in the raw BHTs (the variability as a result of different shut-
in time) obviously can be overcome by applying the proposed numerical correction
procedures. The general noise in the raw data and the unknown quality of the measurements
however are major problems in working with BHTs, especially when applying empirical
correction methods. Empirical corrections are based on the relation between the BHT and
'true' (undisturbed) formation temperature in the measured depth interval (for example the
AAPG correction method, Kehle 1972, 1973), but the original data scatter (error) will not be
improved, only shifted towards a higher value. Deming (1989) concluded that the true error
in raw BHT data is really unknown and that rigorous studies that compare true equilibrium
temperatures to BHTs derived from the logging of oil/gas wells are warranted.
A discussion on possible influences of data quality is provided by Speece and others
(1985). Some problems with the BHT quality are concerned with the way in which the
measurements were taken and the conditions under which they were made. There is no
evaluation as to the accuracy of the instrument nor recognition of the problems in obtaining,
recording, and filing the data. The instrumental precision of industrial temperature data
normally is about 1 K.
A single BHT value, even corrected to some formation temperature, is suspect, but
the problem of the inaccuracy can be reduced by using large BHT data sets. From oil and
gas fields tight clusters of corrected, as well as uncorrected, BHTs are known, which show
that the scatter of the BHTs is about 15 to 20°C. This suggests that the random noise in the
corrected BHT data may be on the order of ± 1OOC (Deming and others, 1990). A study in
the deep Northeast German Basin utilizing multiple BHTs (three or more BHTs measured at
the same depth, but at different shut-in times) that were corrected using the Bullard (1947)
equation and its simplification (the Homer plot) have shown for boreholes with multiple-
depth BHTs that the data after correction scatter around some mean value by about ± 10-
15°C (A. Forster and coworkers, unpublished data). In addition it can be shown that the
BHT correction surprisingly underestimates the equilibrium temperature known from
continuous temperature logs by about 5-15°C.
Considering the high variability in the data even after the correction, and that only
one corrected BHT value may be available for a well, it is necessary to know how many
BHTs are needed for (1) composite temperature-depth plots to obtain a meaningful
temperature-depth relation and geothermal gradient for an area or vice versa, and (2) how
large a geologic area can be to obtain a sufficient number of BHTs for composite plots with
no increase in the basic data scatter. In addition, investigations of possible sources of
variation other than depth in BHT measurements are desirable.
This study is a reconnaissance of commercial temperature data from the shallow,
cratonic Cherokee Basin in southeastern Kansas (Midcontinent, USA: Fig. 1) where the
Precambrian crystalline basement is overlain by a thin sedimentary veneer. In the U.S.
Midcontinent, archives of commercial wireline logs are available. Petrophysical logs vastly
outnumber high-precision temperature logs and data on subsurface thermal conductivity.
However, by integrating thermal measurements with commercial logs, a high-quality
database can be extended to viable synthetic thermal conductivity profiles for large regions
which then can be used in combination with commercial temperature data to determine heat
flow and evaluate the thermal structure.
38 FORSTER AND MERRIAM

c:=:J 'A1.EOZCJC 'ElSIC Vo..CAHIC ROCKS (St. (IIi. I I .


CJ PAlEOZOIC OJ4C.. ", fACC:S , ~ht ~ , . n)
A.,.. . t~ C"tUUYI 1.w~ I )

1:1=,:::'!f:t=,:::f:=:::':f:=~'YO ..... n

DASSM ENT Sl"I\UCTURE


0' t h~
MID. CONTINENT

Figure I. Generalized regional map of Midcontinent USA showing major basement structures and outcrop of
Precambrian rocks. Cherokee Basin in southeastern Kansas where studies were made is highlighted. Also
shown are Nemaha Anticline and Sedgwick and Forest City Basins (not identified) west and north of
Cherokee Basin. Contours on top of basement (subsea); CI = 1,000 ft (from Adler and others, 1971).

Here, we focus on the problems and limitations of working with single temperature
measurements in such a shallow sedimentary environment and to what extent available data
might add information on subsurface geothermal conditions using simple plotting and
mapping techniques. The area we have selected is one of several subareas in southeastern
Kansas of well-known geology where the structural situation is simple, which is a
prerequisite to learn about the general variability of commercial temperature data.
Figure 2 is an index map outlining the subareas in southeastern Kansas within this
framework that were investigated previously by us. Forster and Merriam (1993), utilizing a
map comparison/integration technique, and Forster, Merriam, and Brower (1993), utilizing
multivariate statistics, analyzed the usefulness of (1) average gradients from continuous
thermal logs at shallow depth, and (2) average gradients from uncorrected bottom-hole
temperatures (BHTs) obtained from depths deeper than the continuous temperature logs in
their relation to known geologic conditions. It was shown that the patterns of both types of
geothermal gradient maps, as originally published by Stavnes (1982), have only vague
similarity and that the interpretation of average geothermal gradients in terms of subsurface
geology can be misleading. To work on the methodology of interpreting BHT data in
southeastern Kansas, Forster and Merriam (1995) and Forster, Merriam, and Davis (1997)
analyzed in more detail a subset of available BHTs and DSTs as a basis for making
interpretations on a basinwide scale.
PROBLEMS AND POTENTIAL OF INDUSTRIAL TEMPERATURE 39

a SO MILES
I I
I
o 50 KILOMETERS

Figure 2. Map of eastern Kansas showing in detail subareas of fonner investigations: hatched = area of map
comparison/integration and cross sections for thennal modeling (Forster and Merriam, 1993); stippled = area
analyzed by multivariate statistics (Forster, Merriam, and Brower, 1993); cross-hatched = area ofBHTIDST
analysis (Forster and Merriam 1995; Forster, Merriam, and Davis, 1996; this paper); double outline = area of
thennal subsurface model (this paper).

GEOLOGICAL CONDITIONS

Southeastern Kansas is part of the stable US Midcontinent where Precambrian


crystalline rock is covered by a thin veneer of Phanerozoic sediments. This veneer consists
of Paleozoic strata ranging in age from Late Cambrian to Permian, capped by some
Pleistocene and Recent sediments (Merriam, 1963). Figure 3 shows the main stratigraphic
units, their general lithology, and approximate thickness.
The total thickness of the sedimentary veneer ranges from about 460-1,375 m (1,500
to 4,500 ft) (Fig. 4). Generally, the sediments dip gently westward 30 to 40 feet per mile
except where a reversal of dip occurs on local structures. The sediment veneer thins toward
the western slope of the Ozark Uplift located in western Missouri. Whereas the sedimentary
section in the east comprises mostly carbonates (Mississippian and Cambro-Ordovician)
because of the truncation, the section increases in clastic content to the west because of the
presence of the Permo-Pennsylvanian on top of the Lower Paleozoic carbonates. The
northeast-trending, southerly plunging Nemaha Anticline forms the western flank of the
basin and is recognizable by a thinning of sediments over the crest. A south-plunging
syncline subparalleling the eastern side of the Nemaha Anticline is the axis of the Cherokee
Basin, which is a northern shelf extension of the Anadarko/Arkoma basin complex located
farther south in Oklahoma (see also Fig. 1).
The regionally continuous Lower Paleozoic carbonates comprising the Mississippian
and the Cambro-Ordovician Arbuckle Group are parts of the Ozark Plateau and the Western
Interior Plains Aquifer Systems of the Midcontinent. As outlined in papers by Jorgensen
40 FORSTER AND MERRIAM

General
Stratigraphic age thickness I ft. rock type
Column

............ AHemating thick shale,


- lenticular sandstone, and
Permo-Penn 0- 2000 thin limestone in cyclic
7':--:.7.-::7-:: :- pattern.

~'~':.::'':'

.;-;- .. ..-:::;,.
7:'::+::';';:.~
-;-;
Mainly limestone with
some dolomite and shale,
200 -450 cherty in places. Upper
Miss-Dev
(0 -100) surface is karsted. Black
shale at base (Dev.l
///

Mainly dolomite wnh some


limestone. Sandstone at
Cambro-Ord 700 -950 base (Cambrian). Upper
surface karsted; lower
surface uneven.

~~
If'
v v· . ..,'"
Granite I rhyolne crystallines
v v v wnh some metasediments.
~,,'~'...v v"'.., PC + Faulted in block pattem.
,'..' ...' ...' v vv
1.1 -1.4 Ga
~;~;~; ;~~
',... ,","',',

Figure 3. Generalized stratigraphic column of units present in southeastern Kansas and their relative
thickness and lithologic type (3.28 ft = 1 m).

(1989), Macfarlane and Hathaway (1987), and Musgrove and Banner (1993), these two
aquifer systems have three regional patterns of groundwater flow. Groundwater in southern
Missouri migrates radially outward from the Ozark Uplift through the Ozark Plateau
Aquifer System; the resulting flow is westward into southwestern Missouri and southeastern
Kansas. To the south in Oklahoma, groundwater flows out of the Anadarko/Arkoma basin
complex northward into southern Kansas. In western and central Kansas, groundwater
migrates to the east and southeast through the Western Interior Plains Aquifer System.
These three regional flow components of groundwater meet and mingle in and near the
study area.

GEOTHERMAL CONDITIONS

Only two heat-flow measurements are published for southeastern Kansas (Blackwell
and Steele, 1989). The heat-flow density value determined in Miami County is 60 mWm-2 ,
the other value from Labette County yields 62 mWm2 • From high-precision temperature
logs in these boreholes and from additional investigations in boreholes in central Kansas, it
is known that formation temperature gradients differ vertically through a large range
PROBLEMS AND POTENTIAL OF INDUSTRIAL TEMPERATURE 41

because of the contrasts in thermal conductivity of the sediments. Blackwell and Steele
determined that the mean gradients in the carbonate sections of different boreholes were
nearly identical, averaging 17-21°Ckm·1 for those which are predominantly limestone
(Mississippian) and 13_17 o Ckm· 1 for those which include dolomite (Arbuckle Group).
Gradients in the predominantly shale sections (Pennsylvanian) range from 35 to 53°Ckm· 1
depending on the shale content. It also was recognized that the heat-flow density in the
boreholes is constant because thermal conductivity offsets the variation in formation
temperature gradients. Thermal-conductivity contrasts in the subsurface result in a ratio of
2.5 : 1 between limestone and shale and up to 4 : 1 between dolomite and shale. As a
consequence of the different lithology and thermal conductivity in the sedimentary
sequence, significantly different temperatures occur at the same depth in those parts of the
area that differ structurally and stratigraphically.

T
o
W
N
S
H
I
p

31

J1

JJ

J4

JSS
10 II 12 13 14 15 16 17 18 19 10 21 22 2J 24 lS I'!

RANGE
Figure 4. Total thickness of sedimentary section in southeastern Kansas (from top Precambrian to surface).
CI = 500 ft. Prominent Nemaha Anticline in western part of area; note thinning of stratigraphic section to
east. Area is subdivided into counties and townships (N-S) and ranges (E- W); township is 36 sq mi.

Because of the lack of measurements of radioactivity in the basement, the influence


geographically of a changing heat-production rate within the basement on the sedimentary
temperature changes can not be determined. However, no extreme changes in the heat-
42 FORSTER AND MERRIAM

production rate in the underlying basement rocks that would cause lateral temperature
changes in the area are suspected from the range of heat-flow density measured in eastern
and central Kansas. Also lateral temperature variations in the sediments resulting from fluid
flow and causing large-scale lateral transfer of heat are not evident in any of the aquifer
systems (for example, see Blackwell and Steele, 1989).
Geothermal information on southeastern Kansas and other parts of the Midcontinent
also is available from isotherm and gradient maps that have been published resulting from
an analysis ofBHT data (AAPG - USGS, 1976a, 1976b). These average gradients computed
on the basis of a two-point-straightline computation between BHT data and a mean-surface
temperature value are manyfold in their interpretation as already shown by Forster and
Merriam (1993) and Forster, Merriam, and Brower (1993) on two types of gradient maps
originated by Stavnes (1982) and therefore did not add information on the area not already
known from the few thermal logs measured.

MODEL OF THE THERMAL STRUCTURE

In areas where no, or only a few, subsurface temperature data are available and
steady-state heat conduction is assumed, temperatures as a function of depth can be
predicted using Fourier's law, q = k(dT/dz), where q is heat-flow density, k is thermal
conductivity, t is temperature, and z is depth. In our situation of known heat-flow density for
the area and sedimentary units with contrasting thermal conductivity, the subsurface
temperature should be modeled considering the thickness of the various subsurface units
and their characteristic thermal conductivity as:

n
T = To + q I (Z; / k) (1)
;=1

where Tis temperature eC), To is ambient surface temperature eC), z; (m) and k; (Wm· 1K· 1)
are thickness and thermal conductivity of the i th unit from the surface, and q (mWm·2) is
heat-flow density.
The equation was used in computing temperature-depth profiles at borehole locations
along two E - W cross sections (Fig. 2) (Forster and Merriam, 1993), and for the purpose of
this study, temperature-depth data at additional locations were generated to give a relatively
homogeneous data distribution for mapping. In such a manner, a thermal structure within
the sediments of southeastern Kansas, comprising all or parts of 32 counties from T 13 S to
T 35 S, and R 1 E to R 25 E, was developed that is dependent on the thickness and lithology
changes within the sedimentary section.
Table 1 lists the units (stratigraphic groups) in which the stratigraphic sequence of
Figure 3 is organized. To involve the units in the subsurface temperature modeling, a
lithotype was assigned to each group summarizing the percentages of lithology making up
the interval. Considering the uniformity of the stratigraphic section regionally, the
percentage of lithologies for each stratigraphic unit does not differ significantly in the area
modeled. Thermal-conductivity values as measured by Blackwell and Steele (1989) in four
boreholes in eastern and central Kansas were assigned to each lithotype by weighting the
values according to the thickness of lithology. In such a way, the model reflects the
conditions in the area in which lateral variation of measured thermal-conductivity values is
small in comparison to the vertical variation because of alternating lithological units.
PROBLEMS AND POTENTIAL OF INDUSTRIAL TEMPERATURE 43

Table 1. Thennal-conductivity data used in modeling.

Stratigraphic Unit Lithotype Thennal Conductivity


(Wm·1K· 1)
Sumner Group sh4/salt1 1.40
Chase Group sh11ls1 1.62
Council Grove Group Is3/sh1 2.05
Admire Group 1sl/sh1 1.70
Wabaunsee Group 1sllsh3 1.45
Shawnee Group 1s lIshb2 1.52
Douglas Group Isllsh3 1.70
LansinglKC Groups Isllshllsstl 1.90
Pleasanton Group sstllsh2 1.51
Mannaton Group Isllshllsstl 1.90
Cherokee Group sstllsh2 1.25
Mississippian units Islldol 3.33
Chattanooga Shale sh 1.10
Hunton Group Is 3.15
Maquoketa Shale sh 1.25
Viola dolomite do 2.92
Simpson Group shllsslllslldol 2.20
Arbuckle Group do 4.00
Reagan Sandstone ss 2.47
sh = shale, ss = sandstone, Is = limestone, sst = siltstone, do = dolomite, salt - salt, shb - black shale

A surface temperature of l3°C (ambient present-day surface temperature in Kansas)


and a heat-flow density of 60 m Wm'2 were used as input parameters. Figure 5 shows
exemplarily for one borehole that the modeled temperature-depth values based on these
input parameters coincide reasonably well with the actual measured temperature log (see
Blackwell and Steele, 1989, for the temperature logs). Differences between modeled and
measured temperatures that can be observed for example in the 500-m (1,650-ft) depth map
(Fig. 6) are caused partially by differences in the measured surface temperature (differs
between 13 and 15°C) compared to the input surface temperature for modeling (constant
l3 0 C). The temperature conditions at the 500-m depth level are 35°C (95°F) in the western
part of the area, but are 25°C (77°F) or lower in the eastern part. Differences in the
thickness and lithology (thennal conductivity) of the sediments making up the sequence
account for the different temperature conditions.
To illustrate the influence of sediment thickness/structural conditions on the
subsurface temperature in detail, the calculated temperatures were mapped separately for
four different subsurface (stratigraphic) units (Fig. 7); these are (from youngest to oldest):
(A) Pennsylvanian (LansinglKansas City); (B) top of Mississippian; (C) top of Cambro-
Ordovician Arbuckle; and (D) top of Precambrian. The Nemaha Anticline in the west
(compare with Fig. 4) is reflected by local temperature anomalies over the crest. The
anomalies are more pronounced on maps of the older units (Fig. 7B-D) and are an effect of
shallower depths of these stratigraphic units and a thinner Penno-Pennsylvanian section.
The thennal model in general shows that all stratigraphic units are characterized by a slight
temperature increase to the west and southwest which is the result of the westward regional
44 FORSTER AND MERRIAM

Temperature ( DC )
o 10 20 30 40 50 60
O~~~~~~~~~~~-r~~~~~

200

400

E
..c 600
aQ)
Cl

800

1000

1200

14ooL-----------~----~-----L----~----~

Figure 5. Comparison of measured (solid line) and modeled temperature-depth profile (+) from Butler
County. Measured temperature profile from Blackwell and Steele (1989).

dip of the units. Lowest temperature values on top of all units occur in the southeastern part
of the area. The temperature differences between the southeast and the southwest of all
maps are on the order of 40°C, except for the Pennsylvanian map (32°C) where the mean
ambient surface temperature for the outcrop area is l3°C.
In contrast, at a depth of 500 m (Fig. 6), the temperature difference over the area is
smaller and on the order of about 10°e. Again, higher temperatures occur in the
southwestern part of the area because of the relatively thick, low heat-conducting
Pennsylvanian clastic (shaly) sequence. Small anomalies of temperature highs and lows are
associated with the Nemaha Anticline as indicated on the stratigraphic temperature maps
(Fig. 7B-D). Temperature highs shown on the 500-m map are caused by thinning of the
high heat-conducting Arbuckle and Mississippian units.
From this regional temperature model for southeastern Kansas, we selected the area
of Elk and Chautauqua counties for the analysis of commercial well-log temperatures (see
Fig. 2). This area is not affected by pronounced structural features, and therefore the
modeled temperature for each stratigraphic unit increases steadily over the entire area with
increasing depth of the layers. There is no complication in the structure (see Fig. 4) nor in
the modeled regional temperature field.
PROBLEMS AND POTENTIAL OF INDUSTRIAL TEMPERATURE 45

o 100 miles
~I----------------~I----------~I
o 100 kilometers

Figure 6. Modeled temperatures at 500m depth. Temperatures at locations A-D are from temperature logs; at
locations A-C thermal-conductivity measurements were made (Blackwell and Steele, 1989); in parentheses are
the measured surface temperatures. CI = 1·C.

ANALYSIS OF WELL-LOG TEMPERATURES

The Data

Bottom-Hole Temperatures. The data for Elk and Chautauqua counties, which is an
area of about 60x75 km (36x45 mi), were compiled from well-log headers from the files of
the Kansas Geological Survey. BHTs were recorded on resistivity and spontaneous potential
logs, laterologs, density, and sonic logs. The data set comprises 637 BHT values, each value
taken at total depth at a single borehole site. Despite the high density of boreholes, in most
instances, available infonnation is limited. There is practically no infonnation on time since
the circulation of drill mud ceased and the time the BHT value was obtained. Also, there are
no multiple BHT recordings at the same depth, which would allow the application of a
numerical correction procedure based on the temperature recovery in the well. This is the
typical situation when working with BHTs in a mature hydrocarbon province such as the US
Midcontinent.
BHTs as reported in this study, can be likened to working with well data such as
fonnation tops reported on scout tickets, which are widely and successfully used in the
46 FORSTER AND MERRIAM

CJ <20'C

CJ 20' 28

[]]I] 28·36

IillIllTIiiil 36' 44
44·52

> 52 aC

Figure 7. Contour maps of modeled temperature on top of different stratigraphic horizons in southeastern
Kansas. CI = 4°C. (+) stands for well sites that were modeled. A, Temperature on top of Pennsylvanian; B,
temperature on top of Mississippian; C, temperature on top of Arbuckle Group; D, temperature on top of
Precambrian.

petroleum industry. Scout tickets may be all that is available when lacking well-log
infonnation. Scout tickets may have errors in the reported depth of tops, misidentification of
stratigraphic units, incorrect well elevations, etc., but in most instances these errors are
negligible or obvious. IfBHTs are used in a regional study, then individual measurements
are important only in contributing to the general pattern. Adjacent, similar values fonn part
of the pattern and erroneous individual values usually are obvious and appear as anomalous
values. Values deemed incorrect can be given less, or no, weight when interpreting the
pattern. This is the same procedure geologists use when using scout tickets or other well
data of unknown quality for structural mapping. We acknowledge that BHTs also may be
PROBLEMS AND POTENTIAL OF INDUSTRIAL TEMPERATURE 47

incorrect in value or depth, but can be screened for useful information prior and after
applying any empirical correction factor.
In our approach, we collected the BHTs and coded them according to the three major
stratigraphic units of relatively homogeneous geology: (1) the Pennsylvanian; (2) the
Mississippian; and (3) the Cambro-Ordovician Arbuckle. The homogeneous nature of the
lithology is obvious for the Mississippian and Arbuckle carbonate units. Because of the
thin-bedded nature of the Pennsylvanian section comprised of alternating clastic and
carbonate rocks (see Fig. 3), the entire sequence can be considered homogeneous
representing a typical mean geothermal gradient derived from logging on the order of
38±6°Ckm· 1•
To test the possible range of subsurface temperatures for each of the three
stratigraphic units in the area we applied two empirical correction factors: (1) the
ForsterlMerriamlDavis (1996) correction factor; and (2) the average AAPG correction factor
(Kehle, 1972, 1973). Maps made using original raw data and the two corrected BHT data
sets are shown in Figure 8A, 8B, and 8C. The patterns are almost identical indicating the
differences are merely shifts in average values. All three maps were contoured using the
same algorithm and are directly comparable. The FIMID correction changes values from 0
to 3SC with larger changes in the west and small changes in the east (Fig. 8D) in the
shallower part of the basin. The AAPG correction ranges from 1.5 to 2SC with a gradual
increase to the west (Fig. 8E). This correction factor was designed for deeper measurements
and has little effect on the values in the shallow Cherokee Basin, whereas the FIMID
correction was constructed specifically for the Cherokee Basin and provides the 'best'
empirical correction (Forster, Merriam, and Davis, 1996).

Drillstem-Test Temperatures. In processing available well-log data, it was obvious,


as expected, that the temperatures recorded in drillstem tests usually were slightly higher
than those recorded as BHTs (maximum mud temperatures). Vik and Hermanrud (1993)
reported that temperatures from drillstem tests with high flow rates and low pressure
drawdown can be expected to yield values close to 'true' formation temperatures. They
applied a t-test statistic to their two subsets of data and determined that the BHT and DST
data sets were independent and could not be integrated without a correction factor. This,
also not surprisingly, is the situation for the Cherokee Basin BHT and DST data sets.
On the whole, 88 temperatures from drillstem tests (DSTs) were collected from the
same three stratigraphic units as the BHTs. Only those values from the tests which showed
oil or water inflow into the borehole were used. Excluded were those temperatures
measured mostly in mud which do not represent 'true' formation temperature (Forster and
Merriam, 1995).

IDENTIFICATION OF SIGNAL AND NOISE

A statistical test of month differences in BHTs showed there is no significant change


of temperature with season. BHTs were measured for a 40-year period and show a slight
increase of about 1°C through this time. The increase may result from (1) drilling to test
deeper targets; (2) a change from drilling with cable tools to rotary rigs; (3) a technical
improvement in the BHT measurement tool; or (4) unknown factors. Although drilling
deeper will result in higher BHT measurements, the regression uses depth as an independent
variable so this factor already is considered; there is an increase with time that cannot be
attributed to deeper wells. The most likely cause for the increase in BHTs may
48 FORSTER AND MERRIAM

D E

Figure 8. Maps showing pattern of: A, uncorrected Mississippian BHT data; B, FIMID correction applied to
BHT data; C, AAPG average correction applied to BHT data; A, B, and C contoured with same algorithm; CI
= 2°C. D, Difference in uncorrected and FIMID corrected values; E, difference in uncorrected and AAPG
average corrected values; CI = O.5°C. Configuration shows influence of westward regional dip and local
structure. See text for explanation.
PROBLEMS AND POTENTIAL OF INDUSTRIAL TEMPERATURE 49

be changes in drilling practices through time. The frequency with which specific
temperature values were recorded was tested to see if some temperature, such as 100°F,
occurred with unusual frequency. However, the distribution of values are normal suggesting
that the values were recorded correctly on the rig.

Composite TemperaturelDepth Plots

In general, it should be possible to calculate an average temperature gradient on the


basis of a BHT/depth plot if temperatures are well distributed along the depth profile. To get
a statistically meaningful result, it is practice to generate a composite temperature/depth plot
and work with confidence bounds for the BHT data scatter. A disadvantage of using
composite plots can arise when the basic scatter in data (noise) is enlarged by signal from
changes in structural conditions in the area.
To illustrate the problems, we plotted the BHTs from Elk and Chautauqua counties
versus depth separately for each stratigraphic/lithologic unit (Fig. 9A-C). Our area of 60x75
km from which the BHTs are compiled for example is larger than the area of 30x30 km
used by Jessop (1990) for comparison of uncorrected BHTs and equilibrium temperature
logs in the Western Canada Sedimentary Basin. Each of our BHT data sets shows a large
scatter in values with maximum departure from a mean value on the order of about 10°C.
For comparison, Jessops' BHT values, even from a smaller area also scatter by the same
amount determined by data regression. These scatter plots alone would confirm previous
results on the variation of BHT data as reported by Deming and others (1990). However,
the modeled temperature-depth profiles related to the plots (Fig. 9A-C; see also the section
on modeling) reflect that in our situation some of the variability in the BHT values is a
result of the gentle dipping of the layers in the area and a thickening of the Pennsylvanian
units towards the west.
The modeled temperature-depth profiles El and Cl (Fig. 9A-C) reflect the thermal
conditions in western Elk and Chautauqua counties in a comparable structural situation. The
profile E2 represents eastern Elk and C2 eastern Chautauqua County. The changes in the
stratigraphic section from east to west are indicated by the break in the temperature-depth
curves caused by high heat-conducting Mississippian and Arbuckle carbonates, which in the
eastern part of the area is located at shallower depths. On all plots it can be observed that
some BHTs are higher than the 'true' subsurface temperature from modeling. This might be
an effect of the used surface temperature of l3°C. If a value of 15°C would be used, as it
was observed by continuous temperature logs at some locations, the entire temperature-
depth curve would shift towards a higher value by 2°C and better approach the upper bound
of the BHT scatter.
Pennsylvanian BHTs (Fig. 9A) show a scatter of 12-18°C. At about 500 m, the BHTs
are near 'true' formation temperature; at depths shallower than 500 m, the BHTs are higher;
and at depths greater than 500 m, they are slightly below the modeled formation
temperature. In this depth range almost all BHTs are from the western part of the area and
from slightly deeper depth and therefore have to be related to profiles El and Cl. The
abnormally high values at shallow depths give rise to a couple of questions. Normally, by
rotary drilling a temperature obtained at the bottom of a well shortly after cessation of
drillmud circulation will result in a value lower than the equilibrium formation temperature.
So the BHTs in the shallow subsurface might have been affected by other influences, which
are not defined yet.
50 FORSTER AND MERRIAM

Temperature rC) Temperature ('"C) Temperature rC)

,0 ,S 20 25 30 35 40 45 50 ,0 ,S 20 25 30 3S 40 45 50 ,0 'S 20 25 30 3S 40 45 50
0 0 0

,
.
__ .l __ __ L __ _________
200 200
_C~ ~

0 200
!: OC
6
: c fl'
400 --~--~---~c -" ~- - - - - ---- 400 400
, ;'lI
000

I : ~ I
' I I 0 I
~ 600 --7---'---'---0
c!l ,, ~ 5
0
c
600
~c 600

eo'
-
o Penn. ~ ___ ~ __ ~ __
600 - - 800 800
-E,
-C,
,000 -E2 -- - --- '000 -E2 '000
-C2 -C2

'200
-'CO, -'CC, ,200
'200
A B C

Figure 9. Plots ofBHTs versus depth for Elk and Chautauqua counties in comparison with modeled (C1, E1,
C2, E2) temperatures and temperature profile (lCQ1) in shallow depth logged by Stavnes (1982). A,
Pennsylvanian BHTs; B, Mississippian BHTs; C, Arbuckle BHTs. Data are compiled from area about 60x75
km (36x45 mi).

Mississippian BHTs (Fig. 9B) are mostly from the central and western part of the
area with a few from the eastern part (shallow depth). These values scatter around the 'true'
formation temperature. BHTs from the central and western part plot in the depth interval
where the two sets of curves spread apart. Again, it is observed that the BHTs are scattered
by about 18°C. The scatter in the data masks any relationship with the generally assumed
trend of temperature disturbance by drilling.
Arbuckle BHTs (Fig. 9C) also are mostly from the central and western part of the
area and therefore have to be related mostly to curves El and Cl. Again as observed with
the Mississippian data, the BHT values generally fit in the envelope given by the two sets of
curves, whereby the values from the west (from deeper depth) logically are more reduced
and thus lower as the formation temperature indicated by the corresponding El and Cl
curves. BHTs from the eastern and central area (depth interval 550-800m) are partly in ex-
cess of the modeled temperatures (E2 and C2) and partly below.
The formation geothermal gradients based on regression analysis are about 16°Ckm· 1
for all three stratigraphic units (Table 2), but a poor fit is indicated by the correlation
coefficient. Whereas, the regressed BHT gradients in the Mississippian and Arbuckle car-
bonates fall in the range of those gradients observed by thermal logging under equilibrium
conditions, which are 21±3°Ckm-1 and 15±I°Ckm-1, respectively (see Forster and others,
1997), the gradient in the Pennsylvanian is far too low compared with the equilibrium
gradient of 38±6°Ckm-l . The Pennsylvanian is the uppermost part of the sequence in
southeastern Kansas and the regression line calculated for this unit should intercept the X
axis of the plot at an ambient surface temperature of about 13°C (Fig. 9A). The actual value
however is about 24°C indicating that the BHTs in the upper part of the sequence (at a depth
of 200-500 m) are higher than 'true' formation temperature resulting in a temperature
gradient that is too low. To analyze how robust those formation gradients are in terms of
the number of data points and the depth range from which they are compiled, we computed
formation gradients of BHTs from an interval of 500-800 m (Table 2); this is the depth from
which most BHTs are available. The gradients obtained for this depth range differ slightly
PROBLEMS AND POTENTIAL OF INDUSTRIAL TEMPERATURE 51

from the previous gradients comprising data from the entire depth range: the highest gra-
dient is for the Arbuckle Group (21.6°Ckm- 1) followed by the Mississippian (18_7°Ckm- 1)
and Pennsylvanian (13.0°Ckm- 1). In both situations, the order of formation gradients is a
mismatch with the order of gradients known from continuous temperature logging. Also it is
shown that the gradients do not reflect the expected disturbance effect on the raw BHT data
because of drilling. As a result from plotting the data from different depth intervals it can
be summarized that the general data scatter provides equivocal formation gradients.

Table 2. Statistical data from BHTIDST-depth plots.

BHT BHT DST

Unit Depth: 0-1100m Depth: 500-800m Depth: 0-1100m

no.point sd roc gradient no. point sd roc gradient no. sd roo gradient

Pennsylvanian . 141 4.0 0.60 16.0 65 3.4 0.28 13.0 40 4.5 0.95 32.8

Mississippian 313 3.4 0.45 15.7 251 3.3 0.37 18.7 38 3.3 0.78 24.2

Arbuckle 183 3.3 0.52 15.9 119 3.3 0.45 21.6 10 4.0 0.97 41.3

sd = standard deviation, rc:c = correlation coefficient

In a next step to investigate the sensitivity of the data scatter and its impact on the
formation gradients, composite temperature-depth profiles from several different-size areas
were analyzed. We determined that in our particular geological situation a reference area of
one township (10xlO km; 6x6 mi) is too small to obtain a meaningful relation of
temperature with depth (Fig. 10). In this situation the data scatter may be large compared to
depth difference. Moving into larger areas of 3x3 townships (30x30 km; 18x18 mi), it was
observed that the composite BHT-depth relations plotted for two different 30x30-km
clusters give different results in terms of formation gradients, but also in terms of mean
gradients averaged over the entire borehole depth. The difference in average gradient
obtained from the two clusters was on the order of 8°Ckm- 1, which amounts to 21 % gradient
change compared to mean gradient known from logging. The number and distribution of
data points available in the different areas also affect the correlation coefficient.
Temperatures measured during drillstem tests in Elk and Chautauqua counties are
plotted in Figure llA. Where the data points are from small, producing oil fields (about
1I2x1l2 mi), only 1 or 2 values from the remainder of data for each field were used for the
analysis. The DST data, which are mostly from the western part of the area better match the
modeled temperature profiles (E1 and C1) than the BHTs (compare with Fig. 9), but in
general are slightly higher. This might be an indication that the modeled temperature
profiles are in error by about 2°C as a result of the surface temperature used in modeling as
it was reported previously in this study. However, the overall match between DSTs and
modeled temperature is most striking for the Pennsylvanian where the DST scatter is less
than 8°C at any depth. The DSTs in the Mississippian and Arbuckle carbonates show a
scatter of about 12°C which is comparable to the scatter of BHT data. Results of the
regression analysis of DSTs for all three stratigraphic/lithologic units are shown in Figure
lIB. The formation gradient indicated by the regression line is highest in Arbuckle
dolomites and lowest in Mississippian limestones (Table 2). Although the DST gradients in
the Pennsylvanian and Mississippian are about in the range of temperature gradients
52 FORSTER AND MERRIAM
Temperature (OC) Temperature ("C)

15 20 25 30 35 40 45 50 15 20 25 30 35 40 45 50
O~~~~~~~~~ O~~~~~~~~~

1
I 1
I I I j I I 1
_ L __1__ 1. __ 1__ .1 __ _1.. __ 1_ _ ...i __
200 I I I I I
200 1 I 1

1 1
1 1 I I 1

400 --1--t-- 400 --l--+-- 1- - +- -1- - -f - -


1 1 I 1 1 1
I I 0 I I
1 1
I I I I
_J __ !...~_I_Q _
1 1 1 600

.
I

800 . -~--~--:--~-:- 800


I I I I Q I
10 1

1000
D

• Miss.
Penn.
f» 0: 1000 • Miss. - -;- - ~
I
- -:- - ~ --
o Arb. o Arb.
-Cl -Cl

A B
Figure 10. Temperature/depth plots for two townships in Chautauqua County. A, depth/plot for T 32 S, R8 E
has data points from area with no data from oil fields; B, township T 32 S, R 9 E contains Hylton oil field
with anomalously high temperatures that give large data spread for both Mississippian and Arbuckle. This
example demonstrates differences that can be obtained by different data sets in similar geologic settings.

Temperature rC) Temperature rC)

10 15 20 25 30 35 40 45 50 55 60 65 10 15 20 25 30 35 40 45 50 55 60 65
O~~~~~~~~~~~~~~ O~~~~~~~~~~~~~~~

,, ,,,
,, ,
---:, , ,,
.._-, --I -1--
200 --f-- , .. --'

, , 200 -T---,---r"--l---,---r
I I I I I I
"-II
,,, ,, I I I I I
" ,,
,, , ,, ,,
"
"
, "

.-r'. --:-,
1 I, I I I I
_1 ___ L __ J. __ J ___ L __ .1 __ _
400 --t--~---~
,, ,, ,, -" --~- -{--
,, 400 I I I I I I

.: , I : :
,
§: §:
600 ~ 600
~
Q)
0
DST Gl
0 ,,
-E1 ,, , , , ,,
,, ,,
I I I I I I

, , ,,
600 -C1 800 - -+- - -/- - -1--- -+-- -1- - - l - ---1---1----4---
I I L I I I
' ,
,,,
I t I I I ,0
-E2 : : : .i
-C2 ,
~ ~
DST :
I
,, ,,
: :

: : : I
, ,
. t---i---
I I I I I
1000 o Penn. --~---~--~--
, , 1000 -e-Penn. -i--~---:---i--~-
,' ,, ,
Miss.
,,
-Miss. : :
,
::
"
,,,
o Arb. ..... Arb.
, I
, ,
::

1200 1200

A B
Figure II. A, Plots of DSTs versus depth for Elk and Chautauqua counties for Pennsylvanian, Mississippian,
and Arbuckle units compared with modeled (EI, CI, E2, C2) temperatures; B, data analyzed by linear
regression. Statistical data given in Table 2.

obtained from high-precision temperature logging, the obtained temperature gradient for the
Arbuckle carbonates of 41 °Ckm- I is too high compared to the logging results (13-17°Ckm- I ).
The misfit of the DST Arbuckle gradient and temperature gradient from logging is
PROBLEMS AND POTENTIAL OF INDUSTRIAL TEMPERATURE 53

surprising considering the high correlation coefficient of the DST data, but could be the
result of too few data points.
The observed results in the BHTIDST-depth plots can be summarized as follows:
(1) Uncorrected BHTs from shallow depths (less than 500 m) reflect temperature
conditions that are higher than undisturbed formation temperatures. In a depth range of
500-700 m, the BHT values scatter around a 'true' formation temperature. At greater depths
a cooling effect on BHTs can be observed. The disturbance effect can be seen best on a
BHTIDST depth plot (Forster and Merriam 1995).
(2) The composite BHT-depth plots for the three stratigraphic units show a total
scatter of about 18°C. Some of this scatter is related to background, some is the result of
different adjustment to thermal equilibrium, and some seemingly is the result of regional
structure. Different BHT subsets show different statistical characteristics in the temperature-
depth plots depending on area size and number of data points. Consequently, each data set
reflects different temperature gradients for die stratigraphic units as well as the entire depth
interval.
(3) As was expected, the DSTs better reflect 'true' formation temperatures and
apparently are less variable than the BHTs. The 'best' correlation with depth is observed in
the relatively thick Pennsylvanian and consequently, the DST gradient is closest to the
formation gradient obtained from logging.
(4) We have observed in our study area that both, BHTs and DSTs are well suited
for determining mean temperature gradients when averaged over some depth interval of at
least several hundred meters. Even with a reduced DST scatter compared to the scatter of
uncorrected BHTs, the variability of the values does not allow to calculate formation
gradients unequivocally.

Temperature Maps

To analyze the temperature distribution in the different stratigraphic/lithologic units


in more detail and delineate local and regional trends that might provide hints for signal in
the data, the BHTs were mapped separately for the three different stratigraphic units (Fig.
12A-C). The maps help to interpret the overall BHT data scatter observed in the
temperature-depth plots and identify the basic error in the data.
Although all three maps show a similar overall pattern, the distribution and amount
of data differed for each map. A general trend of increasing temperatures towards the west
can be observed on all three maps. This trend is controlled by the increased depth of the
westward-dipping stratigraphic units. The pattern of the BHT maps however is not as
regular as observed on the modeled temperature maps (Fig. 7A-C).
Data for the Pennsylvanian map (Fig. 12A), comprising a depth interval from about
200 to 800 m, are scarce compared to the other two maps. Because of BHTs are scarce in
the northern part of the area this part of the map cannot be interpreted. The data over the
remainder of the map range from 21 to 43°C with most values in the range of25 to 35°C. A
few small hot spots with temperatures from 35 to 43°C occur in the western part of the area.
The pattern on the Pennsylvanian BHT map is difficult to interpret because of the
considerable thickness of the unit and different depths from which the BHTs were taken.
This is because most of the Pennsylvanian is an exploration target for both structural and
stratigraphic traps.
The MiSSiSSippian map (Fig. 12B) comprising a depth interval of about 450 to 900 m
is the most detailed because the most data are available for that unit. The Mississippian is a
54 FORSTER AND MERRIAM

major oil and gas producer in this part of Kansas, and therefore the large number of
boreholes drilled to this unit. For the Mississippian as well as the Arbuckle units, the BHT
values all are compiled from near-top of the unit. The exploration target for these units are
hydrocarbon traps linked with local structural features. Although the data are from near top
of the units, the contours of BHTs on the Mississippian are highly irregular showing an
alternation of highs and lows especially in the central and western part where most of the
boreholes are located. The lateral change in temperature on the map ranges between 25 and
45°C. The regional trend however shows an increase in values from 32°C in the east to 38°C
(~T=6°C) in the west. For comparison, the temperature model for the Mississippian surface
(Fig. 7) shows for the same area a more pronounced increase of temperature from about 28°
to 44°C (~T=16°C), indicating that the BHTs in the west (in the deeper part of the basin) are
somewhat cooler than the near-equilibrium conditions, but in the eastern (shallower) part
may not deviate from them. Temperature anomalies as isolated by trend analysis are in the
range of ±2SC with a few values of ±7.5°C. The anomalously high temperatures in excess
of 38°C are scattered mostly over the central and western part of the map. The remainder of
the map in the southeastern and northern part is devoid of data so that the smooth pattern of
low values there is not reliable.
Values for the Arbuckle Group overlying the Precambrian basement (Fig. 12C) are
from a depth interval of about 550 to 1050 m and range from 26°C (eastern part) to 43°C
(western part). This temperature range is the same as recognized for the overlying
Mississippian. The map has the best spatial data distribution of all three maps. Again, as
outlined for the Mississippian, there is an overall trend of increasing values from 32°C in
the east to 40°C (~T=8°C) in the west. For comparison, modeled temperatures range from
32° to 46°C (~T=14°C). Several hot spots of temperatures between 37° and 43°C are scat-
tered over the central and western part of the map. The values of hot spots on the
Mississippian BHT map are in about the same range as the maximum values in the Ar-
buckle.
The regional trend of increasing BHT values towards the west can be related visually
to the westward dip of the stratigraphic units as shown on a configuration map on top of the
Mississippian (Fig. 12D). Although the top of the Mississippian is a karst surface, at this
scale and contour interval the pattern reflects the regional structure. In addition to the
regional trend, a series of small high and low features is visible. These local features are
elongated northeast-southwest and northwest-southeast, reflecting the Precambrian
basement fracture/fault system. These local features, known as 'plains-type folds,' were
developed by differential compaction over the buried fault blocks (Merriam and Forster
1994, 1996).
From a visual comparison of the Mississippian BHT map (Fig. 12B) with the
Mississippian structural map (Fig. 12D), it can be seen that the alignment of small highs and
lows on the temperature map semiparallels the grain of the structural pattern. It also is
obvious that some of the anomalies of higher temperatures locally occur coincident with
structural highs, for example in the southwestern and south-central part of the area. The
higher temperature at shallower depth is in contradiction to the general prediction in the area
that temperature increases with increasing depth of a unit.
A detailed study of several subtle anticlines in the area substantiates the relationship
of higher temperature and local positive structures in the Lower Paleozoic carbonates (Fig.
13). Temperature anomalies on the order of about 7°C (15°F) occur in both the BHT and
the DST data sets. These anomalies are not a reflection of variation in
stratigraphic/lithologic properties within these subtle structures, and if not a result of noise,
PROBLEMS AND POTENTIAL OF INDUSTRIAL TEMPERATURE 55

D<35 0
C
A B

25 Miles
i' ,

25 Kilometers

Figure 12. Contour maps of bottom-hole temperatures (BHTs) recorded in Elk and Chautauqua counties; CI =
SoC. A, BHTs recorded in Pennsylvanian. B, BHTs on top Mississippian. C, BHTs on top Arbuckle Group;
and D, structural contour map on top of Mississippian rocks in Elk and Chautauqua counties; CI=lOO ft.
Datum: sealevel.

could be a reflection of a change in the reservoir conditions in the carbonates. The


contained hydrocarbons in the anticlines in conjunction with fluid movements along the
developed fracture system on top of those structures may cause the temperature anomalies.
56 FORSTER AND MERRIAM

SUMMARY OF RESULTS

The BHTIDST analysis of our area demonstrates the problems of working with
temperature data from well logs to obtain information on the subsurface temperature
distribution and temperature gradients. The major problem with single temperature data is
their reliability. In our situation, a lack of data on circulation time and on the more
important shut-in time does not allow temperature corrections fur drilling disturbance, but it
is questionable whether such a correction actually would improve the data resolution.

o 1 mile
~I----------~i------~i
o 1 kilometer

32

S
11

14
R 9 E
Figure 13. Hylton oil field in Chautauqua County showing structure on top of Mississippian (CI = 10 ft, 3.3
m) and DSTs (CI = 5°F, 2.8°C). Anticline has closure of'" 40 ft (12 m). Well control shown by solid circles;
wells with DSTs shown by solid diamonds. Temperature difference on structure'" 15°F (7°C).

Although we have checked the BHT data carefully for several influences that might
affect their value during measurement, none of them actually explain or reduce the
variability (basic noise) inherent in the data. Therefore, clusters of BHT data are needed to
average values and extract a meaningful temperature/depth relation, which then can be
corrected to formation temperature using empirical methods. However, different data
scatter obviously occur in different size areas, which also impacts the geothermal gradients
that are obtained. A data scatter ofBHTs around some mean value of ±5-9°C was observed
in the 60x75 km area comprising Elk and Chautauqua counties. Seemingly, the data scatter
can be reduced to ± 7°C when working with composite BHT-depth plots for smaller areas of
lOxlO km. But this requires good data coverage, which in most situations is not available.
The disadvantage of working with small areas also can be that the data do not have vertical
resolution because of clustering at about the same depth. There is no way to determine from
the scatter plots how much scatter actually is caused by different adjustment of the BHTs to
steady-state equilibrium temperature. The DSTs, which in general are closer to true
PROBLEMS AND POTENTIAL OF INDUSTRIAL TEMPERATURE 57

formation temperatures, seemingly are less scattered and may contain less error than raw
BHT data (see Forster and others, 1997).
We used a mapping technique to investigate the variability of BHTs as shown in the
scatter plots in terms of signal and noise. We attribute signal in the BHTs to be related to
regional and local geological conditions in the area. By separating the BHTs according to
their stratigraphic unit and plotting their spatial distribution, it is obvious that, in addition to
a regional trend of increasing temperature with increasing depth, local anomalies complicate
the temperature pattern. Thus, for example, the regional BHT increase from east to west
amounts 6°C on the Mississippian surface. The internal 'noise' not related to regional
geology (here the depth change of the subsurface) is on the order of ±2.5°C with a few
values of ±7 .5°C as indicated by the residuals of the trend. This amount of noise is less than
would be obtained from BHT scatter plots, where the regional geological variation is not as
obvious.
Part of the anomalous values (some of the temperature highs) are shown to be
correlated with local, subtle anticlines developed in the sedimentary cover over faulted
basement blocks and contain signal. Moreover, from the BHT trend being coincident with
the trend of regional structure it also can be concluded that for deeper units below 500 m
depth the amount of cooling in the BHTs seemingly is correlated with depth and that BHTs
measured at the same depth have undergone about the same disturbance during drilling.
This implies that anomalies on the map not related to local structure could be a result of
different shut-in times compared with the remainder ofBHT data, but this cannot be proved.
The mapping approach we used in our study is a valid tool to search for the
variability in the industrial temperature data to delineate and emphasize local influences.
The search for spatial properties in the data set helps identify and select proper values to be
used in temperature-depth plots and thus to improve the data resolution prior to the
application of empirical correction procedures, which do not affect the statistical properties
of the data set. The separation of different effects on BHTs is a necessity to improve the
reliability of geothermal gradient and heat-flow density estimations using these data.

ACKNOWLEDGMENTS

We would like to thank Lynn Watney and David Newell of the Kansas Geological
Survey and Phil Armstrong of the University of Utah for reading a preliminary copy of the
manuscript and making helpful suggestions. John Davis of the Survey did the statistical
analysis and helped with the interpretation.

REFERENCES

American Association of Petroleum Geologists and U.S. Geological Survey 1976a, Subsurface temperature
map of North America: U.S. Geoi. Survey Map, 1:5,000,000.
American Association of Petroleum Geologists and U.S. Geological Survey, 1976b, Geothennal gradient map
of North America: U.S. Geoi. Survey Map, 1:5,000,000.
Adler, F.J., and others, 1971, Future petroleum provinces of the Mid-Continent, in Cram, I.H., ed., Future
Petroleum Provinces of the United States - Their Geology and Potential: Am. Assoc. Petroleum
Geologists Mem. 15, v. 2, p. 985-1042.
Bachu, S., 1985, Influence ofiithology and fluid flow on the temperature distribution in a sedimentary basin:
a case study from the Cold Lake area, Alberta, Canada: Tectonophysics, v. 120, no. 3-4, p. 257-284.
Ben Dhia, H., 1988, Tunisian geothennal data from oil wells: Geophysics, v. 53, no. 11, p. 1479-1487.
58 FORSTER AND MERRIAM

Blackwell, D.D., and Steele, J.L., 1989, Heat flow and geothermal potential of Kansas: Kansas Geo!. Survey
Bull. 226, p. 267-291.
Bodner, D.P., and Sharp, Jr., J.M., 1988, Temperature variations in South Texas subsurface: Am. Assoc.
Petroleum Geologists Bull., v. 72, no. 1, p. 21-32.
Bullard, E.C., 1947, The time necessary for a borehole to attain temperature equilibrium: Royal Astronomical
Monthly Notices, Geophys. Supp!., v. 5, no. 5, p. 127-130.
Chapman, D.S., Keho, T.H., Bauer, M.S., and Picard, M.D., 1984, Heat flow in the Uinta Basin determined
from bottom-hole temperature (BHT) data: Geophysics, v. 49, no. 4, p.453-466.
Deming, D., 1989, Application of bottom-hole temperature corrections in geothermal studies: Geothermics, v.
18, no. 5-6, p. 775-786.
Deming, D., and Chapman, D.S., 1988, Inversion of bottom-hole temperature data: the Pineview field Utah-
Wyoming thrust belt: Geophysics, v. 53, no. 3, p. 707-720.
Deming, D., Nunn, J.A., Jones, S., and Chapman, D.S., 1990, Some problems in thermal history studies, in
Nuccio, V.F., and Barker, C.E., eds., Applications of Thermal Maturity Studies to Energy
Exploration: SEPM, Rocky Mountain Section, p. 61-80.
Drury, M.J., Jessop, A.M., and Lewis, TJ., 1984, The detection of groundwater flow by precise temperature
measurments in boreholes: Geothermics, v. 13, no. 3, p. 163-174.
Forster, A., and Merriam, D.F., 1993, Geothermal field interpretation in south-central Kansas for parts ofthe
Nemaha Anticline and flanking Cherokee and Sedgwick Basins: Basin Research, v. 5, no. 4, p. 213-
234.
Forster, A., and Merriam, D.F., 1995, A bottom-hole temperature analysis in the American Midcontinent
(Kansas): implications to the applicability ofBHTs in geothermal studies: Intern. Geothermal Assoc.,
World Geothermal Congress (Florence, Italy), Proc., v. 2, p. 777-782.
Forster, A., Merriam, D.F., and Brower, J.C., 1993, Relationship of geological and geothermal field
properties: Midcontinent area, USA, an example: Math. Geology, v. 25, no. 7, p. 937-947.
Forster, A., Merriam, D.F., and Davis, J.C., 1996, Statistical analysis of some bottom-hole temperature (BHT)
correction factors for the Cherokee Basin, southeastern Kansas: Tulsa Geo!. Society Trans., AAPG
Mid-Continent Section Meeting, p. 3-9.
Forster, A., Merriam, D.F., and Davis, J.C., 1997, Spatial analysis of temperature (BHTIDST) data and
consequences for heat-flow determination in sedimentary basins: Geo!. Rundschau, v. 86, no. 2, p.
252-261.
Forster, A., Schrotter, J., Merriam, D.F., and Blackwell, D.D., 1997, Application of optical-fibre temperature
logging, an example in a sedimentary environment: Geophysics, v. 62, no. 4, p. 1107-1113.
Hermanrud, C., Cao, S., and Lerche, I., 1990, Estimates of virgin rock temperature derived from BHT
measurements: bias and error: Geophysics; v. 55, no. 7, p. 924-931.
Jessop, A.M., 1990, Comparison of industrial and high-resolution thermal data in a sedimentary basin:
Pageoph., v. 133, no. 2, p. 251-267.
Jorgensen, D.G., 1989, Paleohydrology of the Anadarko Basin, central United States, in Johnson, K.S., editor,
Anadarko Basin Symposium, 1988: Oklahoma Geo!. Survey Circ. 90, p. 176-193.
Kehle, R.O., 1972, Geothermal survey of North America: Am. Assoc. Petroleum Geologists, 1971 Annual
Progress Report, 31 p.
Kehle, R.O., 1973, Geothermal survey of North America: Am. Assoc. Petroleum Geologists, 1972 Annual
Progress Report, 28 p.
Lam, H.L., and Jones, F.W., 1984, Geothermal gradients of Alberta in western Canada: Geothermics, v. 13,
no. 3, p. 181-192.
Lucazeau, F., and Ben Dhia, H., 1989, Preliminary heat-flow density data from Tunisia and the Pelagian Sea:
Can. Jour. Earth Science, v. 26, no. 5, p. 993-1000.
Macfarlane, P.A., and Hathaway, L.R., 1987, The hydrogeology and chemical quality of groundwaters from
the Lower Paleozoic aquifers in the Tri-State region of Kansas, Missouri, and Oklahoma: Kansas
Geo!. Survey Ground-Water Ser. 9, 37 p.
Majorowicz, lA., and Jessop, A.M., 1981a, Regional heat flow patterns in the Western Canadian
Sedimentary Basin: Tectonophysics, v. 74, no. 3-4, p. 209-238.
Majorowicz, J.A., and Jessop, A.M., 1981b, Present heat flow and a preliminary paleogeothermal history of
the Central Prairies Basin, Canada: Geothermics, v. 10, no. 2, p. 81-93.
Majorowicz, J.A., Jones, F.W., Lam, H.L., and Jessop, A.M., 1984, The variability of heat flow both regional
and with depth in southern Alberta, Canada: effect of groundwater flow?: Tectonophysics, v. 106,
no. 1-2, p. 1-29.
PROBLEMS AND POTENTIAL OF INDUSTRIAL TEMPERATURE 59

Majorowicz, J.A., Jones, F.W., Lam, H.L., and Jessop, A.M., 1985, Terrestrial heat flow and geothennal
gradients in relation to hydrodynamics in the Alberta Basin, Canada: Jour. Geodynamics, v. 4, no. 1-
4, p. 265-283.
McPherson, B.J.O.L., and Chapman, D.S., 1996, Thennal analysis of the southern Powder River Basin,
Wyoming: Geophysics, v. 61, no. 6, p. 1689-1701.
Merriam, D. F., 1963, The geologic history of Kansas: Kansas Geo!. Survey Bull. 162,317 p.
Merriam, D.F., and FOrster, A., 1994, Precambrian basement control on 'plains-type folds'in the Midcontinent
region, USA (abst.), in Oncken, 0., and others, eds, 11th Intern. Conf. on Basement Tectonics '94:
Intern. Basement Tectonics Assoc. (potsdam, Gennany), p. 101-102.
Merriam, D.F., and FOrster, A., 1996, Precambrian basement control on 'plains-type folds' (compactional
features) in the Midcontinent region, USA, in Oncken, 0., and Janssen, C., eds., Basement Tectonics
11: Kluwer Acad. Pub!., Dordrecht, p. 149-166.
Musgrove, M., and Banner, J.L., 1993, Regional ground-water mixing and the origin of saline fluids:
Midcontinent, United States: Science, v. 259, no. 5103, p. 1877-1882.
Speece, M.A., Bowen, T.D., Folcik, J.L., and Pollack, H.N., 1985, Analysis of temperatures in sedimentary
basin: the Michigan Basin: Geophysics, v. 50, no. 8, p. 1318-1334.
Stavnes, S.A., 1982, A preliminary study of the surface temperature distribution in Kansas and its relationship
to the geology: unpubl. masters thesis, Univ. Kansas, 311 p.
Vik, E., and Hennanrud, C., 1993, Transient thennal effects of rapid subsidence in the Haltenbanken area, in
Dore, A.G. and others, eds., Basin Modelling: Advances and Applications: Norwegian Petroleum
Society (NPF), Spec. Publ. 3, p. 107-117.
Willet, S.D., 1990, Stochastic inversion ofthennal data in a sedimentary basin: resolving spatial variability:
Geophys. Jour. Intern., v. 103, no. 2, p. 321-339.
Willet, S.D., and Chapman, D.S., 1986, On the use ofthennal data to resolve and delineate hydrologic flow
systems in sedimentary basins - an example from the Uinta Basin, Utah, in Hitchon, B., Bachu, S.,
and Sauveplane, C.M., eds., Hydrology of Sedimentary Basins - Applications to Exploration and
Exploitation: Proc. Third Ann. Canadian/American Conference on Hydrogeology, National Water
Well Assoc., Dublin, OH, p. 159-168.
Willet, S.D., and Chapman, D.O., 1987, Analysis of temperatures and thennal processes in the Uinta Basin, in
Beaumont, C., and Tankard, A. J., eds., Sedimentary Basins and Basin-fonning Mechanisms: Can.
Soc. Petroleum Geologists Mem. 12, p. 447-461.
PRESENT HEAT FLOW ALONG A PROFILE
ACROSS THE WESTERN CANADA SEDIMENTARY
BASIN: THE EXTENT OF HYDRODYNAMIC INFLUENCE

Jacek A. Majorowicz,1 Grant Garven,2


Alan Jessop,3 and Christopher Jessop4

INorthern Geothermal Consultants


Edmonton, Alberta, Canada
2Johns Hopkins University
Baltimore, Maryland, USA
3333 Silver Ridge Crescent NW
Calgary, Alberta, Canada
4Jessop Scientific Software
Calgary, Alberta, Canada

ABSTRACT

Compilation of bottom-hole temperature (BHT) data (3,466 readings) and their


correction along the 680-km profile from the Rocky Mountain Foothills to the western edge
of the Canadian Shield has allowed detailed analysis of the 2D temperature distribution,
calculation of the thennal gradient, and estimate of heat flow. Temperature data are available
typically from only a section ofthe depth interval making 2D contouring of the temperature
field difficult. Surface temperature values were used in calculations ofthe thermal gradient.
The quality of the BHT data deteriorates towards the shallow part of the basin where larger
errors of heat-flow estimates of the profile are calculated. Estimated heat flow increases from
the southwest towards the northeast. Anomalously high values between 300 km - 500 km of
the profile are questionable.
Thennal gradient values based on BHTs and resulting estimates of heat flow have been
compared with calculated thermal gradient and heat flow from 2D steady-state models of
coupled fluid flowlheat flow in saturated porous media along the profile for 5 situations of
hydraulic conductivity in the major aquifers. The results of this comparison show that, for
uniform basement heat flow (assumed 70 mW/m2), it is difficult to explain simultaneously the
observed heat flow lows and highs by the same fluid-flow system within the range of
acceptable permeabilities.

61
62 MAJOROWICZ, GARVEN, JESSOP, AND JESSOP

Other possible explanations of high thennal gradients based on BHTs in 'the shallow
part of the basin need to be sought. It has been detennined that the effective thennal
conductivity can be a partial explanation especially if the influence of compaction of the shales
is considered. An alternative explanation is the systematic overestimate of temperature by
BHTs in the shallow northeastern part of the basin. The comparison of the shallow basin
temperature data with high-precision temperature measurements in two areas in the shallow
part of the basin shows that the industrial temperatures tend to be systematically higher. This
disagreement is interpreted to be a result of the systematic error in shallow BHT's. The
validity of thousands of industrial BHT's from shallow depths (from 500 to 900 m)
documenting the regional thennal high is in question. This observation calls for reevaluation
of thennal gradient maps based on BHTs only; especially in the shallow part of the basin
towards the shield where anomalously high thennal gradients (>40 mKlm) have been reported
previously.
Precise temperature measurements in three sites in the shallow part of the basin and the
geothennal gradient map based on shut-in well temperatures in oil pools show that the change
from low gradients in the Foothills «25mK1m) to highs in the foreland basin (>35mK1m)
takes place over the horizontal distance of less than 300 Ian and the latter change can be
explained easier by the hydrodynamic heat flow coupled system based on a 2D model.

INTRODUCTION

In the Western Canadian Sedimentary Basin (WCSB) heat flow is estimated, from
industrial bottom-hole temperature data (BHT), and estimates of effective thennal conductivity
which range from 40 mW/m2 to 120 mW/m2 for a distance of 400 Ian between the topographic
high and low (Majorowicz and Jessop, 1981; Majorowicz and others, 1995). South of the
border, heat flow in the Denver Basin of the Great Plains in the USA ranges from 40 mW/m2
to 80 mW/m2 for a distance of some 100 Ian (Gosnold, 1985). In a recent work, Gosnold
(1996) detennined that heat flow ranged from 20 mW/m2 in the margins of the Black Hills to
140 mW/m2 along the eastern flank of the basin in South Dakota and northern Nebraska.
All of the described heat-flow variations have been interpreted to a large extent to be
the result of basin-wide gravity-driven regional fluid flow "sweeping" terrestrial heat from the
high-topography recharge zone towards a low-topography discharge zone. The wedge shape
of foreland basins seems to amplify the effect (Jessop, 1989) by focusing the up flow. A
general model of steady state, coupled fluid and heat flow has been proposed for foreland
sedimentary basins and the accumulation of both hydrocarbon fields and lead-zinc ore deposits
in the shallow part of the Western Canada basin (Garven,1989).
Although regional variations of heat flow have been documented for the Western
Canada foreland basin, proof of advective disturbance has not been documented and is
controversial (Majorowicz, 1989; Bachu, 1988,1993). Bachu (1988) used dimensional
analysis to argue that the aspect ratio of the proposed flow system does not allow the amount
of heat transfer proposed by Majorowicz and Jessop (1981). Theoretical models for
topography-driven flow usually have been based on numerical studies of generic
hydrogeologic systems (Garven and Freeze, 1984; Deming and Nunn, 1991). Only a few
attempts have been made to calibrate basin-scale flow models with observed heat-flow data.
Probably the most successful field study is the analysis by Deming and others (1992) for the
North Slope of Alaska. The model by Deming and Nunn (1991) of brine migration by
topographically driven recharge for the wedge shape basin 400 Ian long and 6 Ian deep
HEAT FLOW IN THE WESTERN CANADA SEDIMENTARY BASIN 63

(maximum) showed that the high fluid velocity required to carry heat to the basin margins
exhausts the supply of solute and reduces heat flow to low values approaching 0 mW/m2 •
In some basins, knowledge of topography, permeability, thermal conductivity, and
basement heat flow permit the regional-scale permeabilities to be constrained by the heat-flow
observations in the sedimentary succession. In this paper we compare observed thermal data
on a profile across central Alberta (Fig. 1) with regional trends predicted by numerical
simulations of the fluid-flow system.

'" PHANEROZOIC ISOPACH

c:=:J ,q.~-~---."
e:::g _ . , _.... .. ::::~'!.--- .
~.--.-.-

". \
\
\ ...
\
\
\ .,.

\
\ G·

". \

...
Figure 1. Location of profile across Western Canada Sedimentary Basin, Profile D-D' of Wright (1984).

TEMPERATURE DATA

Industrial Temperature Data

Industrial bottom-hole temperature data (BHT) have been collected from 1795 wells
along the profile shown in Figure 1. The profile coincides with the Section D-DI of the
Canadian Society of Petroleum Geologists Map of the WCSB (Wright, 1984). BHT data from
a wide band of townships along the profile have been projected to the 2D plane along the
profile. The original data with the locations are to be included in the Canadian Geothermal
Database (Jessop and others, in preparation).
The raw data require a correction to remove the thermal disturbance of drilling, which
may be calculated if the time of end of drilling, time of end of circulation, and time of
measurement oftemperature are recorded for a series of measurements. The time interval from
64 MAJOROWICZ, GARVEN, JESSOP, AND JESSOP

the end of drilling to the end of circulation is known as the time of circulation (t 1), and the time
from the end of circulation to the time that the logging tool leaves the bottom of the well is
known as the time since circulation (tz).
Unfortunately, only 10 to 20% oflogs have the necessary recorded three time values.
Some records give only the end of circulation and time on bottom, and some give none of these
data. Where the time data are inadequate some assumptions must be made from which to
derive a correction. Where a series oftemperature with full time data are available a corrected
temperature, approximately representing the undisturbed rock temperature, may be derived by
the Homer method (Lachenbruch and Brewer, 1959).

Temperature V is plotted against a function t, given by

(1)

The point where t = 0 implies that the time since circulation is long compared with the time
of circulation. Thus, the intercept of the straightline through these data on the V -axis may be
taken as the best estimate of the equilibrium rock temperature.
Any series of measurements from logs that start within a depth interval of 10m is
acceptable. In a normal geothermal gradient of 20 to 30 mKlmz this implies a temperature
range of 0.2 to 0.3 K, which is within the limit of probable error for the data. The effective
depth is taken as the mean of the individual depths.
Wherever possible corrected temperatures have been calculated by this process. In
addition to the direct corrections, this provides a file of values of the gradient dV/dT of the
straightline. This gradient depends on various factors of which the well diameter and depth
probably are the most significant. This gradient has been analyzed as a function of depth and
well diameter and has been used where there is a single temperature data or where time data
are inadequate for a direct correction.
Where the time data are partially or completely missing, or where careless recording
is suspected or obvious errors are present, assumptions have been made in order to obtain the
corrected temperatures. The procedure used with a series of temperatures has been as follows.
(1) Where there is a sequence of temperature at different times:
a - Ifthe temperatures are different, a correction has been calculated and the value of
the gradient dV/dt has been recorded:
b - If two or more ofthe temperatures are the same, the earlier one has been accepted
as more likely to be correct than the later one, the later one has been discarded, and
the procedure of I-a has been followed;
c - If all the temperatures are the same, it has been assumed that all except the earliest
are false, the loggers having failed to make and record measurements after the first,
and the earliest has been used as a single temperature, as in 2-a
(2) Where there is a single temperature:
a - If all times are given, the gradient-function has been used with the single point
given by the data to calculate a correction,
b - If one or more time is missing, it has been assumed that the missing time periods
are 2.0 hours, and the correction has been calculated as in 2 - a.
(3) Where there are multiple temperatures but missing times:
a - Where no times are known it has been assumed that the lowest temperature is the
earliest and that any missing time is 2.0 hours, and the correction has been
calculated as in 2 - a;
HEAT FLOW IN THE WESTERN CANADA SEDIMENTARY BASIN 65

b - If all temperatures are the same and no times are known, the times are assumed to
be 2 hours and the correction has been calculated as in 2 - a.

A total of 1795 wells with a direct correction, based on two or more different
temperatures with full time data were used. From these data the gradient of temperature with
time function, dV/dt, as a quadratic function of depth in the well, z, and diameter of the well,
d, is determined to be

dV/dt = -14.3-6.156*10. 3 z+104 d+1.815*1Q-6 z2-2.216*1Q-2 d z-396 d2 (2)

where z is depth and d is diameter, both in meters.

The analysis of the data grouped for the large sections ofthe profile and plotted with
depth (Fig. 2) shows that the data are not evenly distributed over the full depth interval of the
sedimentary strata and that the representation of the thermal field of the 2D section by the
BHTs deteriorates towards the shallow basin in the northeast where most ofthe data are at 400
m ±100 m. It also has been observed that the shallow depth data can differ by 25 0 K for the
same depth. Examples of the plots of temperature with depth for 10 Ian sections of the profile
are shown in Figure 3.

Precise Temperature Logs

Comparison has been made between precise temperature logs in stabilized wells
(Majorowicz, 1996; Majorowicz and Safanda, 1998) and BHT in two regions in northeastern
WCSB in Alberta known for their high thermal gradients; Cold Lake (54.5 0 N, 110.5 0 W) and
Fort McMurray (57.3 0 N, I11.r W). The plots of precise temperatures are shown in Figure
4 with limits of the gradients implied by the BHT data. The precise logs reach a depth of only
350 m, but these examples suggest that at 500 m thermal gradients and temperatures are lower
than those indicated by the BHT. These data seriously question high gradients in the shallow
northeastern part of the basin based on the BHTs. It is possible that BHTs taken at shallow
depths in winter conditions, when the mud was warm and did not represent virgin rock
temperatures. Alternatively, in summer it is possible for atmospheric temperature of20° C
or more to exceed BHT, making the maximum-reading glass thermometers useless. Both
effects would produce systematically high-temperature readings.
The disagreement between shallow depth BHTs and precise temperature measurements,
the lack of complete time data, and obvious errors in recording data all call for caution in the
analysis of the shallow basin thermal gradients based on industrial data.

GEOTHERMAL GRADIENT CALCULATED FROM OBSERVED DATA

The geothermal gradient has been calculated from the reduced BHT data using two
approaches. Most of the wells lack temperature data at shallow depths. Surface temperature
constraint has been used as described by Majorowicz (1996). The surface temperature is
disturbed by the ground warming over the last century, which has been observed in the Prairies
provinces. Present surface temperatures are on average 2.1 0 K higher than before the climatic
warming and land development ofthis century so a surface temperature of 4 0 C ±1 0 K has
66 MAJOROWICZ, GARVEN, JESSOP, AND JESSOP

TEHPERRTURE (CI TEHPERATURE (el


0- 25 0 25 50 75.100125 ISO I7S 0- 25 0 25 SO 75 100 125 150
0 0
l/1
l/1
0 0
0
~ 0
l/1 ~

l/1
.... "'"
0
~N ::E~
:5
~~N 0
:r:'
:r:o t-'"
t
tU
cri+ "-
tUl/1
O.
0l/1 tv
en
0 0

.; ..;
l/1
.; l/1
..;
.. . A
0
ul 0 B
l/1 .;
ul
TEMPERRTURE ICI TEMPERATURE ICl
0- 25 0 25 50 75 100 125 0- 25 0 25 50 75 100
ci ci
Ul l/1
ci +
ci
0
l' f

+.*
0

2'1.' CMI
~ +
~Ul :5
~....: l/1
:r:...; +
:r: t-
1-0 "-
tU
IL
tUN 00
0
ru
Ul

N Ii
0
i l/1
ru
:E
M "
0
e
0

Ill-
! 0
en
M

Figure 2 _ Plots ofBHT data against depth for sections of profIle D-D': A - 0-50 km; B - 50-100 km; C - 100-150
km; D - 150-200 km.

been assumed. This is the temperature with which the deep subsurface temperatures are in
approximate equilibrium.
Geothermal gradients calculated for individual well data are shown in Figure 5. The
northeastward increase along the profile culminates with an anomalous high at 400 - 500 km.
The northeastward increase of values has been approximated by a 2nd_degree polynomial. At
the extreme northeastern end of the profile where data are scarce the temperature gradient
declines.
The average geothermal gradient is calculated from the plots oftemperature with depth
for 10 km sections ofthe profile (example 5 km ±5 km, 15 km ±5 km, 25 km ±5 km, etc.).
It is apparent from the analysis of the plots that the validity of thermal gradients is the highest
HEAT FLOW IN THE WESTERN CANADA SEDIMENTARY BASIN 67

TEHPERATURE 1[1 TEHPERATURE 1[1


0- 25 0 25 50 75 100 0- 25 0 25 50 75 100
ci ci

III
III
ci ci
0 i: +
~
0 +
i: r·
!--
~ Q..
III
w 39.1 CJI(I1
r· a
!-- +
-
"- III
W
0 0 +
N
0 F
III
N
N

0
m
E
TEHPERATURE 1[1 TEHPERATURE 1[1
0 25 50 75 0- 25 0 25 50 75
0- 25

ci ci

++
III III

I'ci ++ I'ci
~ + +
~
++
r r
!- + +
!-
"- "-
wo
wo
a
- a
- +
GRADIENT
+
53.7
GAAlJIENT ~ ClKH ~

III G III H

Figure 2. Plots ofBHT data against depth for sections of profile D-D': E - 200-250km; F - 250-300 km; G-
300-400 km; H - 400-500 km.

in the first 200 km of the profile whereas the data from the shallow parts depend on a cluster
of points over small depth intervals. The plot of the average thermal gradient in 10 km
sections is shown in Figure 5B. A 6th-degree polynomial fit to the average also is shown. The
northeastward increase of the thermal gradient is apparent to about the 500 km point then
followed by a decline.
However, the thermal gradient observed is based on data of limited depth. Thermal
gradient increases smoothly from 25 mKlm2 to 35 mKlm2 in the first 200 km from the edge
of the Rocky Mountains, where the best and most numerous data are available. Both plots
show this feature.
The geothermal gradient map based on the temperature measurements in shut-in wells
in oil and gas areas (Fig.6) does not extend to the northeastern shallow part offue basin, but
it shows an increase of geothermal gradient along the profile perpendicular to the strike of the
basin similar to the gradients based on deep basin BHTs only.
68 MAJOROWICZ, GARVEN, JESSOP, AND JESSOP

TEMPERRTURE ICI TEMPERRTURE ICI


0-25 0 25 50 75 100 125 150 175 o 25 50 75 100
o o
III

o
o III

o
III +
o
~{\J
o
I'
~lIl t- ......
D....
{\J l1.J GRRDIENT l!7. I ClKH
IO o
t-,
~(Tl III
OlIl
(Tl

o
.. 380.0 KM 0 seo.o KM
:r
III ~~----------------------------~B
:r
o
III
0.0 KM 10.0 KM
III
A
0

~------------------~~
III

Figure 3. Example of plot of temperature with depth for 10 !an sections of profile DoD': A deep part of basin;
0

B shallow part of basin.


0

EFFECTIVE THERMAL CONDUCTIVITY

The effective thermal conductivity has been calculated for the same depth intervals as
the average temperature gradients, The method of calculation has been described as model (1)
in Majorowicz and Jessop (1993). The variations of the effective thermal conductivity are
based on the known lithology and assumed average rock conductivities (Beach, Jones, and
Majorowicz, 1987). The resulting conductivity values were assigned to major
litho-stratigraphic units.
The variations of the effective conductivity shown on the profile in Figure 7 do not
reflect variations of the thermal conductivity of the entire sedimentary section. They show
thermal conductivity for the part of the sedimentary strata for which thermal gradient could be
determined. The values decrease through the first 450 km of the profile partly accounting for
the increase in thermal gradient (Figs. 5 and 6). The effective thermal conductivity ranges from
2 W/mK to 2.6 W/mK.
The estimates of thermal conductivity depend on the range of assumed thermal
conductivity values for different rock types because there are no measurements from the core
samples available for the wells along the profile. Additional measurements of thermal
conductivity on cores are needed in the future.
HEAT FLOW IN THE WESTERN CANADA SEDIMENTARY BASIN 69

30 40
fOB.T YcMUlLKAY AREA
COLDLAXB

25
30
........ 20
~ Q
cL 15 cL 20
:2 :2
W w
I- 10 I-
10
5

o
o 100 200 300 400 500 100 200 300 400 500 600
B
OEPTH(m) A OEPTH(m)
Figure 4. Comparison of precise temperature logs: A - in area of Cold Lake and B - Fort McMurray. Thennal
gradients are 20.2 and 28.9 mK/m, respectively, shown with limits oftemperatures from BHTs (broken lines).

ESTIMATED PRESENT HEAT FLOW

Heat-flow distribution is calculated from the average thermal gradient and effective
thermal conductivity and is shown in Figure 8. Calculated heat flow increases from 60 mW/m 2
to maximum of 140 mW/m2 with the highest values between 400 km and 500 km of the
profile. Calculated heat flow ranges between 60 mW/m2 and 75 mW/m2 in the first 200 km
of the profile where the best temperature data are available.

100 100
AVI!:RAOE &niT nON PL01'S' OP T VI DEPTH

.. "
....
80 80
E E
~ ~ 60
.s 60 E
I- f="
o ~
« 40
IT: 40 IT:
CJ CJ
20
20
FOR EVERY IIIKm+/-5Km AL0N3ll-iE PROfiLE

o 200 400 600 A 200 400 600


018T ANCE(Km) 018T ANCE(Km) B
Figure 5. Geothenna1 gradient calculated from BHT data and surface temperature plotted against position on
profile D-D': A - individual values, with sections of profile with large spread marked by shading; and B - average
values calculated from plots of temperature with depth for 10 km sections along profile. Apparent anomalous
section of profile is marked by shading.
70 MAJOROWICZ, GARVEN, JESSOP, AND JESSOP

The calculated heat-flow values in the eastern part of the basin are almost twice as high
as the basement heat-flow values (70 ± 20 mW/m2) interpreted from heat-generation data
(Jessop, 1992). Anomalously high heat-flow values are mostly in the shallow northeastern part
of the basin where the noisy industrial temperature data comes from a limited depth interval
and shallow depth. In contrast, relatively low heat flow is suggested in the far northeastern

I
I SASK.
I
~

Geotr\ermOI
Groo,er\1
h "-
i '-'-'"
,

_>~Om.r<.m·1
D 3~'<O
D:w-~ I .. ,
DZ5-:W .
0<25
' ,I
- ._ ., 4'1 I """""
1 .. ··· · 6\
2····· ·
,
o
. 42 ·u .'-;:

Figure 6, Geothennal gradient map based on thermal measurements in oil pools in shut-in wells, Position of
profile is shown. Heat flow values (squares) are in mW/m2 (Majorowicz and Jessop, 1981).

part of the profile, both by the low thermal gradients (20-29 mKlm) from precise shallow
temperature measurements (Fig. 4) and by its proximity to the exposed Canadian Shield, which
is known to have generally low heat flow (Drury, 1985).
HEAT FLOW IN THE WESTERN CANADA SEDIMENTARY BASIN 71

3
ES'IDoIATB OF THE EFFECTIVE THERMAL CONDucnvrn'

q2.5
E
~
LL
LL
W 2
~

1.5
o 200 400 600
018TANCE (Km)
Figure 7. Estimates of effective thennal conductivity for depth intervals for which average thennal gradients
were available, at 10 km intervals on proflle 0-0'.

150

....
(\j100

E
~
E
0: 50

200 400 600


018T ANCE(Km)
Figure 8. Estimated present heat-flow profile (0-0') from BHT data and effective thermal conductivity, at 10
km intervals. Heat flow from crystalline basement = 70 mW/m2 (Jessop, 1993) is shown.

COMPARISON OF MODELED THERMAL GRADIENT AND HEAT FLOW BASED


ON THE STEADY-STATE MODELS OF COUPLED FLUID FLOW AND HEAT
FLOW WITH OBSERVATIONS

Quantifying conductive and convective heat flow in sedimentary basins is possible


provided material properties such as permeability, porosity, and thermal conductivity are well
72 MAJOROWICZ, GARVEN, JESSOP, AND JESSOP

enough known and if hydraulic and thermal boundary conditions can be assigned.
Hydrogeologic models for regional flow in foreland basins are based mainly on numerical
models for fluid and heat flow, either in transient or steady flow modes (Garven, 1995). One
hydro stratigraphic model of the profile is presented here, based on the geological section
Section D-DI of the CSPG Map of the Western Canada Sedimentary Basin (Wright, 1984).
The hydro stratigraphic model by Garven coincides with this profile. The model and the finite
element mesh are shown in Figure 9. The mathematical model and assumptions for fluid
migration and heat flow are given by Garven (1986,1989) and will not be repeated. This is a
steady-state model of coupled fluid heat flow in saturated porous media, assuming present
geometry. Flow is assumed to be confined to the vertical plane and laterally bounded by the
thrust belt in the west and Precambrian shield in the east. The watertable follows topography.
The assigned thermal conductivities are identical to the ones used in this work for heat-flow
calculations. An average basement heat flow of70 mW/m3, typical for the Alberta Basin
(Beach, Jones, and Majorowicz, 1987; Jessop,1992) has been assumed for the basement. The
temperature at the watertable is assumed to be 50 C. The geologic units are assumed to be one
hundred times more permeable parallel to bedding than perpendicular to bedding. Hydraulic
conductivity varies from one hydrostratigraphic unit to another but it is assumed to be uniform
within a given unit.
Two-dimensional simulations of temperature on the profile allowed calculation of
thermal gradients for the same depth intervals for which average thermal gradients were
calculated from the data. Heat-flow profiles have been calculated for five cases, each having
a combination of assumed hydraulic conductivity for the hydro stratigraphic units.

HYDROSTRATICRAPHY

A
.00

FINIte:: E:Le:ME:NT "-4ESH

B
150 ZOO '2'5.0 .soo 3~ 400 .$0 !tOO ~~ eoo
KII..O..-t'TCfIt$

Figure 9. Hydrological model of flow in Western Canada Sedimentary Basin, profile D-D'. Example shown is
case 2. A - hydrostratigraphy; B -fmite element mesh.
HEAT FLOW IN THE WESTERN CANADA SEDIMENTARY BASIN 73

HYDRAULIC HEADS

100 fSO '00 400 4~O 500 .550 fS-OO &~O 700 000
KILOMETERS

VELOCITY VECTORS

KILOMETERS

Figure 9. Hydrological model of flow in Western Canada Sedimentary Basin, profile D-D'. Example shown is
case 2. C - hydraulic head (contours every 50 m); D - groundwater velocity (maximum velocity vectors - 1.1
m/yr).

STREAM FUNCTIONS

TEMPERATURES

KILOMETERS

Figure 9. Hydrological model of flow in Western Canada Sedimentary Basin, profile D-D'. Example shown is
case 2. E - stream functions (1\1jI=50rn3/yr/m); F - temperature distribution (every 10 C).
74 MAJOROWICZ, GARVEN, JESSOP, AND JESSOP

Case 1: low hydraulic conductivity estimates from Bachu (1988) are considered. The Tertiary
and Cretaceous shale (unit 6) has a hydraulic conductivity of 10-2 m/yr. Devonian carbonate
(unit 4) has conductivity 10-6 m/yr.

Case 2: hydraulic conductivity of Devonian aquifer (unit 4) is assumed to be 200 m/yr


(permeability of 660 millidarcys) and of the basal Cambrian sand to be 50 m/yr.

Case 3: hydraulic conductivity ofthe Devonian aquifer (unit 4) is set to 2000 m/yr which is
considered the upper limit.

Case 4: the basal Cambrian sandstone aquifer is assumed to be the most permeable aquifer in
the section (unit 1), with a conductivity of 500 m/yr. Devonian carbonate (unit 4) has 100m/yr.

Case 5: the basal Cambrian aquifer (unit 1) has a conductivity of 500 m/yr, the Cambrian
shale/silt/carbonate (unit 2) has 500 m/yr and Devonian Carbonate (unit 4) has 10 m/yr.

Thermal gradients calculated from these five systems are shown in Figure 10. All the
simulations illustrate regional effects of fluid migration. The best control is the permeability
assigned to the Upper Devonian aquifer. However topography plays a major role also in that
regional groundwater discharge today is confined to the zone from 10 Ion to 350 Ion along the
profile (see case 2 for example, Fig. 9B). As a result of the topography-controlled fluid flow,
thermal gradient is mostly accentuated in the same areas, regardless of the magnitude of
assigned permeability values.
Calculations of heat flow have been based on the thermal gradient profiles and effective
thermal conductivity estimates from the model (Fig.9A), and are shown in Figure 11.
None of the simulated heat-flow and thermal gradient variations along the profile
match the observed thermal gradient and heat flow for the whole length of the profile. If
observed measurements are accurate, the poor match may be the result of the assumed
permeability field, which is unconstrained by exact permeability and porosity information.
The simulations seem to replicate the general trend of observed heat flow only in the southwest
(Fig.8).
In the situations of high-estimated permeabilities, high heat flow is predicted for the
central part of the profile, but the contrast with heat flow in the southwest is high. For low
permeabilities some reduction of the heat flow in high topography deep southwestern part of
the basin takes place, which agrees with observations, but the predicted high heat flow for the
shallow, low-topography part of the basin disappears. Although only five examples are
presented, their diversity suggests that no single deep flow system can uniquely account for
both the lows and highs in the observed heat flow along the profile.

DISCUSSION

The comparison of modeled heat flow and thermal gradient with observations along
the profile does not support the hypothesis that heat-flow variations across the strike of the
basin are controlled entirely by heat transport by deep, large, regional-scale fluid flow.
However, it is possible to explain reduction of the observed thermal gradients in the high
topography deep basin (Fig.5) by the set of low hydraulic conductivities given in Bachu
(1988). Extremely high hydraulic conductivities on the order of 1000 m/yr are needed to
HEAT FLOW IN THE WESTERN CANADA SEDIMENTARY BASIN 75

100 100 100


..a'CMCIIU.'nII _ _ m .-rCII..C.'III.AtID_ .....ca
pofl'CALCUlA1IDftOM ..... M

eo 80 80

E E E
~eo ~eo ~ 60
E E E
i=" i=" i="
Q 40 Q 40 Q 40
a: a: a:
<.'J <.'J <.'J
20 20 20

0 0 0
0 200 400 eoo A 0 200 400 800 B 0 200 400 eoo C
OISTANCE(Km) OISTANCE(Km) OISTANCE(Km)

100 , . - - - - - - - - - - , 100 . - - - - - - - - - - ,

eo 80
E E
.s~eo
I-
~eo
E
i="
~ 40 Q40
<.'J
a:
<.'J
20 20

°0~-~200~--4~00~--6~00-D~ 00L--~20~0--~40~0--~eo~0~E
OIST ANCE(Km) OISTANCE(Km)

Figure lO. Modeled thennal gradient along profile D-D' for cases 1-5, A - E respectively.

explain observed heat flow high in the shallow northeastern part of the basin. However, such
values would cause a reduction of heat flow in the deep, high-topography part of the basin to
less than 10 mW/m2, which is not supported by any observations. The reasonable assumption
of a moderate heat-flow value of 70 mW/m2 from the Precambrian basement is enough to
provide a sufficiently large supply of heat in the deep part ofthe basin.
If a large regional flow system cannot support the observations of heat flow for the
whole length of the section other possible explanations of the observed anomalous heat flow
need to be sought, as follows:
(1) The assumption ofunifonn basement heat flow, typical for the cratonic basins, of
70 mW/m2 is not valid. This cannot be verified without coring and temperature logging of
deep wells drilled for scientific purposes. Theoretical estimate of the basement heat flow from
the data on heat generation of the basement cores, age, geological division of the basement
provinces and assumed "reduced" heat flow, gives the basement heat flow to be mainly in the
60-80 mW/m2 range (Jessop, 1992). Bachu (1993) based his estimate of the basement heat
flow (40-80 mW/m2) on the BHTs from the base of the sedimentary succession, surface
temperatures, and estimates of thermal conductivity for the sedimentary fill. He assumes that
the transfer of heat in the basin is entirely by conduction with no influence of fluid flow and
corrects for the heat generation of the sediments.
(2) Effective thermal conductivity differs from the estimates. The thennal conductivity
distribution required to maintain unifonn heat flow (70 mW/m2 ±1O) across the basin is
reasonably close to values measured for sedimentary rocks. However, in order to explain
76 MAJOROWICZ, GARVEN, JESSOP, AND JESSOP

150,------ 150 , - - - - - - - - - - - , 150,---------,

r~1 i
fi100 fi100

0'50 - 0'50
I
0'50

o ~I_~--=--~-' O~--~--~--~~
B
o 200 400 600 A o 200 400 600 200 400 600 C
OIST ANCE(Km) OISTANCE(Km) OISTANCE(Km)

150 150
~"'TI'UI'I'

""~

fi100 fi100
'E 'E
~ ~
E E
0'50 0'50

0 0
0 200 400 600 D 0 200 400 600
OISTANCE(Km) OISTANCE(Km) E
Figure 11. Modeled heat flow along profile D-D' for cases 1-5, A - E respectively.

thennal gradient variations without heat-flow change along the entire profile, low
conductivities of 1.0 to 1.3 WImK would be needed for the observed high thennal gradient
zone between 400-500 km of the profile. These figures apply to the average thennal
conductivity throughout the vertical sedimentary sequence. Although some individual shale
units may have these conductivities, or variations in compaction of clastics, particularly shales,
may influence conductivity, it is unlikely that the average conductivity of the whole
sedimentary sequence can differ by so much. Such a variation is not supported by laboratory
measurements of thennal conductivity of preserved shale core samples (Jessop and Issler,
1998, unpublished results). Measurements on fragments are not sufficiently numerous or
precise, particularly for anisotropic rocks.
(3) Quality ofBHTs is low especially in the shallow part of the basin where anomalous
high temperatures have been recorded by industry in thousands of wells at shallow depths of
about 500 m. Alternatively, it may be that the correction applied to the deeper data is not valid
for the shallow data. The apparent high thennal gradients (>40 mKlm) are a result of these
high temperatures. As a result BHT-inferred gradients are too high (USGS/AAPG,1976;
Jones, Majorowicz, and Lam, 1985; Majorowicz and others, 1986) and the resulting heat-flow
values are higher than the 70 mW/m2 or less that has been assumed in the past (Jones,
Majorowicz, and Lam, 1985; Majorowicz and others, 1986). This thennal gradient
discrepancy has been recognized only after comparison of the BHT data with precision
borehole temperature logs made for climate-change purposes (Majorowicz, 1996; Majorowicz
and Safanda, 1998). The problem probably is limited to the shallow «1 km) data.
Comparison of industrial and the few precise temperature data by Jessop (1990) show no
HEAT FLOW IN THE WESTERN CANADA SEDIMENTARY BASIN 77

systematic error. Further, precise logs in wells of intermediate depth would be a valuable
check on BHT data there. The revised geothermal gradient map from BHT data in the Western
Canada Sedimentary Basin is based on the intermediate and deep temperatures only (Fig. 12).
The shallow data in the northeastern part of the basin used in previous works (Majorowicz and
others, 1986; Jones, Majorowicz, and Lam, 1985) has been removed. This revised map of
thermal gradients for the sedimentary strata above the Paleozoic erosional surface, as well as
the map of average thermal gradients for the entire sedimentary succession based on
temperatures measured in "shut in" wells, show a smaller range oftemperature gradient across
the basin. The range of thermal gradients is 20 mKlm-40mK/m and the change from low in the
foothills to the high in the central part of the foreland basin is for the distance less than 200
km.
(4) The inferred lateral groundwater flow needed to generate the presumed heat-flow
change from southwest towards northeast is operating on a smaller lateral scale «200 km).
The hypothesis of basin-wide thermal transport by the flow of water from the Rocky
Mountains to the lowlands has generated much discussion (Bachu, 1988, 1993; Gosnold, 1985;
Majorowicz, 1989). Bachu (1988,1993) has contended that the aspect ratio ofthe flow system
is such that no realistic set of parameters will generate the apparent high heat-flow values
proposed in the past for the eastern part of the section. These concerns were based on the

- - - -,-,."
110' ".,
- 1- _ - ~ - -
__'02'
\-.0•

SASKATCHEWAN

, .
rOO

'.
" .....
f,
~~.

Figure 12. Revised geothermal gradient map based on BHTs above Paleozoic erosional surface (in mKlm).
Note that umeliable BHTs in shallow NE part of basin have been rejected.
78 MAJOROWICZ, GARVEN, JESSOP, AND JESSOP

assumption that the heat-flow change from low in the Rocky Mountains to high in the shallow
basin is along the distance of some 500 km. This is now in question (see point 3). The smaller
horizontal scale of the lateral heat-flow change (less than 300 km distance) allows the low heat
flow in the Foothills and high in the foreland basin to be explained at least partly by the
hydrodynamic effect upon heat transfer.
The hypothesis of basin-wide distortion of heat transport by water flow is not supported
by these observations and simulations, at least for present-day hydrogeologic conditions. It
seems that the observations can be duplicated by numerical simulations only in the first 200
km northeast of the Rocky Mountains. However, the quality of the data is questionable and
further reliable geothermal surveying may be required before a definite judgement of the
hypothesis can be rendered.

ACKNOWLEDGMENTS

The authors gratefully acknowledge the careful work of Marcela Deuma in transcribing
the temperature data from log headings to electronic files. Part of this work was supported by
a research contract of the Geological Survey of Canada (No. 23294-5-0669/01-XSH) to
Northern Geothermal (J.A.M., A.J. & CH.J.). Help of H. Abercombie and G. Mossop is much
appreciated.

REFERENCES

Bachu, S., 1988, Analysis of heat transfer process and geothermal pattern in the Alberta Basin, Canada: Jour.
Geophys. Res., 93, no. B7, p. 7767-7781.
Bachu, S.,1993, Basement heat flow in the Western Canada Sedimentary Basin: Tectonophysics, v. 222, no. 1,
p. 119-133.
Beach, R,D.W., Jones, F.W., and Majorowicz, lA., 1987, Heat flow and heat generation estimates for the
Churchil basement of Western Canadian Basin, Alberta, Canada: Geothermics, v. 16, no. 1, p. 1-16.
Deming, D., and Nunn, lA., 1991, Numerical simulations of brine migration by topographically driven recharge:
Jour. Geophys. Res., v. 96, no. 2,2485-2499.
Deming, D., Sass, J.H., Lachenbruch, A.H., and De Rito, R.F., 1992, Heat flow and subsurface temperature as
evidence for basin-scale ground-water flow, North Slope of Alaska: Geol. Soc. America Bull., v. 104,
no. 5, p. 528-542.
Drury, M.J., 1985. Heat flow and heat generation in the Churchill Province of the Canadian Shield, and their
palaeotectonic significance: Tectonophysics, v. 115, no. 1-2, p. 25-44.
Garven, G., 1985, The role of regional fluid flow in the genesis of the Pine Point deposit: Economic Geology,
v. 80, no. 2, p. 307-24.
Garven, G., 1986, Quantitative models for strataboumd ore genesis in sedimentary basins, in Hitchon, B., Bachu,
S., and Sauveplane, C.M., eds., Hydrogeology of Sedimentary Basins: Proc. 3rd Canadian/American
Conference on Hydrogeology, Banff, Canada, p.69-74.
Garven, G., 1989, A hydrogeological model for the formation of the giant oil sands deposits of the Western
Canada Sedimentary Basin: Am. Jour. Science, v. 289, no. 2, p. 105-166.
Graven, G., and Freeze, R.A., 1984, Theoretical analysis of the role of groundwater flow in the genesis of
stratabound ore deposits; 1. Mathematical and numerical model: Am. Jour. Science, v. 284, no. 10, p.
1085-1124.
Gosnold, W.D., 1985, Heat flow and groundwater flow in the Great Plains of the United States: Jour.
Geodynamics, v. 4, no. 1-4, p. 247-265.
Gosnold, W.D., 1996, Basin scale ground water flow and advective heat flow: an example from the N. Great
Plains (abst.): IHFC, IASPEI Heat Flow Workshop Trst, Chech Rep. June 9-15, 1996.
HEAT FLOW IN THE WESTERN CANADA SEDIMENTARY BASIN 79
Jessop, A.M., 1989, Hydrogeological distortion of heat flow in sedimentary basins: Tectonophysics, v. 164, no.
2-4, p. 211-218.
Jessop, A.M., 1990, Comparison of industrial and high-resolution thermal data in a sedimentary basin: Pure and
Applied Geophys., v. 133, no. 2, p. 251-267.
Jessop, A.M., 1992, Thennal input from the basement of the Western Canada Sedimentary Basin: Bull. Can.
Petroleum Geologists, v. 40, no. 3, p. 198-206.
Jones, F.W., Majorowicz, J.A., and Lam, H.L, 1985. The variations of heat flow density with depth in the Prairies
Basin of Western Canada, Tectonophysics, v. 121, no. 1, p. 35-44.
Lachenbruch, A.H., and Brewer, M.C., 1959, Dissipation of the temperature effect of drilling a well in Arctic
Alaska: U.S. Geol. Survey Bull. 1083C, p. 73-109.
Majorowicz, J.A., 1989, The controversy over the significance of the hydrodynamic effect on heat flow in the
Prairies Basin: Am. Geophys. Union. Geophys. Mon. 47, p. 101-105.
Majorowicz, J.A., 1996, Accelerating ground warming in the Canadian Prairie Province: is it a result of global
warming?: Pure and Applied Geophys., v. 147, no. 1, p. 1-24.
Majorowicz, J.A., and Jessop, A.M., 1981, Regional heat flow patterns in the Western Canadian Sedimentary
Basin: Tectonophysics, v. 74, no. 3-4, p. 209-238.
Majorowicz, J.A., and Jessop, A.M., 1993, Relation between basement heat flow and thermal state of sedimentary
succession of the Alberta Plains: Bull. Can. Petroleum Geology, v. 41, no. 3, p. 358-368.
Majorowicz, J.A., and Safanda, J., 1998, Ground surface temperature history from inversions of underground
temperatures - Western Canadian Sedimentary Basin case study: Tectonophysics, v. 291, no. 1-4, p.
287-298.
Majorowicz, J.A., Jones, F.W., Lam, H.L. and Jessop, A.M., 1995, Terrestrial heat flow and geothermal gradients
in relation to hydrodynamics in the Alberta Basin, Canada: Jour. Geodynamics, v. 4, no. 1-4, p.
265-283.
Majorowicz, J.A., Jones, F.W., Ertman, M.E., Linville, A., and Osadetz, K., 1986, Heat flow in the Edmonton-
Cold Lake region of the Western Canada Sedimentary Basin and the influence of fluid flow: Proc.3 m
Canadian/American Conference of Hydrology, Banff, Canada, p. 151-158.
Osadetz, K.,G., Jones, F.W., Majorowicz, J.A., Pearson, D.E., and Stasiuk, L.D., 1990, Thennal history of the
Cordilleran Foreland Basin in Western Canada: a review, in Macqueen, R.W., and Lechie, D.A., eds.,
Foreland Basins and Fold Belts: Am. Assoc. Petroleum Geologists Mem. 55, p. 259-278.
USGS/AAPG, 1976, Geothermal gradient map of North America, scale 1:500,000,2 sheets, Washington, D.C.
Wright, G.N., ed., 1984, The Western Canada Sedimentary Basin, a series of geological sections illustrating basin
stratigraphy and structure: Can. Soc. Petroleum Geology/Geol. Assoc. Canada.
REGIONAL-SCALE GEOTHERMAL AND HYDRODYNAMIC
REGIMES IN THE ALBERTA BASIN: A SYNTHESIS

Stefan Bachu

Alberta Geological Survey, Edmonton, AB, Canada

ABSTRACT

The flow of water and heat in a sedimentary basin may be coupled through buoyancy
effects, caused by temperature variations, and through heat advection by flowing formation
waters. However, there are situations when the two processes can be partially or totally
decoupled. Such situations are when variations in formation water salinity offset variations in
water density caused by temperature differences, and when rock permeability is so low that the
velocity offluid flow does not affect significantly the conduction ofterrestrial heat through the
sedimentary succession. The role of heat advection versus heat conduction in a sedimentary
basin can be established through either numerical or dimensional analysis, based on the
geothermal and hydrodynamic characteristics of the basin or parts thereof.
The Alberta Basin in western Canada is complex in terms of hydrostratigraphy and flow
offormation waters. Several basin-scale aquitards and aquicludes separate various aquifers and
aquifer systems. On a regional scale, the flow of formation waters in aquifers is driven in
various systems and directions by past tectonic compression, erosional rebound in thick shales,
and regional- and local-scale topography. The salinity increase with depth offsets the decrease
in density resulting from temperature increase. Formation-scale rock permeability in aquifers
is low, leading to formation water velocities of the order of 10.2 m/yr. Dimensional analysis
shows that conduction dominates the terrestrial heat flow in the basin, except for the Middle
Devonian Elk Point aquifer system at the northern edge of the basin.
The thermal conductivity of the sedimentary succession increases generally eastward,
from 1.4 to 2.2 Wlm OK. The basement heat flow, calculated on the basis of rock lithology,
surface and bottom-hole temperatures, and thermal conductivity measurements, increases
generally northward, from 40 to 80 mW/m2• This corresponds on a basin scale to changes in
basement structure from old Archean rocks in the south to younger magmatic arcs in the north.
Accordingly, geothermal gradients range from less than 20 mKlm in the south to more than
45 mKlm in the north. Local-scale anomalies are superimposed on this general trend. These
anomalies are the result of basement heterogeneity, variations in radiogenic heat production
by basement rocks, and stratigraphic and lithologic variability caused mainly by erosion. Only

81
82 BACHU
at the northern edge of the basin, high permeability of the Elk Point aquifer system, caused by
reefs, dolomitization, fracturing and karst processes, leads to focused flow along the reef
barrier and to heat advection by formation waters stronger than heat conduction. The advection
of heat by formation waters discharging from this aquifer at outcrop near Great Slave Lake
explains the high geothermal gradients (> 70 mKlm) observed at shallow depths near the
Precambrian Shield at the northeastern comer of the basin featheredge.

INTRODUCTION

Sedimentary basins are of great economic importance because of their energy and mineral
resources. These are the result of a whole series of physical and geochemical processes among
which the flow of formation waters and geothermal regime playa crucial role. The flow of
formation waters may play an important role in the redistribution of terrestrial heat and
minerals, in the migration and accumulation of hydrocarbons, and in the genesis of ore
deposits. Depending on the flow strength, the geothermal regime in the basin, or parts thereof,
can be significantly distorted, with cooling effects in recharge areas, and warming effects in
discharge areas. On the other hand, understanding the geothermal regime in a sedimentary
basin is important because of the link between temperature and the various physical and
chemical processes leading to the generation of hydrocarbon resources and the genesis of
Mississippi Valley and other types of ore deposits. In some situations, the temperature
distribution in a basin may affect the flow of formation waters by inducing free convection in
a cellular pattern. This also has an effect on redistribution of terrestrial heat and of energy and
mineral resources. Thus, it is evident that the two main transport processes, fluid and heat
flows, may be either intertwined or independent of each other, depending on the stage of basin
evolution and its particular conditions and characteristics. Therefore, knowledge of both
processes and of their interdependence, or lack of, is essential to understanding the formation
of energy and mineral resources in a basin, and, therefore, it is an important exploration tool.
The geothermal regime in a sedimentary basin is influenced by the magnitude and
distribution of heat sources and by the various mechanisms for transport of the terrestrial heat
to the surface and their interaction. The main mechanisms for heat transfer in a sedimentary
basin are conduction and convection by formation waters. If the flow of formation water is
driven by external forces (forced convection or advection), then the relative importance of the
two heat transfer mechanisms is indicated by the geothermal Peelet number (Bachu, 1988)
defined as:
pwCwq D Gh
Pe*= .- (1)
Am Gv
where A. is thermal conductivity, p is density, c is specific heat, q is the Darcy velocity ofthe
water, D is the thickness of the respective aquifer, G is geothermal gradient, and the subscripts
w, m, h, and v stand, respectively, for water, saturated porous medium, horizontal and vertical.
The Darcy velocity is given in terms of hydraulic head (de Marsily, 1986) by:

kpwg
q =- - - (VHo +
~p
-Vz) (2)
~ po
where k is absolute permeability, J.l is water dynamic viscosity, g is the gravitational constant,
WI" is the freshwater hydraulic gradient driving the flow of formation waters, ap is the density
difference in the flow domain, and z is the vertical coordinate oriented upward.
GEOTHERMICS AND HYDRODYNAMICS OF ALBERTA BASIN 83

If the flow of formation water is driven by internal forces (buoyancy) in cellular


convection, then the onset of free convection takes place when the Rayleigh number Ra* for
flow in porous media, generally defined as:

Ra* =
(XW pw g k D ~T cos e (3)
JlKm I
is greater than a critical value, Rae" which depends on basin structure and boundary conditions
(Nield, 1968; Ribando and Torrance, 1976; McKibbin and hvand, 1983). In relation to (3)
a is the coefficient of water thermal expansion, Km is the thermal diffusivity of the saturated
porous medium, II T is the temperature difference between the top and the bottom of the
aquifer, and e is the angle of the aquifer with the horizontal.Where a saturated porous layer
subjected to a constant temperature difference, Raer = 39.5, whereas for a porous layer
subjected to a constant heat flux at the bottom, R~r= 27.1 (Nield, 1968; Ribando and Torrance,
1976).
Examination of relations (1) - (3) shows that rock permeability k is the most critical
parameter in determining either the strength of forced convection relative to heat conduction,
or the onset of free convection in the aquifers of a sedimentary basin. Many numerical studies
of groundwater and heat flows in sedimentary basins implicitly take into account the
interdependence between the two processes by simulating them, but seldom express explicitly
this interdependence and the factors which affect it. On the other hand, most studies of the
geothermal regime in various sedimentary basins based on observational data and their
interpretation do not take into account the relation between the two processes. Yet in other
situations, conclusions regarding the geothermal regime are drawn based on postulated
assumptions, without ground-truthing them with data from the field of basin hydrodynamics.
In such examples, wrong conclusions may be reached. However, without using sophisticated
numerical simulations, just field data and dimensional analysis, it is easy to assess if indeed
the two flow processes are interrelated or not.
The Alberta Basin in Canada is such an example where the geothermal regime was
interpreted first in isolation, based only on temperature data and a postulated model of the flow
of formation waters, as being controlled by forced convection. Subsequent geothermal and
hydrodynamic studies, to be discussed in the following, have shown that the transport of the
terrestrial heat to the surface is dominated by conduction. This paper showcases the Alberta
Basin as an example of the importance of integrating basin hydrodynamics in the study of the
geothermal regime in a sedimentary basin without having to use complex numerical models.

THE ALBERTA BASIN

The Alberta Basin, located in western Canada, is sitting on a stable Precambrian platform
and is bounded by the disturbed Foothills thrust and fold belt and Rocky Mountains to the west
and southwest, the Tathlina high to the north, by the Canadian Precambrian Shield to the
northeast, and the Williston Basin to the east and southeast (Fig. lA). A brief basin history and
geology is presented here based on Porter, Price, and McCrossan (1992) and Mossop and
Shetsen (1994). The basin comprises a wedge of sedimentary rocks increasing in thickness
from zero at the Canadian Shield in the northeast to close to 6000 m in the southwest (Figs.
IB and 2A). The basin was initiated during the late Proterozoic by rifting of the North
American craton. Subsequent thermal contraction led to the transgressive onlap of the cratonic
platform from middle Cambrian to middle Jurassic time during the passive-margin phase of
84 BACHU

basin evolution. The corresponding sedimentary succession is dominated by shallow-water


carbonates and evaporites, with some regional-scale intervening shales. Subsequent accretion
to the western margin of North America of allochtonous terranes during the Columbian and
Laramide orogenies led to isostatic flexure of the lithosphere which formed the foreland basin.
Westward dipping of the passive-margin succession from middle Jurassic to early Cretaceous
led to extensive erosion of eastwardly progressively older strata, which subcrop along the pre-
Cretaceous unconformity (Fig. lB). During the Cretaceous and early Tertiary active-margin
phase, the basin filled with synorogenic clastics, mainly shales, derived from the emerging
Cordillera. Tertiary-to-Recent erosion has removed from up to 3800 m of sediments in the
southwest to only 1000 m in the north. As a result ofthese depositional and erosional events,
the present-day topography ranges in elevation from close to 1200 m in the southwest at the
edge of the thrust and fold belt to slightly less than 200 m at Great Slave Lake in the
northeastern comer of the basin near the Precambrian Shield (Fig. 2B).

B
w Cretaceous Sub·Cretaceous
Unconformity
E

Figure 1. Major features of Alberta Basin: A, location; and B, hydrostratigraphic cross section.
GEOTHERMICS AND HYDRODYNAMICS OF ALBERTA BASIN 85

Figure 2. Characteristics of sedimentary succession in the Alberta basin: A, isopach; and B, ground surface
(isolines in m).

PREVIOUS FLUID AND HEAT-FLOW MODEL

In the 1970's and early 1980's, the accepted model of flow offonnation waters in the
Alberta Basin was that the basin-scale groundwater flow is at steady state with and driven by
the present-day topography (Fig. 3). This model was based on the work of Hitchon (1969) and
Toth (1963, 1978), who analyzed the distribution of hydraulic heads in various units in the
basin using a limited database, and promoted the idea of gravity-driven flow of fonnation
waters. In the general model of Hitchon (1969,1984), the fonnation waters in the southern and
central parts of the Alberta Basin are driven in a single basin-scale flow system from the
southwest to the northeast, with recharge in the Foothills and discharge in the Athabasca area
near the basin edge at the exposed Precambrian Shield. In the northern part of the basin, the
flow of fonnation waters, driven also by topography, is directed southwestward, toward the
same discharge area in the Athabasca region. In the extreme south, topography is driving the
flow of fonnation waters eastward from the Foothills to lake Winnipeg, across the Alberta
Basin and the Canadian part of the Williston Basin.
Anglin and Beck (1965) were the first to note, based on a few carefully measured data,
that heat generation in the Precambrian basement shows a trend of northerly increase, which
they attributed to crustal structure. In the 1980s, some authors, whose most relevant papers are
cited here, used only measured bottom-hole temperature (BHT) data (uncorrected first, and
corrected for drilling disturbances in later papers) to analyze the geothennal pattern in the
whole or various parts and stratigraphic intervals of the Alberta Basin (Majorowicz and Jessop,
1981,1993; Hitchon, 1984; Majorowicz and others, 1984, 1985; Jones, Majorowicz, and Lam,
1985). The limitations of this approach are that: (1) BHT data collected by industry are
notorious for poor quality and inconsistencies (Chapman and others, 1984); (2) there are many
uncertainties in correcting for thennal equilibrium, when these corrections are made; and (3)
there is a great variability (scatter) in data distribution and resolution both areally and with
depth because the measurements are made selectively in target areas of interest to the industry.
Calculated values for geothennal gradients were distorted further by these authors by
averaging gradients based on BHT data measured at various depths in stratigraphic intervals
86 BACHU
OCKY INTERIOR PLAINS

UNT"NS~~
d_ . . .
___ -------1
"",-
[
. nallateral flow cischarge

-----
lOCal k:Jca1oge _
local _

./

~ Main flow directions

... B
Figure 3. Diagrammatic model of assumed steady-state topography-driven flow offonnation waters in Alberta
Basin (after Hitchon, 1984): A, dip cross section; and B, plan view.

of different lithologies. In addition, some of the authors (Majorowicz and others, 1984, 1985;
Jones, Majorowicz, and Lam, 1985) smoothed the results further by using an areally moving
window of 3 x 3 townships (9.6 x 9.6 km2), and made the results more difficult to compare by
using different data controls in different profiles, and different averaging techniques.
Moreover, no documentation was provided with respect to the corrections applied, when this
was done, and for the methodology used in the statistical processing of the respective data
distributions. Nevertheless, regardless of the data distributions and methodology used, a
general pattern of increase in geothermal gradients from the south-southwest to the north-
northeast was observed by all of the mentioned authors. On the basis of the previously
described hydrodynamic model, they explained this trend as the result of heat transport by
vigorous flow of formation waters in a single regional-scale system across the entire
sedimentary succession and basin, from recharge areas in the southwest to discharge areas in
the northeast. However, neither pressure and water salinity data, nor permeability data were
used in their analysis to establish the direction and quantify the strength of groundwater flow,
and the relative importance of convective versus conductive heat transport. Their model of the
geothermal regime in the Alberta Basin was based on a simple statistical trend analysis of
uncorrected and corrected BHTs, and the postulated hydrodynamic model of Hitchon (1969,
1984). This model of convection-dominated heat transport implicitly assumes uniform or
GEOTHERMICS AND HYDRODYNAMICS OF ALBERTA BASIN 87

quasi-uniform basement heat flow. Based on available data at the time (Burwash and
Cumming, 1976), Majorowicz and Jessop (1981) tried to determine a spatial correlation
between the geothermal state of the sedimentary succession in the basin and basement heat
production, but reached the conclusion that such a relation does not exist.
Based on corrected BHTs, and pressure and permeability data for the study of the
hydrogeological and geothermal regimes in the Cold Lake and Swan Hills areas ofthe Alberta
Basin, Bachu (1985, 1988) argued that the main mechanism for heat transport in the Alberta
Basin is conduction and not advection by formation waters, because the regional-scale
permeability values for aquifers in these areas are too low (of the order of 10.15 m 2), leading
to flow velocities of the order of 10.2 m/yr. At such low velocities, dimensional analysis shows
that no significant heat transport by formation waters can take place (pe* < 0.1). Bachu (1985,
1988) linked regional- and local-scale variations in geothermal gradients, both vertically and
areally, to the lithology of the sedimentary succession, and with assumed variations in
basement heat flow. Using data of radiogenic heat production and U-Pb dates of basement
rocks, combined with the concepts of heat-flow provinces and decline of heat flow with age,
Jessop (1992) estimated the heat flow from the tectonic provinces of the Precambrian basement
of the Alberta Basin. He noted that the derived patterns of basement heat flow show little
agreement with the maps of heat flow in the Paleozoic and Mesozoic successions of the
Phanerozoic cover produced previously by Majorowicz and others (1985), and pointed out to
the effect of different data distributions and various methodologies used in processing and
interpreting the geothermal data.

HYDRODYNAMICS OF THE ALBERTA BASIN

The stratigraphy and lithology ofthe sedimentary succession in the Alberta Basin plays
an important role in determining the interplay between conductive and advective heat transfer
in the basin and in establishing its geothermal regime. Extensive, basin-scale aquitards, such
as the Ireton, Exshaw-Banff, and Cretaceous shales, and aquicludes, such as the Lotsberg, Cold
Lake, Muskeg, and Prairie evaporites, do not allow communication between various aquifers
in the sedimentary succession, except where reefs breach through these aquitards or
aquicludes, or along erosional unconformities near the eastern edge of the basin. Thus, full-
scale basin circulation across the entire sedimentary succession, of the type postulated initially
by Hitchon (1969, 1984), is precluded. Convective heat transport can take place only in
aquifers, whose thickness D ranges between several meters and up to 300 m (Mossop and
Shetsen, 1994).
Until the 1990's, what was missing from the debate regarding the influence or lack thereof
of advective heat transport in the Alberta Basin was a good knowledge and understanding of
the directions and mechanisms driving the flow of formation waters in the basin, of the
chemistry of formation waters, and of permeability distributions both areally and with depth
in the hydro stratigraphic succession. As a result of several hydrogeological regional-scale
studies (Hitchon, Bachu, and Underschultz, 1990; Bachu and Underschultz, 1993, 1995; Parks
and Toth, 1993; and Bachu, 1995, 1997), a clear image of the basin-scale flow in the Alberta
Basin emerged, totally different from the topography-driven model prevalent in the 1980's. The
flow of formation waters is more complex, both areally and stratigraphically (Fig. 4). South
of the Peace River Arch, several regional flow systems are active, driven in different directions
by different mechanisms (Bachu, 1995). A long-range flow system is active in the Upper
Devonian-to-Carboniferous succession, driven north-northeastward by basin-scale topography,
88 BACHU

from a recharge area at outcrop in Montana in the south to a discharge area in northeastern
Alberta (Athabasca region) near the Precambrian Shield (Bachu and Underschultz, 1993,
1995). An underlying flow system in Cambrian and Middle Devonian aquifers, driven east-
northeastward from the Rocky Mountains in the southwest to the Precambrian Shield in the
east, most probably is driven by past tectonic compression during the Laramide Orogeny
(Bachu, 1995). Finally, inward flow systems are present in the overlying Cretaceous and
Tertiary aquifers, driven toward the thrust and fold belt by erosional rebound in thick
Cretaceous and Tertiary shales (Bachu and Underschultz, 1995). Hydraulic heads in these
aquifers reach values near the thrust and fold belt lower than the lowest elevation in the basin
at Great Slave Lake, situated at more than 1,500 km to the northeast. The erosional shale
rebound, leading to these flow systems, is the result of Tertiary-to-Recent erosion in southern
Alberta of up to 3,800 m of sediments (Bustin, 1992). Local-scale flow systems are present at
the top of the sedimentary succession, driven by local topography. The complex pattern of
formation-water flow south of Peace River Arch shows that the flow systems did not reach yet
steady state and are evolving toward reaching equilibrium with the present-day boundary
conditions.

A
Local-scale lopography dn'o'(l:n Row
RegIOnal-scale rlow doveo by erOSlooal rebound
BaSin-scale lopography--drlvon flow
Basin-scale !low of leclonlc origin with sirong
buoyancy cnecls
Am.basca and Cold Lake 0<1 SandS

w E

== = MIXing zone
Basln·scale lopoglaphy dnven
!low normal 10 the plane of cross section
B
Figure 4. Diagrammatic representation of main fluid-flow systems in Alberta Basin: A, in plan view; and B,
in cross section (see Fig. I for location of line of cross section W-E).
GEOTHERMICS AND HYDRODYNAMICS OF ALBERTA BASIN 89

North of the Peace River Arch, the flow pattern is simpler (Bachu, 1997), but different
from the one postulated by Hitchon (1984). Namely, the flow of formation waters in the entire
Phanerozoic succession is driven northeastward by regional-scale topography, from a recharge
area near the fold belt and Bovie Lake fault, to discharge at aquifers outcrop at Great Slave
Lake (Bachu, 1997). Because only up to I km of sediments were removed by Tertiary-to-
Recent erosion (Kalkreuth and McMechan, 1988), the process of erosional rebound in the thick
Cretaceous shales ended in this area, and the flow systems are at steady state with the present-
day topography.
The chemistry of formation waters in the Alberta Basin (Hitchon, Bachu, and
Underschultz, 1990; Bachu, 1995, 1997) shows that the connate waters in Paleozoic aquifers
(corresponding to the platform-margin in basin development) are generally saline, except for
small regions in recharge and discharge areas where water of meteoric origin enters the system.
This indicates incomplete flushing of the original basinal waters, consistent with low flow
velocities which would not distort the conductive heat regime in a sedimentary basin (Deming
and Nunn, 1991). In contrast, the formation waters of meteoric origin in Mesozoic aquifers
(which correspond to the foreland stage in basin evolution), are less saline. A zone of mixing
(Fig. 4B) occurs along the sub-Jurassic and sub-Cretaceous unconformities (Bachu, 1995). The
high salinity of Paleozoic formation waters overcomes the decrease in density caused by
increased temperature with depth. Thus, the Paleozoic formation waters in the Alberta Basin
are heavier than the Mesozoic waters, particularly in areas where Lower and Middle Devonian
evaporite beds are present. This difference in salinity adds another component to the flow of
formation waters in that buoyancy plays an important role, in this case impeding full-scale
water circulation from the surface to the basement.

GEOTHERMAL REGIME IN THE ALBERTA BASIN

As described previously, the Alberta Basin is sitting on a stable crystalline Precambrian


basement. No volcanic centers were active in the past nor are present in the basin. Tectonic
activity in the basin also is limited (Burwash, McMechan, and Potter, 1994). According to
Ross, Broome, and Miles, (1994), the basement is comprised of Archean rocks in the south,
younger accretted terranes, Proterozoic plutonic and metaplutonic rocks, and magmatic arcs
in the north. After the rifting in late Proterozoic which initiated the basin, and compression in
Jurassic-Cretaceous time, the most recent thermal activity is related to an Eocene tectono-
extensional event at 50 Ma localized at the southern edge of the basin. The effects ofthermal
events older than 100 Ma have already dissipated, whereas the effects of the Eocene event, if
present yet, are local and limited in nature. Based on the maximum thickness of the
sedimentary cover, the characteristic time for propagation to the surface of basement thermal
events is less than 3 Ma. Thus, the heat flow from the basement is at steady or quasi-steady
state, depending on location. Based on basin history and current conditions, it follows that the
two main sources of heat are the mantle heat flow and the heat generated internally in the crust
by the decay of radioactive elements. All other heat sources are either absent or minor
(Rybach, 1981; Hitchon, 1984; Bachu, 1993). Having established the main sources of heat
flow in the basin, the next step in the analysis of the geothermal regime is to determine the
relative importance of the two heat-transport mechanisms: conduction through the entire
sedimentary succession, and convection (forced and free) in aquifers. Further evaluation of the
relative importance of convective versus conductive heat transport needs estimating regional-
90 BACHU

scale representative values for the aquifer hydraulic and geothennal parameters in the
expressions of the Peelet and Rayleigh numbers (relations 1-3).

Rock Permeability and Porosity

On a regional scale, the rock penneability k of aquifers in the Alberta Basin is low, of the
order of 10-15_10- 14 m2 (Bachu, 1985, 1988, 1997; Hitchon, Bachu, and Underschultz, 1990;
Bachu and Cao, 1992; Bachu and Underschultz, 1992, 1993), notwithstanding the even lower
permeability of the intervening shaly aquitards and evaporitic aquieludes. This low
penneability, in conjunction with regional-scale hydraulic gradients V'H of the order of 1-10
mIkm, leads to low fluid-flow velocities, of the order of 1O-QO-1 rn/yr. Only along the
Presqu'ile barrier reef in Middle Devonian carbonates along the northern edge of the basin,
permeability values in the reef complexes are high (up to 10- 12 m2; Bachu, 1997), leading to
relatively high flow velocities (of the order of 1 m1yr), capable of focusing the flow of
formation waters along the barrier reef. The porosity cp of aquifer rocks ranges between 2 and
35 % (Bachu and Underschultz, 1992,1993; Bachu, 1997).

Effective Thermal Conductivity and Heat Capacity

Based on complete logs of lithological analysis, the local thennal conductivity A,,20 of
sedimentary rocks comprising fractions f; ofN lithological types An was calculated from rock
thennal conductivity measurements performed at laboratory conditions for various lithologies
(with values ranging between 1.2 W/m K for shales and 5.8 W/m K for anhydrite; Bachu,
1993), using the geometric average (Chapman and others., 1984; Bachu, 1991, 1993):
N

A..20 = II Anfi (4)

These values were corrected for in-situ conditions at the temperature T eC) corresponding to
the burial depth according to the relation (Chapman and others., 1984):
A. = A.,.20 [293/(T+273)] (5)

Finally, the thennal conductivity A,., of the water-saturated porous medium was calculated also
as the geometric average of the two components of the binary system (Chapman and others.,
1984; Bachu, 1991):
(6)

after correcting the water thennal conductivity Aw for variations with temperature (depth) based
on the relation (Deming and Chapman, 1988):
Aw =0.5648 + 1.878 x 10-3 T -7.231 x 10-6 T2 for T ~ 137°C (7)
which corresponds to the temperature range encountered in the Alberta Basin.
Using this methodology, surface-to-basement logs ofthennal conductivity variations with
depth were constructed for 1453 selected wells (one per township, or 9.6 x 9.6 km2) which
reach the Precambrian crystalline basement and for which complete lithological logs were
available (Bachu, 1993).
The heat capacity (pc)m of the saturated porous medium was calculated as the porosity-
weighted average of water and rock heat capacities (Cheng, 1978):
GEOTHERMICS AND HYDRODYNAMICS OF ALBERTA BASIN 91

(pc)m = ql (pc)w + (l-ql ) (pc), (8)

where c is specific heat. Literature values were used for rock specific heat (de Marsily, 1986)
and density (Daly, Manger, and Clark, 1966), resulting in values for (pc)m of the order of2.3
x 106 J/m3 K.

Dimensional Analysis

The thermal diffusivity of the saturated porous medium, K,n, was calculated according to
its definition as the ratio of thermal conductivity to heat capacity. The coefficient of water
thermal expansion, IXw, ranges from 0.1- 1.0 x 10-4 K-I for temperatures encountered in the
Alberta Basin (de Marsily, 1986). Finally, the water dynamic viscosity /l and density Pw at in
situ temperature and salinity conditions were calculated according to the values published by
Kestin, Khalifa, and Correia, (1981) and Rowe and Chou (1970), respectively. They range
between 300 and 1300 /lPaos for viscosity, and 990 and 1350 kglm3 for density, the former
being influenced mainly by temperature and the second by salinity, particularly in the vicinity
of evaporitic beds. Finally, regional-scale geothermal characteristics such as gradients Gv and
Gh across and along aquifers are on the order of30 mKlm and 0.2 mKlm, respectively (Bachu,
1985, 1988; Bachu and Cao, 1992).
Using the given characteristic values for the parameters needed in dimensional analysis,
the following regional-scale values were obtained for the geothermal Peelet and Rayleigh
numbers representative for aquifers in the Alberta Basin:

Pe* < 0.01 and Ra* < 0.1

which indicate that heat transport by forced convection (flow of formation waters in aquifers)
is negligible with respect to heat conduction through the sedimentary succession, and that
aquifer permeability is too low for free convection to develop in a cellular pattern (Ra* < Ra.,).
With respect to the high-permeability Presqu'ile reef barrier at the northern edge of the basin,
where permeability reaches formation-scale values of 10- 12 m 2, the corresponding Peelet and
Rayleigh numbers are of the order of:

Pe* '" 4 andRa*:;;; 1

This indicates that, although cellular free-convection flow would not develop in this aquifer,
the lateral transport of terrestrial heat by the flow offormation water is significant, at least of
the same order as heat conduction, thus locally distorting the conductive geothermal pattern.
It must be emphasized here that all the parameter values and dimensionless numbers Pe* and
Ra* are representative for the Alberta Basin at a regional- to basin-scale. Locally, areas of high
permeability, high hydraulic gradients or even high geothermal gradients may lead to local-
scale disturbances ofthe conductive heat flow by convection of formation waters.

Relation between Basin Hydrodynamics and Heat Transport

The analysis of the hydrodynamic regime of formation waters in the Alberta Basin, and
the dimensional analysis of convective versus conductive heat transport in the basin show that
the previous basin-scale model of topography-driven flow of formation waters and convective
92 BACHU

heat transport from recharge in the southwest to discharge in the northeast is not correct for the
following reasons:
(1) The velocity of formation waters is too low, because of low regional-scale
permeability, such that advective heat transport is negligible in comparison with conductive
heat transfer, except only for the Presqu'ile reef barrier at the northern edge of the basin. The
low permeability of rocks in aquifers precludes also the onset of free convection in a cellular
pattern.
(2) The rebounding Cretaceous and Tertiary strata in southwestern Alberta act as sinks
for any meteoric waters recharging the basin in this region, impeding their reaching of deeper
aquifers. In addition, the flow in the aquifers in this succession is oriented downdip,
southwestward. Thus, these waters do not flow and carry heat on a basin scale from the
southwest to northeast, regardless of their velocity (although too low).
(3) Extensive, basin-scale aquitards and aquicludes also impede deep vertical flow and
penetration, hence cooling and subsequent heating, of any meteoric water recharging the basin.
(4) The salinity difference between Paleozoic and Mesozoic formation waters leads to
"negative" buoyancy which also impedes the deep penetration of cool, meteoric water where
it would be heated and carry out the terrestrial heat toward discharge areas in the northeast.
(5) The various basin- and regional-scale flow systems carry the formation waters in
different directions, such that there is no single flow system carrying heat from the southwest
to the northeast, as previously postulated.
With respect to the relation between basin-scale hydrodynamics and geothermal regime
in the Alberta Basin, the main conclusion of this analysis is that the two are independent, and
that the main mechanism for the transport of terrestrial heat from the basement to the surface
is vertical conduction through the sedimentary succession. As a result, geothermal gradients
and basement heat flow can be calculated using a simple one-dimensional conductive model,
given the layered structure of the basin and the absence of significant short-scale topographic
variations (like in mountain regions) which would introduce heat-refraction effects.

Geothermal Pattern and Basement Heat Flow

Average geothermal gradients across the entire sedimentary succession were calculated
based on multiannual ground-surface temperatures (Fig. 5A) and temperatures at the top of the
Precambrian basement (Fig. 5B). Although geothermal gradients based on temperature
measurements at both ends of a stratigraphic succession do not explicitly take into account the
lithological variations within that interval, they represent a true weighted-average of the
geothermal gradients in each individual lithologic unit in the respective succession (Bachu,
1985). Ground-surface temperature values were calculated based on multi annual air-
temperature averages recorded at climate stations across the basin, after applying a correction
to account for the ground snow-cover during winter (Bachu and Burwash, 1991). The
temperature distribution at the bottom of the sedimentary succession was constructed based
on BHTs measured immediately below the crystalline Precambrian surface in the 1453 control
wells distributed across the basin which reach the basement (Bachu and Burwash, 1991). For
1086 wells with multiple BHT readings, the formation temperature was individually estimated
in each well by applying the Homer method to the raw data (Chapman and others, 1984; Bachu
and Burwash, 1991). For the 367 wells with a single temperature measurement, a statistical
correction with depth was applied based on regression analysis of the multiple-readings
temperature measurements (BHT correction versus depth; Bachu and Burwash, 1991).
Different statistical corrections apply in the southern, north-central, and northern parts of the
GEOTHERMICS AND HYDRODYNAMICS OF ALBERTA BASIN 93

basin (Bachu and Burwash, 1991). The ground-surface temperature distribution (Fig. 5A)
exhibits an expected pattern consistent with latitude and altitude variations (Fig. 2B), whereas
the temperature distribution at the top of the Precambrian basement (Fig. 5B) presents
generally a pattern consistent with the westward dip of the basin (Fig. 2A) and local
topographic variations (Fig. 2B). A few local-scale anomalies, such as the one in the southeast
near the Bow Island Arch (Fig. 5B) are related to local high heat-generation by radioactive
decay in the crystalline basement (Bachu and Burwash, 1991). A significant geothermal
anomaly near Great Slave Lake in the northeast, where temperatures in the 40°C range were
measured at shallow depths « 600 m) is the result of convective effects of heat transport by
formation water in the highly permeable Presqu'ile reef barrier aquifer which discharges near
the lake under a thin veneer of unconsolidated Quaternary sediments.

A B

--~'- - ------
USA ..oN1',,"A USA
MONT"'~

Figure 5. Main characteristics of geothennal regime in Alberta basin: A, multiannual ground-surface temperature
(OC); B, temperature distribution at top of Precambrian basement eC); C, geothermal gradients (mKlm); and D,
effective thennal conductivity of sedimentary succession (W/m K)
94 BACHU

The distribution of geothermal gradients (Fig. 5C) shows a general trend of increasing
values from less than 20 mKlm in the south to more than 40 mKlm in the north, consistent
with the general trend observed by previous authors. The differences in detail between this
map and previously published maps stem from using different data distributions, and different
methods for BHT correction and statistical processing. Local-scale anomalies are better
evidenced here than by the temperature distribution (Fig. 5B) where the temperature increase
with depth plays a masking effect, and are all the result of variability in radioactive heat
generation at the top of the basement (Bachu and Burwash, 1991). The geothermal anomaly
near Great Slave Lake is caused by local convective heat transport in the Presqu'ile reef barrier
and reaches values in the 70 mKlm range (Fig. 5C). Although all the studies to date of the
geothermal regime in the Alberta Basin indicate the north-northeastward trend of increasing
geothermal gradients, they differ in: (1) data distributions and methods used for processing;
(2) resolution and detail; and (3) interpretation of the observed patterns.
In a steady- or quasi-steady-state conductive heat transfer problem, the temperature
distribution, hence geothermal gradients, inside any domain of interest are controlled by the
boundary conditions, and heat generation and variations in thermal conductivity inside the
domain. For sedimentary basins, the respective boundary conditions are the temperature at the
top ofthe succession and the heat flow at the bottom of the succession. In the inverse problem,
the heat flow at the base of the succession can be calculated from knowledge of the top
boundary condition (ground-surface temperature), distributions of geothermal gradients and
thermal conductivity, and location and strength of internal heat sources caused by radiogenic
decay mainly in shales. Another way to calculate the heat-flow distribution at the base of the
sedimentary succession would be to calculate local geothermal gradients in the same well
using BHTs from different depths in the basement, close to its top, and thermal conductivity
values for the respective rocks. Unfortunately, such data are not available because of the depths
involved and lack of industry interest in collecting these data.
In order to estimate the basement heat flow, the well-scale thermal conductivity of the
sedimentary succession was calculated using the Bullard method (harmonic average) for the
selected 1453 wells which reach the Precambrian basement, based on the thermal conductivity
distribution across the sedimentary succession in each well. The resulting effective thermal
conductivity of the entire sedimentary succession (Fig. 5D) ranges between 1.4 W/m K in the
northwest, where shales are predominant, to 2.2 W/m K in the east, where the post-Jurassic
siliciclastic cover is almost absent because of Tertiary-to-Recent erosion, and higher
conductivity carbonates and evaporites dominate the sedimentary succession (Bachu, 1993).
Local-scale variations in effective thermal conductivity are the result of local depositional and
erosional events which affected the lithology of the sedimentary succession.
The conductive surface heat-flow density in the Alberta Basin was calculated everywhere
except in the vicinity of Great Slave Lake (because of the convective heat-transport effects)
based on the distribution of geothermal gradients and effective thermal conductivity. Because
measurements of radiogenic heat production in the sedimentary rocks in the basin are missing,
literature values (Rybach, 1988) were used to estimate the heat production in the sedimentary
column and its areal distribution. The basement heat flow at the top of the Precambrian was
estimated by subtracting the heat generated in the sediments from the surface heat flow. Its
distribution (Fig. 6A) shows a regional-scale northward trend of increasing values from less
than 40 mW/m2 in the south to values in the 50-60 mW/m2 range in the north, with local
values reaching 70-80 mW/m2 • This trend is consistent in a broad sense with the basement age
and structure (Fig. 6B), which is comprised of old Archean rocks in the south and younger
plutonic rocks and magmatic arcs in the north (Ross, Broome, and Miles, 1994). This broad
GEOTHERMICS AND HYDRODYNAMICS OF ALBERTA BASIN 95

correlation trend was noted also by Jessop (1992). A detailed statistical analysis of the degree
of correlation between basement heat flow and structure is not possible because ofthe different
data distributions and resolution of determining the two, that is aeromagnetic surveys and
broad delineation for structure (Ross, Broome, and Miles, 1994), and point (well) data for heat
flow. Local-scale thermal anomalies superimposed over this general trend are the result of
basement heterogeneity (Bachu, 1993) and variable basement heat generation caused by the
radiogenic decay ofU and Th, and K metasomatism (Bachu and Burwash, 1991). It is worth
noting here the total dissimilitude between the regional-scale patterns of formation-water flow
(Fig. 4) and geothermal regime (Figs. 5C and 6A) in the basin.

--""'"---,-----
USA MONTANA
U.SA MONTANA

Arcnean
0 [Z2;l Magmallc arcs

D Accretod
lerranes GSLSZ Greal Stave Lake
Sheaf Zona

Figure 6. Characteristics of Precambrian basement underneath Alberta Basin: A, heat flow (mW/m2); and B,
structure (after Ross, Broome, and Miles, 1994).

CONCLUSIONS

In assessing the geothermal regime in a sedimentary basin, it is important to consider all


possible sources of terrestrial heat and mechanisms for heat transport to the surface.
Particularly, the interplay between conductive and convective heat transport can be evaluated
using either numerical simulations or dimensional analysis. The latter is particularly well
suited for situations of complex hydrodynamic systems, extremely variable rock properties and
a multitude of factors influencing both the hydrodynamic and geothermal regimes in a basin,
situations which are too difficult to be treated numerically unless significant simplifying
assumptions are made. A proper dimensional analysis is based on a good understanding of
flow processes in a sedimentary basin, and on a good data set which allows a proper estimation
of the numerical values ofthe parameters involved in the analysis.
Based on the wealth of data available for the Alberta Basin and on recent studies of basin
hydrodynamics and geothermics, application of dimensional analysis to heat transport
processes in the basin shows that:
96 BACHU

(1) Because of low rock permeability, the main, basin-scale mechanism for heat transport
in the basin is conduction through the entire sedimentary succession. Only at the northern edge
of the basin, high permeability of the rocks in the Presqu'ile reef barrier leads to convective
heat transport in this aquifer, with the result of a significant thermal anomaly (high geothermal
gradients) at aquifer discharge near Great Slave Lake.
(2) At the basin-scale, the south-northward increase in terrestrial heat flow and associated
geothermal gradients most probably corresponds to changes in basement structure from older
and "cooler" Archean rocks in the south, to younger and "hotter" rocks in the north.
(3) The stratigraphy and lithology of the sedimentary cover plays an important role in
establishing the geothermal pattern in the basin because of variability in thermal conductivity
and thickness of various strata.
(4) Local-scale anomalies in the basin-scale geothermal pattern are caused by
heterogeneity in basement structure and heat generation, to local changes in the stratigraphy
and lithology of the Phanerozoic cover, and to local topographic variations. Beside the thermal
anomaly at Great Slave Lake, the flow of formation water does not affect the terrestrial heat
flow. Local influences are possible in zones of high permeability, but these are undetectable
with the existing data resolution and are insignificant on a regional scale.
This example of application of hydrodynamics, geothermics and dimensional analysis can
be applied to any sedimentary basin for assessing the importance of conductive versus
convective heat transport in the basin, and, consequently, of basin thermal history and
maturation. Furthermore, it shows that a multidisciplinary approach, use of various data
categories, and data quality and distributions play an important role in the study of flow and
transport processes in sedimentary basins, and in developing sound and consistent conceptual
models oftheir hydrodynamic and geothermal regimes.

REFERENCES

Anglin, F.M., and Beck, A.E., 1965, Regional heat flow pattern in western Canada: Can. Jour. Earth. Sci., v. 2,
no. 3, p. 176-182.
Bachu, S., 1985, Influence of lithology and fluid flow on the temperature distribution in a sedimentary basin: a
case study from the Cold Lake area, Alberta, Canada:Tectonophysics, v. 120, no. 3-4, p. 257-284.
Bachu, S., 1988, Analysis of heat transfer processes and geothermal pattern in the Alberta basin, Canada: Jour.
Geophys. Res., v. 93, no. B7, p. 7767-7781.
Bachu, S., 1991, On the effective thermal and hydraulic conductivity of binary heterogeneous sediments:
Tectonophysics, v. 190, no. 2-4, p. 299-314.
Bachu, S., 1993, Basement heat flow in the Western Canada Sedimentary Basin: Tectonophysics,
v. 222, no. 1, p. 119-133.
Bachu, S., 1995, Synthesis and model of formation water flow in the Alberta basin, Canada: Am. Assoc.
Petroleum Geologists Bull., v. 79, no. 8, p.1159-1178.
Bachu, S., 1997, Flow of formation waters, aquifer characteristics, and their relation to hydrocarbon
accumulations in the northern part of the Alberta basin: Am. Assoc. Petroleum Geologists Bull., v. 81, no.
5, p. 712-733.
Bachu, S., and Burwash, R. A., 1991, Regional-scale analysis of the geothermal regime in the Western Canada
Sedimentary Basin: Geothermics, v. 20, no. 5/6, p. 387-407.
Bachu, S., and Cao, S., 1992, Present and past geothermal regimes and source rock maturation, Peace River arch
area, Canada: Am. Assoc. Petroleum Geologists Bull., v. 76, no. 10, p. 1533-1549.
Bachu, S., and Underschu1tz, J. R., 1992, Regional-scale porosity and permeability variations, Peace River arch
area, Alberta, Canada: Am. Assoc. Petroleum Geologists Bull., v. 76, no. 4, p. 547-562.
Bachu, S., and Underschultz, J. R., 1993, Hydrogeology of formation waters, northeastern Alberta basin: Am.
Assoc. Petroleum Geologists Bull., v.77, no. 10, p. 1745-1768.
GEOTHERMICS AND HYDRODYNAMICS OF ALBERTA BASIN 97
Bachu, S., and Underschultz, J. R, 1995, Large-scale erosional underpressuring in the Mississippian-
Cretaceous succession, southwestern Alberta basin: Am. Assoc. Petroleum Geologists Bull., v. 79, no. 7,
p.989-1004.
Burwash, RA, and Cumming, G. L., 1976, Uranium and thorium inthePrecarnbrianbasementofwestern
Canada, I. Abundance and distribution: Can. Jour. Earth Sci., v. 13, no. 2, p. 284-293.
Burwash, RA, McMechan, M. E., and Potter, D.E.G., 1994, Precambrian basement beneath the Western Canada
Sedimentary Basin, in Mossop, G.D., and Shetsen, I., compilers, Geological Atlas of the Western Canada
Sedimentary Basin: Can. Soc. Petroleum Geologists and Alberta Research Council, Calgary, p.49-56.
Bustin, R.M., 1992, Organic maturation of the western Canada sedimentary basin: Intern. Jour. Coal Geology,
v. 19, no. 1-4, p. 319-358.
Chapman, D. S., Keho, T. H., Bauer, M. S., and Picard, M. D.,1984, Heat flow in the Uinta basin determined
from bottom hole temperature (BHT) data: Geophysics, v. 49, no. 4, p. 453-466.
Cheng, P., 1978, Heat transfer in geothermal systems: Advances in Heat Transfer, v. 14, p. 1-105.
Daly, R.A., Manger, E.G., and Clark, S.P., 1966, Density of rocks, in Clark, S.P., ed., Handbook of
Physical Constants: Geol. Soc. America Mem. 97, p. 19-26.
de Marsily, G., 1986, Quantitative hydrogeology: Academic Press: San Diego, 440 p.
Deming, D., and Chapman, D. S., 1988, Heat flow in the Utah-Wyoming thrust belt from analysis of bottom hole
temperature data measured in oil and gas wells: Jour. Geophys. Res., v. 93, no. Bll, p. 13,657-13,672.
Deming, D., and Nunn, J.A., 1991, Numerical simulations of brine migration by topographically
driven recharge: Jour. Geophys. Res., v. 96, no. B2, p. 2485-2499.
Hitchon, B., 1969, Fluid flow in the Western Canada Sedimentary Basin, 1. Effect of topography: Water
Resources Res., v. 5, no. 1, p. 460-469.
Hitchon, B., 1984, Geothermal gradients, hydrodynamics and hydrocarbon occurrences, Alberta, Canada: Am.
Assoc. Petroleum Geologists Bull., v. 68, no. 4, p. 713-743.
Hitchon, B., Bachu S., and Underschultz, J.R, 1990, Regional subsurface hydrogeology, Peace River arch area,
Alberta and British Columbia: Bull. Can. Petroleum Geology, v. 38A, p. 196-217.
Jessop, A.M., 1992, Thermal input from the basement of the Western Canada Sedimentary Basin: Bull. Can.
Petroleum Geology, v. 40, no 3, p. 198-206.
Jones, F.H., Majorowicz, JA., and Lam, H.L., 1985, The variation of heat flow density with depth in the prairies
basin of western Canada: Tectonophysics, v. 121, no.l, p. 35-44.
Kalkreuth, W., and McMechan, M.E., 1988, Burial history and thermal maturity, Rocky Mountain front ranges,
foothills and foreland, east-central British Columbia and adjacent central Alberta, Canada: Am. Assoc.
Petroleum Geologists Bull., v. 72, no. 11, p. 1395-1410.
Kestin, J., Khalifa, H.E., and Correia, R.J., 1981, Tables of the dynamic and kinematic viscosity of aqueous
NaCI solutions in the temperature range 20-150 DC and the pressure range 0.1-35 Mpa: Jour. Physics and
Chemistry Reference Data, v. 10, no. 1, p. 71-87.
Majorowicz, J.A., and Jessop, AM., 1981, Regional heat flow patterns in the Western Canada Sedimentary
Basin: Tectonophysics, v. 74, no. 3-4, p. 209-238.
Majorowicz, J.A, and Jessop, A.M., 1993, Relation between basement heat flow and the thermal state of the
sedimentary succession of the Alberta plains: Bull. Can. Petroleum Geology, v. 41, no. 3, p. 358-385.
Majorowicz, J.A., Jones, F.A, Lam, H.L., and Jessop, AM., 1984, The variability of heat flow both regional and
with depth in southern Alberta, Canada: effect of groundwater flow?: Tectonophysics, v. 106, no. 1-2, p.
1-29.
Majorowicz, J.A, Jones, FA., Lam, H.L., and Jessop, AM., 1985, Regional variations of heat flow with depth
in Alberta, Canada: Geophys. Jour. Roy. Astr. Soc., v. 81, no. 2, p. 479-487.
Mckibbin, R, and Tyvand, PA., 1983. Thermal convection in a porous medium composed of alternating thick
and thin layers: Intern. Jour. Heat and Mass Transfer, v. 26, no. 5, p. 761-780.
Mossop, G.D., and Shetsen, I., compilers, 1994, Geological atlas of the Western Canada Sedimentary Basin:
Can. Soc. Petroleum Geologists and Alberta Research Council, Calgary, 510 p.
Nield, D.A, 1968, Onset oftherrnohaline convection in a porous medium: Water Resources Res., v. 4, no. 3, p.
553-559.
Parks, K.P., and Toth, J., 1993, Field evidence for erosion-induced underpressuring in Upper Cretaceous and
Tertiary strata, west central Alberta, Canada: Bull Can. Petroleum Geology, v. 43, no. 3, p. 281-292.
Porter, J.W., Price, R.A, and McCrossan, R.G., 1982, The Western Canada Sedimentary Basin: Phil. Trans. Roy.
Soc. London, v. A305, no 1489, p. 42-48.
98 BACHU
Ribando, R.J., and Torrance, K.E., 1976, Natural convection in a porous medium: effects of confmement,
variable permeability and thermal boundary conditions: Jour. Heat Transfer, Trans. ASME, v. 98, no. 2,
p.42-48.
Ross, G.M., Broome, J., and Miles, W., 1994, Potential fields and basement structure - Western Canada
Sedimentary Basin, in Mossop, G.D., and Shetsen, I., compilers, Geological Atlas of the Western Canada
Sedimentary Basin: Can. Soc. Petroleum Geologists and Alberta Research Council, Calgary, p. 41-48.
Rowe, A.M., and Chou, J.C.S., 1970, Pressure-volume-temperature-concentration relation of aqueous NaCI
solutions: Jour. Chem. Eng. Data, v. 15, no. 1, p. 61-66.
Rybach, L., 1981, Geothermal systems, conductive heat flow, geothermal anomalies, in Rybach, L., and Muffier,
L.P .J., eds., Geothermal Systems: Principles and Case Histories: John Wiley & Sons, New York, p. 3-76.
Rybach, L., 1988, Determination of heat production rate, in Haenel, R., Rybach, L., and Stegena, L., eds.,
Handbook of Heat-Flow Density Determinations: Kluwer Academic, Dordrecht, p. 125-142.
Toth, J., 1963, A theoretical analysis of groundwater flow in small drainage basins: Jour. Geophys. Res., v. 68,
no. 16, p. 4795-4812.
Toth, J., 1978, Gravity-induced cross-formational flow of formation fluids, Red Earth region, Alberta, Canada:
analysis, patterns and evolution: Water Resources Res., v. 14, no. 5, p. 805-843.
BASIN-SCALE GROUNDWATER FLOW AND ADVECTIVE HEAT FLOW:
AN EXAMPLE FROM THE NORTHERN GREAT PLAINS

William D. Gosnold, Jr.

Department of Geology and Geological Engineering


University of North Dakota
Grand Forks, ND 58202

ABSTRACT

Gravity-driven groundwater flow in a confined aquifer system that extends several


hundred kilometers eastward from the Black Hills causes anomalous surface heat-flow over
an 80,000 km2 area in southern South Dakota and northern Nebraska. Data from 16 new heat-
flow holes, existing heat-flow data, and heat-flow values calculated from BHT data show a
systematic variation in heat-flow from 20 mW m-2 in the recharge zone to 140 m W m- 2 in the
discharge zone. Ninety-four conventional heat-flow values and 62 heat-flow values calculated
from BHT data yield an average heat-flow of 58 ± 9 mW m- 2 for the northern Great Plains
exclusive of the anomalous area. The advective heat-flow system is unusual in that
temperature gradients in boreholes ranging from 2000 meters deep near the Black Hills to 500
meters deep in central South Dakota show only conductive heat-flow. In effect, the advective
system is masked by 500 to 2000 meters oflow permeability marine shales that overlie a 600-
meter thick confined aquifer system. Numerical models of coupled groundwater heat-flow in
the aquifer system suggest that confined water flow at Darcy velocities of 3.17 x 10-8 m S-I to
6.34 x 1O-8 m S-I (1 to 2 my-I») causes the anomalous heat-flow.

INTRODUCTION

The empirical linear relation between heat flow and heat generation (Birch and others,
1968)
Q=Q*+Ab (1)

provides a basis for delineating heat-flow provinces with characteristic values of reduced heat-
flow, Q., and the radiogenic layer parameter, b (Roy and others, 1968; Roy, Blackwell, and
Decker, 1972). In this relationship, Q represents surface heat flow determined from borehole

99
100 GOSNOLD

data, Ab is the product of radioactive heat production at the surface and the slope of linear
regression, and Q* is heat flow from below the heat producing zone. Equation (1) applies
fairly well in the tectonically stable region east of the Rocky Mountains where average surface
heat flow, Q is 51 ± 20 m W m- 2 (n = 530), reduced heat flow is 31 ± 1 m W m-2 (n = 20), and
the radiogenic layer parameter is 8.4 ± 0.4 km (Morgan and Gosnold, 1989). However, the
relationship clearly does not apply in the tectonically active Rocky Mountain provinces where
average surface heat-flow is 149 ± 181 mW m- 2 (n = 110) (Morgan and Gosnold, 1989).
Variability in surface heat flow in tectonically stable provinces results from variation in the
crustal radiogenic component of heat-flow and may be related either to crustal thickness or to
the thickness of the upper crustal radioactive layer (Morgan and Gosnold, 1989). However,
some of the heat-flow variability in the Great Plains province, which has the highest average
heat flow east of the Rocky Mountains (66 ± 26 mW m- 2 , n = 87), has been attributed to
advective heat transport by groundwater flow (Gosnold, 1985, 1990).
A recent analysis of heat flow in the Great Plains, mainly based on bottom-hole
temperature data, suggested that a large region in southern South Dakota and northern
Nebraska is characterized by heat flows greater than 100 mW m- 2 (Gosnold, 1991). If the
anomalously high heat flow is the result of regional groundwater flow, it may be possible to
quantify the advective component by making heat-flow measurements at locations that test the
advection model. In this paper, I present new heat-flow data from sites located specifically to
test the advection hypothesis and use model simulations of the system to quantify the advective
and conductive heat-flow components in the Great Plains. First, heat-flow data and geothermal
gradients acquired during continental heat flow and geothermal resource investigations are
combined to better delineate surface heat flow in the region. Next, the relevant aspects of the
regional groundwater systems are described in the context of an advective heat-flow system.
Finally, analytical and numerical models coupling heat flow and groundwater flow are tested
against both heat flow and hydrogeologic data.

THE HEAT-FLOW ANOMALY

The heat-flow anomaly was recognized first as a "Hot Water Belt" covering an area of
about 7,200 km2 in South Dakota (Fig. 1) and initially was delineated by about 200
temperature gradients calculated from temperatures measured in artesian water wells
(Adolphson and LeRoux, 1968). Later, Schoon and McGregor (1974) compiled temperature
data from more than 1000 wells, including surface temperatures of artesian wells and bottom-
hole temperatures obtained in oil exploration wells, and produced a geothermal gradient map
of South Dakota that shows higher-than-average geothermal gradients occurring over a large
area in South Dakota with apparent extension into north-central Nebraska (Fig. 1). Prior to
this study, 19 heat-flow measurements were made in or close to the anomalous region. Eleven
measurements within the anomaly, excluding the Black Hills Uplift, average 93.5 ± 15.9 mW
m- 2, and eight measurements on the margin of the anomaly average 62.8 ± 5.8 mW m- 2
(Gosnold, 1990). The general pattern of the heat-flow anomaly delineated by these combined
data and published analyses of regional groundwater systems (Downey, 1986) led to earlier
interpretations of an advective heat-flow anomaly (Gosnold, 1985, 1990).
BASIN-SCALE GROUNDWATER FLOW AND ADVECTIVE HEAT FLOW 101

km
Figure I. Geothermal gradient map of South Dakota modified from Schoon and McGregor (1974). Area within
bold line is the "hot water belt" identified by Adolphson and LeRoux (1968). Temperature gradients are °C
lan-I.

NEW HEAT-FLOW DATA

Heat-Flow Sites

In this study, thirteen specially completed heat-flow holes and three holes of
opportunity were occupied to test the advective heat-flow model (Filg. 2). The advective
system is a group of four confined aquifers that recharge on the eastern side of the Black Hills
and discharge over a broad zone some 200-400 km to the east in south-central South Dakota
(Fig. 3). Consequently, most of the new heat-flow locations were selected to provide a profile
from the recharge to the discharge zone. Also of interest was the possibility of heat advection
by localized vertical discharge from the uppermost aquifer to the surface through fractures in
the Pierre Shale (Bredehoeft, Neuzil, and Milley, 1983; Neuzil, Bredehoeft, and Wolff, 1984).
To investigate this question, three of the new heat-flow holes were drilled in an area west of
the confluence ofthe White and Missouri rivers in south-central South Dakota (Fig. 2) where
high geothermal gradients coincide with a region that Neuzil, Bredehoeft, and Wolff (1984)
show as having vertical fracture-leakage velocities of the order of approximately 10 -10 m S-I.

Resolution of Temperature Gradients and Thermal Conductivitif~s

Heat flow is the product of temperature gradient and thermal conductivity, that is

Q=dT/dz)"

and the errors associated with temperature measurements, temperature gradient calculations,
and measurement ofthermal conductivity determine the error of a heat-flow measurement.
102 GOSNOLD

NORTH DAKOTA

---I - ~-
104 103 102 101 100 99 98 97

Figure 2. Locations of heat-flow sites. Sites previously published are shown by dots (Sass and others, 1971; Sass
and Galanis, 1983; Gosnold,1990) and new sites are shown by triangles. TIrree heat-flow holes at White River
site lie within ellipse. Line A - A' locates cross section for Figures 3, 8, and 9.

2.0

km

1.0

o
o 100 200 300
km

Figure 3. Four regional aquifers, Al - A4, and three confming units, QI - Q3 are shown in schematic cross
section along line A - A' in Figure 2. Al - Fall River, Lakota and Dakota Sandstones (Cretaceous); A2 -
Pahasapa (Madison) Limestone (Mississippian); A3 - Minnelusa Formation (Pennsylvanian); A4 - Deadwood
Sandstone and Red River Formation (Cambrian - Ordovician). QI - Cretaceous shales; Q2 - shales of Permian
to Jurassic age; Q3 - crystalline basement.

Temperatures were measured with a thermistor probe having a precision of ±O.OOl cC and an
absolute accuracy of ±O.02 cC. Depths were determined with a precision of ±O.OOl m and
temperature gradients were calculated as least-squares fits to the selected depth interval for
each borehole. The temperature measurements were made several years after drilling so that
the borehole temperatures were at equilibrium conditions.
BASIN-SCALE GROUNDWATER FLOW AND ADVECTIVE HEAT FLOW 103

Thermal Conductivity

Thermal conductivity determinations for the Pierre Shale were problematic and a
reliable technique for measuring thermal conductivity on shales was not available until the
final borehole in this study was drilled at Wall, South Dakota in 1995. Consequently, thermal
conductivities were estimated based on published values for nearby sites and published
descriptions of the lithology of the Pierre Shale. The thermal conductivity of shale is one of
the lowest values for any rock type, for example, granite typically has a conductivity of 2.5 to
3.5 W mol °C- l whereas the most recently published values for the conductivity the Pierre
Shale range from 1.0 - 1.2 W mol °C- l (Blackwell and Steele, 1989; Sass and Galanis, 1983;
Gosnold and Todhunter, 1994). Using a half-space needle-probe technique, we measured a
thermal conductivity of 1.2 ± 0.2 W mol °C- l (n = 60) on Pierre Shale cores from a heat-flow
site in southwestern Manitoba and a thermal conductivity of2.0 ± 0.25 W mol °C-l (n = 60) at
a site in western South Dakota. Clay fraction analysis of the Pierre Shale samples showed that
the difference is the result of a higher quartz content in the samples from the Wall, South
Dakota site. The Pierre Shale is divided into eastern, median, and western facies on the basis
oflithology (Tourtelot, 1962; Gill and Cobban, 1965). All heat-flow sites in this study except
the Wall site lie within the clay-rich eastern facies that is characterized by low-thermal
conductivity. Therefore, the thermal conductivity of the Pierre Shale was assumed to be 1.2
W mol °C- I ±0.2 for all localities but the Wall heat-flow site in South Dakota.

T-z Profiles

Climatic warming since the end of the Little Ice Age, c.a. 1850, has influenced
temperatures in the upper 80 meters of all of the boreholes (Gosnold, Todhunter, and Schmidt,
1997). However, the temperature gradients below 80 meters are remarkably linear in all but
three boreholes (Fig. 4) and the standard error for least-square gradient calculations averages
only 0.9% (Table 1). This observation is significant because linear temperature gradients
below the climatic disturbance in the Pierre Shale suggest that only conductive heat flow
occurs within the formation. The Pierre Shale has a low fluid permeability (Bredehoeft
Neuzil, and Milley, 1983), thus it could mask the thermal signals from any advective system
beneath it. The thickness of the Pierre Shale plus the underlying Niobrara Formation, Carlile

60
50
40
I:-
~
150·c/km

lOO·C/km
50·c/km
u 20
30

25

~
U
Ol
g'30 CD
0 015
20
10
10
0 5
0 200 400 600 800 0 50 100 150 200 250
Depth (m) Depth (m)

Figure 4. Temperature-depth profiles for new heat-flow holes in South Dakota. A, shows T-z profiles for
specially completed heat-flow holes; and B, shows T-z profiles from 5 holes of opportunity. Profiles are spaced
on temperature scale for better visibility_Profiles with lower temperature gradients are from sites in west of study
area and those with higher temperature gradients are from sites in east of study area.
104 GOSNOLD

Table I. New heat-flow data. Temperature gradients are linear least-squares fits to indicated
depth intervals. Thennal conductivities for all sites but Wall-I and Wall-2 are estimates based
on measured thennal conductivites of Pierre Shale from drill cores taken at Hayes, S. D. (Sass
and Galanis, 1983) and at Wawanesa, Manitoba (Gosnold and Todhunter, 1994).
Conductivities at Wall site were measured using half-space needle probe technique (Gosnold
and Todhunter, 1994, Gosnold, Todhunter, and Schmidt, 1997).

LOCATION LAT. LONG GRADIENT DEPTH THERMAL HEAT ELEV


MkM" INTERVAL CONDo FLOW m
m Wm"K' mWm·2
WhiteR. 1 43.83 99.57 54.2±0.2 99-181 1.2±O.2 65.0±O.4 545
WhiteR. 2 43.82 99.72 90.9±O.2 91-179 1.2±0.2 109.1±0.4 540
White R. 3 43.79 99.74 95.1±0.1 95-175 1.2±0.2 114.1±0.3 545
WhiteR. 4 43.74 100.04 108.6±0.1 107-179 1.2+0.2 130.3±0.3 570
WhiteR. 5 43.70 100.04 85.6±0.5 95-175 1.2±O.2 102.7±0.3 480
WhiteR. 6 43.68 100.04 91.1±0.1 95-175 1.2±0.2 109.3±0.3 540
Goodman 43.45 99.18 50.7±0.2 115-260 1.2±0.2 60.8±0.4 547
Dixon 44.90 101.80 76.6±0.3 145-315 1.2±0.2 91.9±0.5 590
Belvidere 43.85 101.26 65.7±0.1 81-181 1.2±0.2 78.7±0.3 707
Kadoka 43.81 101.50 50.1±0.2 81-181 1.2±0.2 60.1±0.2 747
Wall-I 43.74 102.22 32.1±0.01 100-181 2.0±0.3 63.8±0.3 848
Wall-2 43.74 102.23 33.7±0.01 150-202 2.0±0.3 67.4±0.3 840
Koch 34-17 45.84 102.93 48.1±O.1 300-470 1.2±0.2 57.7±0.3 935
Koch 14-15 45.87 102.96 54.0±0.1 300-470 1.2±0.2 64.9±0.3 929
Sheep Draw 43.19 103.90 53.6±0.1 300-470 1.2±0.2 64.3±0.3 421
Sisseton 45.71 96.92 42.9±0.2 200-275 1.2±O.2 51.5±0.4 387

Shale, and Belle Fourche Shale ranges from 400 m to 750 m across the study area and only one
heat-flow hole has been logged below these Cretaceous age shales. That site, near Burton,
Nebraska (Gosnold, 1990), shows a 20% decrease in heat flow below the Pierre Shale which
suggests a possible advective heat-flow component resulting from fluid flow in the underlying
formations.

A Special Case

Three of the T-z profiles in the high heat-flow region along the White River in central
South Dakota show a concave upward curvature from the ground surface to the top of the
Niobrara and a convex curvature within and below the Niobrara formation (Fig. 5). Part of
the concave curvature is the result of climatic warming, but the curvature increases rather than
decreases with depth which suggests a thermal effect in addition to the climate signal.
Assuming a homogeneous thermal conductivity for the overlying Pierre Shale and the
underlying Carlile Shale, these T-z profiles suggest that heat flow systematically increases
toward the Niobrara from above and from below. Neuzil (1993, 1995) demonstrated the
existence of abnormal fluid pressures in the Cretaceous shales in South Dakota and has
suggested (C.E. Neuzil, pers. comm., 1997) that groundwater should be flowing by fracture
leakage into the Niobrara from the Pierre Shale and Carlile Shale. Advective heat flow
calculations using the one-dimensional relationship of Stallman (1963), that is
Qe VDaC p
In - = --.!...

Qe A
where QI is heat flow at the base of the water-flow zone, ~ is heat flow at the top of the zone,
V is Darcy velocity in m S-I, D is the length of the zone in meters, a is density of the fluid in
kg m-3, Cp is heat capacity of the fluid in W s, and kgl °CI,.A. is thennal conductivity in W m 1
BASIN-SCALE GROUNDWATER FLOW AND ADVECTIVE HEAT FLOW 105

26 140
24
120
22
100
() 20
!6'18
-l:1
~ 80
0 16 "0
60
14
12 40
20 L-______________________
10
0 50 100 150 200 110 120 130 140 150 160 170 180 190
Depth (m) Depth (m)

Figure 5. T-z and dT/dz profiles White River heat-flow sites 4, 5, and 6. dT/dz plots were made with Niobrara
Formation as datum. Increase in temperature gradient above the Niobrara is interpreted to be result of
combination of ground-surface warming and heat advection by fracture leakage from surface into Niobrara.
Decrease in temperature gradient below Niobrara is interpreted to be result of heat advection by upward fracture
leakage.

zone in meters, 0 is density of the fluid in kg m-3, Cp is heat capacity of the fluid in W s, and
Kg-I °el , Ais thermal conductivity in W m-I °e l , suggest that downward groundwater flow
in the lower 30 meters of Pierre Shale at 5 X 10-9 m S-I and upward groundwater flow in the
upper 30 meters of Carlile Shale at 5.410-9 m 8"1 could cause the inferred increase in heat flow
toward the Niobrara formation.

Estimated Heat-Flow Values

With the determination of several key parameters for both the heat flow and the
ground-water flow systems, quantitative modeling of the coupled systems should reveal the
degree to which the systems interact. Ideally, the models should include both lateral transport,
represented by two-dimensional cross sections, and areal heat flux models. The data acquired
for the east-west heat-flow profile provide a reasonable basis for two-dimensional modeling
of a cross section of the coupled systems. However, these data do not cover a sufficient
surface area to provide an estimate of the areal heat flux.
The lack of areal coverage by the conventional heat-flow data was overcome by
including heat-flow values estimated from Schoon and McGregor's (1974) temperature
gradient data. Many of the temperature gradient data were determined from the differences
between the surface and bottom-hole or inferred bottom-hole temperatures of wells that
penetrate the top of the Dakota Sandstone. Normally two-point gradients do not adequately
sample subsurface temperatures and are not useful for calculating heat-flow. In this situation,
however, the temperature gradient between the ground surface and the top of the Dakota
Sandstone is the actual temperature gradient in the thick section of Cretaceous shales that
blanket the region. The predominant lithologies between the two points are marine shales,
mostly the Pierre Shale (Table 2). Temperature gradients measured in heat-flow holes in the
Pierre Shale and the underlying shales are typically linear, which suggests that these shales
have a relatively uniform thermal conductivity. Geologically, this is a reasonable inference
because the members of the Pierre Shale tend to be laterally continuous and lithologically
uniform (Tourtelot, 1962; Schultz, 1964, 1965). Therefore, assuming that the effective thermal
106 GOSNOLD

Table 2. Generalized stratigraphy of southern South Dakota. Units included in major regional
aquifers are designated AI, A2, A3, and A4.

SYSTEM FORMATION LITHOLOGY MAX. THICKNESS


(meters)

PIERRE SHALE SHALE 610


NIOBRARA FORMATION CHALK 70
LOWER CARLILE FORMATION SHALE 220
CRETACEOUS GREENHORN FORMATION IMPURE LIMESTONE 110
BELLE FOURCHE SHALE SHALE 170
MOWRY SHALE SHALE 77
NEWCASTLE SANDSTONE SANDSTONE 20
SKULL CREEK SHALE SHALE 83
Ai INYAN KARA GROUP SANDSTONE 280

JURASSIC MORRISON FORMATION SHALE 67


UNKPAPA SANDSTONE SANDSTONE 69
SUNDANCE FORMATION SHALEISANDSTONE 138

TRIASSIC SPEARFISH FORMATION SANDY SHALE 210

MINNEKAHTA LIMESTONE LIMESTONE 16


PERMIAN OPECHE FORMATION SHALE/SANDSTONE 41

PENNSYLVANIAN A2 MINNELUSA FORMATION SANDSTONEILIMESTON 270

MISSISSIPPIAN A3 PASAHAPA LIMESTONE LIMESTONE 192

A4 ENGLEWOOD LIMESTONE LIMESTONE 20


DEVONIAN A4 RED RIVER FORMATION DOLOMITEILIMESTONE 20

ORDOVICIAN WINNIPEG FORMATION SHALE 30

CAMBRIAN A4 DEADWOOD FORMATION SANDSTONE 122

PRE-CAMBRIAN METAMORPHIC and


IGNEOUS ROCKS

conductivity of the Pierre Shale is approximately 1.2 W m· l °e- l heat-flow values were
estimated for 62 sites (Table 3) selected from Schoon and McGregor's (1974) compilation.
It is important to distinguish between these heat-flow estimates and conventional heat-
flow values. A conventional heat-flow value requires temperature gradients measured in a
borehole and mUltiple thermal conductivity values measured on samples from the measured
gradient interval or from a nearby site having equivalent lithology. Statistical analysis ofthese
values provides a measure of the accuracy and quality of the heat-flow determination.
However, no statistical measure of accuracy is possible for the estimated heat-flow values.
The accuracy of these estimated heat-flow values was tested by comparing the
temperature gradients measured at 12 conventional heat-flow holes in the thermal anomaly to
the nearest two-point gradient contours in Figure 1. The results (Table 4) show that four of
the measured temperature gradients lie within the contour values, one is 6% low, two are
approximately 10% high, and the remainder are high by 14, 20, 29, 59, and 101 percent
compared to the nearest gradient contour line. Ifthis comparison is representative of the total
popUlation of estimated heat-flow values, it indicates that approximately 2/3 of the estimated
heat-flow values are accurate within ± 10% and 113 underestimate heat flow by at least 20%.
BASIN-SCALE GROUNDWATER FLOW AND ADVECTIVE HEAT FLOW 107

Table 3. Estimated heat-flow values. Type designations are WW (water well) and BHT
(bottom-hole temperature in oil well).
Latitude Longitude Gradient Type Heat flow Latitude Longitude Gradient Type Heat Flow
43.00 98.60 89.2 VWI/ 107 43.72 99.50 93.3 VWI/ 112
43.02 98.48 69.2 VWI/ 83 43.73 99.69 93.3 VWI/ 112
43.04 98.61 110.8 VWI/ 133 43.79 99.70 93.3 VWI/ 112
43.05 98.25 105.0 VWI/ 126 43.80 99.51 93.3 VWI/ 112
43.10 98.67 67.5 VWI/ 81 43.84 100.28 70.0 VWI/ 84
43.11 98.61 70.8 VWI/ 85 43.86 100.10 83.3 VWI/ 100
43.16 98.38 42.5 VWI/ 51 43.93 102.69 29.9 BHT 36
43.16 98.87 89.2 VWI/ 107 43.98 99.13 41.7 VWI/ 50
43.17 98.82 74.2 VWI/ 89 44.06 99.48 50.8 VWI/ 61
43.17 99.07 107.5 VWI/ 129 44.08 99.53 49.2 VWI/ 59
43.18 98.96 82.5 VWI/ 99 44.11 99.66 42.5 VWI/ 51
43.22 98.95 100.8 VWI/ 121 44.14 101.79 16.4 BHT 20
43.31 99.08 97.5 VWI/ 117 44.17 99.46 47.5 VWI/ 57
43.31 99.13 99.2 VWI/ 119 44.25 100.62 75.8 VWI/ 91
43.38 99.37 116.7 VWI/ 140 44.31 101.52 26.1 BHT 31
43.39 99.15 79.2 VWI/ 95 44.32 102.77 22.4 BHT 27
43.41 99.13 92.5 VWI/ 111 44.34 102.02 18.2 VWI/ 22
43.44 99.25 115.8 VWI/ 139 44.35 100.39 68.3 VWI/ 82
43.44 99.30 115.8 VWI/ 139 44.38 101.58 27.6 VWI/ 33
43.45 99.11 95.8 VWI/ 115 44.46 101.76 21.6 BHT 26
43.46 99.28 102.5 VWI/ 123 44.49 101.59 13.8 BHT 17
43.52 98.97 69.2 VWI/ 83 44.49 102.04 22.3 VWI/ 27
43.54 99.57 97.5 VWI/ 117 44.51 102.01 25.1 VWI/ 30
43.55 99.84 104.2 VWI/ 125 44.52 102.54 18.4 BHT 22
43.58 99.85 100.8 VWI/ 121 44.54 101.59 19.5 BHT 23
43.59 99.39 82.5 VWI/ 99 44.63 102.56 22.2 BHT 27
43.62 99.72 104.2 VWI/ 125 44.75 102.53 32.5 BHT 39
43.67 99.40 70.8 VWI/ 85 44.75 103.16 21.9 BHT 26
43.67 99.44 70.8 VWI/ 85 44.77 102.96 21.8 BHT 26
43.67 99.75 92.5 VWI/ 111 44.80 102.88 14.0 BHT 17
43.71 102.88 29.8 BHT 36 44.82 102.01 15.9 BHT 19

Table 4. Temperature gradients (mK m· l ) measured in heat-flow holes (rows 2-5) compared
to nearest temperature gradient contour lines in Figure 1 (row 1).

Contour range Contour range Contour range Contour range


36 - 54 54 -72 72 - 90 90 -108
44 82 68 90
59 93 90 96
60 108
86
109

The low estimates are not surprising considering that many of Schoon and McGregor's (1974)
geothermal gradients were determined using the discharge temperature from a flowing well
as the bottom-hole temperature. Although several factors may be involved, such as, well
108 GOSNOLD

depth, well age, type of casing, hole diameter, and discharge rate, heat loss as the water flows
up the well results in a low estimate of the bottom-hole temperature. However, the important
point is that none of the estimated heat flows seem to overestimate heat-flow and incorrectly
extend the high heat flow region.

HEAT-FLOW RESULTS AND A HEAT-FLOW


CONTOUR MAP OF THE NORTHERN PLAINS

Eight of the new heat-flow data lie within the anomalous heat-flow region and average
102 ± 20.2 m W m·2 • The eight new data outside the anomalous region average 62.0 ± 4.8 m W
m· 2 • Heat flow at 94 sites in the northern Great Plains, including the 16 new sites reported
here, and the 62 estimated heat-flow values were combined to produce a contour heat-flow
map (Fig. 6). If only heat-flow points lying between the 30 mWm·2 and 70 mW m·2 contours
are included, the average of the conventional heat-flow values is 58 ± 9 mW m·2 • The heat-
flow pattern that appears east of the Black Hills shows the expected variation for advective
heat transport across a tilted basin. Heat flow is low in the recharge area, systematically

Figure 6. Heat-flow contour map of northern Great Plains region. Heat-flow values are in mW m·2•
Conventional heat-flow sites are shown as filled squares and asterisks for previously published data and as filled
circles for data reported in this study. Estimated heat-flow sites are shown as open triangles. Region between
bold lines in central South Dakota is general area where Paleozoic aquifers discharge into Dakota Sandstone
(Cretaceous) through subcrop contacts.
BASIN-SCALE GROUNDWATER FLOW AND ADVECTIVE HEAT FLOW 109

increases to high values in the discharge area and falls to the regional value beyond the
discharge area.

THE AQUIFER SYSTEM

Investigation of the relationship between the heat-flow anomaly and groundwater flow
requires understanding the system of confined aquifers underlying the area. Groundwater flow
in the aquifer system is gravity driven from the recharge area at 1200 m elevation to the
discharge area at about 400 m elevation (Fig. 7 A-7E). Downey (1986) divided the system into
four major confined aquifers that are recharged where eastward-flowing streams cross their
outcrops along the eastern margin of the Black Hills (Swenson, 1968). The three lower

Cambrian..Qrdovician Aquifer Pennsylvanian Aquifer

104 103 102 101 100 99 98 97 104 103 102 101 100 99 98 97
LongHude Longitude

Mississippian Aquifer Cretaceous Aquifer

104 103 102 101 100 99 98 97 104 103 102 101 100 99 98 97
LongHude LongHude

Figure 7. Potentiometric surface contours on tops of four major aquifer systems showing west to east flow
pattern modified from Downey (1986). Contours are in meters and datum is sea level. East-west dashed line
designates heat-flow profile shown in Figure 8.

aquifers discharge at subcrop contacts with the uppermost aquifer beginning about 200 km east
of the Black Hills and extending eastward for another 100 km (Swenson, 1968; Downey,
1986). Waters from the uppermost aquifer discharge primarily at outcrops to the east
110 GOSNOLD

(Downey, 1986) and secondarily through vertical fracture leakage (Neuzil, Bredehoeft, and
Wolff, 1984; Downey, 1986). A detailed description ofthe aquifers is beyond the scope of this
study, but the following general description of the system provides information necessary to
construct analytical and numerical models of the thermal effects of water flow in the system.
The lowermost aquifer is composed of limestones and dolomites of the Red River
Formation (Ordovician) and includes sandstones of the Deadwood and Winnipeg Formations
(Cambrian - Ordovician). Calculated Darcy velocities (Darcy velocity) in this aquifer range
from 0.6 m y.1 to 22.8 m y.1 (Downey, 1986). In the area of the heat-flow anomaly, the
Cambrian-Ordovician aquifer is confined below by the crystalline basement. The
Mississippian aquifer, the Pahasapa Limestone, directly overlies the Cambrian-Ordovician
aquifer in the Black Hills and in southern South Dakota. Computed Darcy velocities for the
Mississippian aquifer range from 0.6 my-I to 6 m yl (Downey, 1986). The third major
aquifer, the Pennsylvanian aquifer, overlies the Pahasapa Limestone and consists of sandstones
and limestones of the Minnelusa Formation. The Minnelusa, with an average thickness of
about 300 meters, is the thickest single aquifer in the area. It is confined above by shales and
siltstones of Permian, Triassic, and Jurassic age. Groundwater flow velocities for the
Minnelusa were not determined in any prior studies, but Downey (1986) determined that the
transmissivity of the Minnelusa Formation is similar to that of the Pahasapa Limestone, that
is 2.7 - 8.1 x 10-4 m2 S-I. Therefore, this aquifer is assumed to permit water flow at velocities
equal to those of the Pahasapa. The Lower Cretaceous aquifer consists of the Fall River,
Lakota, and Dakota Sandstones which have a total porosity of 26% (Rahn, 1985).
Transmissivities range from 2.2 x 10-3 to 1.6 x 10-2 m2 8"1 which are 1 to 2 orders of magnitude
greater than those in the Paleozoic units. The hydraulic gradient of the Lower Cretaceous
aquifer essentially equals that of the Paleozoic rocks, thus, for a first approximation, Darcy
velocities in the Lower Cretaceous aquifer are assumed to be at least an order of magnitude
greater than the Darcy velocities in the Pahasapa Limestone. The Lower Cretaceous aquifer
is confined above by 400 - 750 meters of shale and minor amounts oflimestone.
The three lower aquifers, the Cambrian-Ordovician aquifer, the Mississippian aquifer,
and the Pennsylvanian aquifer are essentially a single system not separated by confining layers
in the area of anomalous heat flow. This lower system has a maximum combined thickness
of about 650 meters in the recharge area and thins to the east where it is truncated and forms
an angular unconformity with the overlying Lower Cretaceous aquifer.
The subcrop contact between the lower aquifer system and the Lower Cretaceous
aquifer is a zone of vertical leakage where waters are discharged into the Dakota Sandstone
(Swenson, 1968; Downey, 1986). Evidence of this vertical leakage is contained in chemical
analyses of waters in the Dakota which indicate that the aquifer is extensively flushed with
meteoric water in the recharge area but contains calcium-sulfate waters derived from the
Paleozoic formations in the discharge area (Swenson, 1968; Schoon, 1971). The region where
the subcrops of the three lower aquifers contact the Dakota lies between the bold lines in
Figure 6. Thus, the area of highest heat flow generally coincides with the zone where waters
from the lower aquifers discharge into the Dakota.

TEST OF THE HYPOTHESIS BY MODELING

For a first approximation, the relation between the heat-flow anomaly and the ground-
water-flow system can be tested with simple analytical models such as the basin model of
Domenico and Palciauskus (1973) and the wedge model of Jessop (1991). The average heat-
BASIN-SCALE GROUNDWATER FLOW AND ADVECTIVE HEAT FLOW III

flow value for the thermal anomaly is about twice the background heat flow, 120 m W m,2 vs.
58 m W m,2, and both analytical models indicate that fluid flow across the basin structure at a
Darcy velocity of I m y'l would cause a doubling of surface heat flow in the discharge area.
However, these simple analytical models do not address the dynamics of the groundwater
system in South Dakota. The system of four aquifers is asymmetrical, in thickness, computed
Darcy velocities differ between aquifers and within the aquifers (Downey, 1986).
To address the complexities of the system, I applied a two-dimensional finite-
difference model which couples heat flow and groundwater flow using the method of Brott,
Blackwell, and Mitchell (1978) and Brott, Blackwell, and Ziagos (1981). That is, groundwater
velocity is a parameter of the model rather than a quantity calculated from the hydrologic and
thermal properties of the aquifers. This method applies appropriately in this situation because
gravity rather than heat is the driving force in the aquifer and velocities have been calculated
independently (Downey, 1986).
The cross section used for the simulation lies along the east-west dashed lines in Figure
7. The finite-difference grid contains 53 vertical and 103 horizontal nodes with 50-m vertical
and 3000-m horizontal spacings. The parameters specified for each node are: thermal
conductivity, radioactive heat production (basement rocks only), fluid heat capacity, rock heat
capacity, Darcy velocity, and flow direction. Thermal conductivities of the distinctive
lithologic units were taken from Gosnold (1991) and were assigned in the model as follows:
The two shale units, QI & Q2 were assigned a thermal conductivity of 1.2 W m'l °C'l. The
crystalline basement was assigned a thermal conductivity of 3,0 W m'l °C'l. The Lower
Cretaceous aquifer, AI, was assigned a conductivity of2.7 W m'l °C I , and the three carbonate
aquifers, A2, A3, A4, were assigned a thermal conductivity of 2.5 W m'l °C,l.
The initial thermal structure of the model was a purely conductive thermal regime with
heat flow in the Black Hills at 90 mW m,2 and heat flow elsewhere at 58 mW m'2. The ground-
surface temperature was fixed at 10°C, and prior to running the advection models, the
conductive model was run for a sufficient time, 300,000 years to achieve equilibrium allowing
for refraction along the margin of the Black Hills Uplift.
The modeling process involved testing different combinations of groundwater
velocities to obtain the best agreement between the modeled and observed heat-flow profiles.
The wedge model predicts that a constant updip velocity of 6.34 x 10,8 m S'l in a single aquifer
400 m thick would generate a heat flow of about 140 m W m,2 at the discharge end of the basin.
Thus, to compare the wedge model to the numerical model, one model run was made using a
constant velocity of 6.34 x 10,8 m S'l in AI, A2, A3, and A4. The resultant heat-flow
maximum for the numerical model was 168 mW m'2. Running the model with no water flow
in Al produced a maximum of 140 mW m,2 which agrees with the wedge model. This
exercise demonstrates agreement between the wedge and numerical models for similar
geometries, but, significantly, it shows the importance of inclusion of the Dakota Sandstone
in the model. Subsequent models were based on the variable flow velocities given by Downey
(1986), that is velocities decreased from the recharge side to the discharge side of the models.
Excellent agreement between observed heat-flow values and the model (Fig. 8) was obtained
with a recharge velocity of6.34 x lO'8 m S'l for A2, A3, and A4 with a decrease to 4.76 x 10,8
m S'l toward the discharge side of the model. The velocity of Al used in the best-fitting model
was 1.902 x 10'7 m S'l although the contribution from Al is relatively small.
Isotherms along the model cross section for the best-fitting model and the initial
(conductive) model are shown in Figure 9. Variation in the geothermal gradient along the
section is shown in Figure 10 which includes two model profiles along the cross section and
two observed temperature-depth profiles. The T-z profile at Wall, South Dakota is
112 GOSNOLD

West East
140
120
N
E 100
3:
.s:;: 80 ...
0
0;;::
60 ... ...
iii 40
OJ
J:
20
0
0 100 200 300 400 500 600
Distance (km)

Figure 8. Model heat-flow profile (solid line) superimposed on observed heat flow and estimated heat-flow
values (triangles).

superimposed on the model curve; both have the lowest temperature gradients. The model
profile in the high heat-flow zone (thin line with dark square symbols) closely matches the
observed T -z profile from Burton, Nebraska (thick line without symbols). The T -z profile at
the Burton site exhibits a drop in heat flow from 106 m W mo2 to 80 m W m o2 at the contact
between the shale and the Dakota. The high heat flow, 80 mW mo2 , in the lower section of the
Burton hole is assumed to be the result of water flow in A3 and A4, both of which underlie the
area.

Initial conditions.

Figure 9. Isotherms (deg C) for initial and advective conditions for model scheme shown in Figure 6. Aquifers
A2-A4 are shown as single unit in this figure for better visibility of isotherms.
BASIN-SCALE GROUNDWATER FLOW AND ADVECTIVE HEAT FLOW 113

80 )!I4-i../U

. ..
70
60
() 50
g>40
0 30

20
10
0
0 200 400 600 800 1000
Depth (m)

Figure 10. Observed T-z profiles agree with T-z profiles generated by numerical model of advective heat
transport in recharge and discharge sections of aquifer system. Thick lines with closed circles and no symbols
are observed T-z profIles at Wall, SD and Burton, NE respectively. Thin lines with triangles and closed squares
are model profiles for those sites.

HEAT-FLUX ESTIMATES

Agreement between modeled and observed heat-flow profiles suggests that


groundwater flow at velocities of6.34 x 1O.sm S·l in the recharge area and 4.76 x 1O.s m S·l in
the discharge area for A2, A3, and A4 can account for both the low and high heat-flow
anomalies east of the Black Hills. In this situation, the positive heat flux in the discharge area
should equal the negative heat flux resulting from downward water movement in the recharge
area. One way of testing this is to compare positive and negative heat fluxes determined using
the heat-flow contours in Figure 6. The heat-source region is defined as the area having heat-
flow values above 58 mW m·2 • The heat sink region includes the Black Hills Uplift where two
heat-flow values of 82 m W m-2 have been recorded (Sass and others, 1971). The high heat
flow in the Black Hills region is partly a residual effect from the uplift of the Black Hills in the
Laramide Orogeny and partly from the thermal effects of mid-to-late Tertiary magmatism.
The background heat-flow value for the area near the Black Hills that provides a balanced heat
flux for the heat source and heat sink is 91 m W m-2 • The heat flux estimates for 10 m W m-2
intervals in the source and sink areas are given in Table 5.

CONCLUSIONS

Conventional and estimated heat-flow data delineate adjacent regions of anomalously


high and anomalously low heat flow between the Black Hills and eastern South Dakota (Fig.
5). A gravity-driven regional groundwater flow system in confined aquifers that underlie the
region is characterized by downward groundwater flow in the region of low heat flow and by
upward groundwater flow in the region of high heat flow (Fig. 2). Analysis of the overall
hydrothermal regime by simple analytical models and a more complex numerical model
suggests that the groundwater flow system causes the areas of anomalous heat flow.
Groundwater flow velocities sufficient to generate the anomalous heat flow are 3.17 x 1O-s m
sol for the analytical models and 6.34 x 10-8 m Sol in the recharge area and 4.76 x 1O-s m sol in
114 GOSNOLD

Table 5. Heat-flux calculations for discharge and recharge areas. Heat flow anomaly (HFA)
used to calculate heat flux is difference between background heat flow, 58 m W m·2 and mean
value of heat flow in each area for discharge area. HFA in recharge area was based on
background heat flow of 58 mW m·2 in plains region having contoured heat-flow value of 43
mW m-2 and background of91 mW m-2 in Black Hills region.
Discharge Area Recharge Area

Heat Flow Area (km 2) HFA x area Heat Flow Area (km 2) HFAx area
mWm-2 (gigawatts) (mWm-2 ) (gigawatts)
73 24674 0_37 43 22309 -0.45
83 13684 0_34 33 20785 -1.14
93 12932 0.45 23 15492 -1.01
103 10104 0.45
113 8590 0.97
123 4678 0_30
133 3154 0_24
TOTAL 2_63 -2.64
FLUX

the discharge area for the numerical model. These flow velocities lie with the range of
velocities calculated from hydrologic parameters by Downey (1986), but the heat-flow results
suggest that the lower end of the Downey's velocity values yield better agreement. Heat-flux
calculations based on the areas defined as high and low heat-flow balance at 2.63 GW for the
discharge area and -2.64 GW for the recharge area. The background surface heat flow of 58
mW m-2 used in the analytical and numerical models was based on the average of all heat flow
data outside the anomalous region. Because this background heat-flow value yielded model
heat-flow results consistent with observations, it is assumed to characterize surface heat-flow
in the region should the groundwater flow system not exist. The heat-flow value of 66 m W
m- 2 for the Great Plains given by Morgan and Gosnold (1989) included some of the
anomalously high heat-flow values from South Dakota and Nebraska. Important remaining
questions concern the length of time that this hydrothermal regime has existed and the
implications for temperatures in the lower crust.

ACKNOWLEDGMENTS

The author gratefully acknowledges reviews by two anonymous reviewers and


constructive comments on an earlier version of this paper by Jeffery Nunn and John Sass. The
heat-flow data used in this research was acquired with support by Department of Energy
contracts DE FG07-85IDI2606, DE FG07-88ID12736, C85-11768, and DE FC03-
90ER61010. Financial support does not constitute an endorsement by DOE of the views
expressed in this article.
BASIN-SCALE GROUNDWATER FLOW AND ADVECTIVE HEAT FLOW 115

REFERENCES

Adolphson, D. G., and F. LeRoux, E. F., 1968, Temperature variations of deep flowing wells in South Dakota:
U.S. Geo!. Survey Prof. Paper 600-D, p. 60-62.
Bachu, S, and Hitchon, B., 1996, Regional-scale flow of formation waters in the Williston Basin: Am. Assoc.
Petroleum Geologists Bul!. 80, no. 2, p. 248-264.
Blackwell, D. D., 1971, The thermal structure of the continental crust, in Heacock, J. G., ed., The Structure and
Physical Properties of the Earth's Crust: Am. Geophys. Union, Geophys. Mon. Ser., v. 14, p. 169-184.
Blackwell, D.D., and Steele, J.A., 1989, Heat flow and geothermal potential of Kansas: Kansas Geo!. Survey
Bul!. 226, p. 267-295.
Bredehoeft, J. D., Neuzil, C. E., and Milley, P. C. D., 1983, Regional flow in the Dakota aquifer: a study of the
role of confming layers: U. S. Geo!. Survey Water Supply Paper 2237,45 p.
Brott, C. A., Blackwell, D. D., and Mitchell, J., 1978, Tectonic implications of heat flow in the western Snake
River Plain, Idaho: Geo!. Soc. America Bul!., v. 89, no. 12, p. 1697-1707.
Brott, C. A., Blackwell, D. D., and Ziagos, J. P., 1981, Thermal and tectonic implications of heat flow in the
eastern Snake River Plain, Idaho: Jour. Geophys. Res., v. B86, no. 12, p. 11,709 - 11,734.
Combs, J., and Simmons, G., 1973, Terrestrial heat flow determinations in the north central United States: Jour.
Geophys. Res., v. 78, no. 2, p. 441-461.
Deming, D., Sass, J. H., Lachenbruch, A. H., and DeRito, R. F., 1992, Heat flow and subsurface temperature as
evidence for basin-scale ground-water flow, North Slope of Alaska: Geo!. Soc. America Bull., v. 104,
no. 5, p. 528-542.
Domenico A., and Pa1ciauskas, S., 1973, Theoretical analysis of forced convective heat transfer in regional
ground-water flow: Geo!. Soc. America Bull., v. 84, no. 12, p. 3803-3814.
Donaldson, 1. G., 1962, Temperature gradients in the upper layers of the earth's crust due to convective water
flows: Jour. Geophys. Res., v. 67, no. 9, p. 3449-3459.
Downey, J.S., 1986, Geohydrology of bedrock aquifers in the northern Great Plains in parts of Montana, North
Dakota, South Dakota, and Wyoming: U.S. Geo!. Survey Prof. Paper 1402-E, 87 p.
Gill, G. R., and Cobban, W.A., 1965, Stratigraphy of the Pierre Shale, Valley City and Pembina Mountain areas,
North Dakota: U. S. Geo!. Survey Prof. Paper, 392-A, p. 1-20.
Gosnold, W.D., Jr., 1985, Heat flow and groundwater flow in the Great Plains of the United States: Jour.
Geodynamics, v. 4, no. 1-4, p. 247-264.
Gosnold, W.D., Jr., 1990, Heat Flow in the Great Plains of the United States: Jour. Geophys. Res., v. B95, no.
1, p. 353-374.
Gosnold, W.D., Jr., 1991, Subsurface temperatures in the northern Great Plains, in Slemmons, D.B., Engdahl,
E.R., Zoback, M.D., and Blackwell, D.D., eds, Neotectonics of North America: Geo!. Soc. America,
Decade Map Volume 1, p. 467-472.
Gosnold, W. D., Jr., and Todhunter, P. E., 1994, Analysis of the geothermal gradient as a method ofpaleoc1imate
reconstruction (abst.): EOS, v. 75, no. 44, Supp!., p. 96.
Gosnold, W. D., Jr., Todhunter, P. E., and Schmidt, W., 1997, The borehole temperature record of climate
warming in the mid-continent of North America: Global and Planetary Change, v. 15, no. 1-2, p. 33-45.
Head, W. J., Kilty, K. T., and Knottek, R. K., 1978, Maps showing temperatures and configurations of the tops
of the Minnelusa Formation and the Madison Limestone, Powder River Basin, Wyoming, Montana, and
adjacent areas: U. S. Geo!. Survey Open-File Rept. 78-905, 2 sheets.
Hildenbrand, T. G., and Kucks, R. P., 1985, Model of the geothermal system in southwestern South Dakota from
gravity and aeromagnetic studies, in Hinze, W. J., ed., The Utility of Regional Gravity and Magnetic
Anomaly Maps: Soc. Exp!. Geophys., p. 258-266.
Jessop, A. M., 1991, Hydrological distortion of heat flow in sedimentary basins: Tectonophysics, v. 164, no. 2-4,
p.211-218.
Lachenbruch, A.H., and Sass, J. H., 1977, Heat flow in the United States and the thermal regime of the crust, in
Heacock, J. G., ed., The Earth's Crust: Am. Geophys. Union Mon. Ser. 20, p. 626-675.
Majorowicz, J. A., Jones, F. W., and Jessop, A. M., 1986, Geotbermics of the Williston Basin in Canada in
relation to hydrodynamics and hydrocarbon occurrences: Geophysics, v. 51, no. 3, p. 767-779.
Neuzil, C.E., 1993, Low fluid pressure within the Pierre Shale: a transient response to erosion: Water Resources
Res., v. 29, no. 7, p. 2007-2020.
Neuzil, C.E., 1995, Abnormal pressures as hydrodynamic phenomena: Amer. Jour. Sci., v. 295, no. 6, p. 742-
786.
116 GOSNOLD

Neuzil, C. E., Bredehoeft, J. D., and Wolff, R. G., 1984, Leakage and fracture permeability in the Cretaceous
shales confming the Dakota aquifer in South Dakota, in Jorgensen, D. G., and Signor, D. C., eds.,
Geohydrology of the Dakota Aquifer: Nat. Water Well Assoc., Worthington, OH, p. IlJ-l20.
Rahn, P. H., 1985, Groundwater stored in the rocks of western South Dakota, in Rich, F. J., ed., Geology of the
Black Hills of South Dakota and Wyoming (2 nd ed.): Am. Geol. Inst., p. 154-174.
Roy, R. F., Blackwell, D. D., and Decker, E. R., 1972, Continental heat flow, in Robinson, E. C., ed., The Nature
of the Solid Earth: McGraw-Hill Book Co., New York, p. 506-543.
Sass, J. H., and Galanis, S. P., 1983, Temperatures, thermal conductivity and heat flow from a well in Pierre
Shale near Hayes, South Dakota: U.S. Geol. Survey Open-File Rept. 83-25, 10 p.
Sass, J. H., Lachenbruch, A. H., Munroe, R. J., Greene, G. W., and Moses, Jr., T. H., 1971, Heat flow in the
western United States: Jour. Geophys. Res., v. 81, no. 76, p. 6376-6414.
Schoon, R. A., Geology and hydrology of the Dakota Formation in South Dakota: South Dakota Geol. Survey,
Rept. ofInvest. 104,55 p.
Schoon, R. A., and McGregor, D. J., 1974, Geothermal potentials in South Dakota: South Dakota Geol. Survey,
Rept. ofInvest. 110, 76 p.
Schultz, L. G., 1964, Quantitative interpretation and mineralogical composition from X-ray and chemical data
for the Pierre Shale: U.S. Geol. Survey Prof. Paper 391-C, 31 p.
Schultz, L. G., 1965, Mineralogy and stratigraphy of the lower part of the Pierre Shale, South Dakota and
Nebraska: U.S. Geol. Survey Prof. Paper 392-B, p. BI-BI9.
Smith, L., and Chapman, D. S., 1983, On the thermal effects of groundwater flow: Jour. Geophys. Res., v. B88,
no. I, p. 593-608.
Swenson, F. A., 1968, New theroy ofrecharge in the artesian basin of the Dakotas: Geol. Soc. America Bull.,
v. 79,no.2,p. 163-182.
Sta1lman, R. W., 1963, Computation of groundwater velocity from temperature data: U.S. Geol. Survey Water
Supply Paper I 554-H, p. 36-46.
Tourtelot, H. A., 1962, Preliminary investigation of the geologic setting and chemical composition of the Pierre
Shale, Great Plains region: U.S. Geol. Survey Prof. Paper 390,74 p.

OTHER REFERENCES

Bachu, S., and Hitchon, B., 1996, Regional-scale flow of formation waters in the Williston Basin: Am. Assoc.
Petroleum Geologists Bull., v. 80, no. 2, p. 248-264.
Blackwell, D. D., 1971, The thermal structure of the continental crust, in Heacock, J. G., ed., The Structure and
Physical Properties of the Earth's Crust: Am. Geophys. Union, Geophys. Mon. Ser., v. 14, p. 169-184.
Combs, J., and Simmons, G., 1973, Terrestrial heat flow determinations in the north central united States: Jour.
Geophys. Res., v. 78, no. 2, p. 441-461.
Deming, D., Sass, J. H., Lachenbruch, A. H., and DeRito, R. F., 1992, Heat flow and subsurface temperature as
evidence for basin-scale ground-water flow, North Slope of Alaska: Geol. Soc. America Bull., v. 104,
no. 5, p. 528-542.
Donaldson, I. G., 1962, Temperature gradients in the upper layers of the earth's crust due to convective water
flows: Jour. Geophys. Res., v. 67, no. 9, p. 3449-3459.
Head, W. J., Kilty, K. T., and Knottek, R. K., 1978, Maps showing temperatures and configurations of the tops
of the Minnelusa Formation and the Madison Limestone, Powder River Basin, Wyoming, Montana, and
adjacent areas: U. S. Geol. Survey Open-File Rept. 78-905, 2 sheets.
Hildenbrand, T. G., and Kucks, R. P., 1985, Model of the geothermal system in southwestern South Dakota from
gravity and aeromagnetic studies, in Hinze, W. J., ed., The Utility of Regional Gravity and Magnetic
Anomaly Maps: Soc. Expl. Geophys., p. 258-266.
Lachenbruch, A. H., and Sass, J. H., 1977, Heat flow in the United States and the thermal regime of the crust,
in Heacock, J. G., ed., The Earth's Crust: Am. Geophys. Union Mon. Ser. 20, p. 626-675.
Majorowicz, J. A., Jones, F. W., and Jessop, A. M., 1986, Geotherrnics of the Williston Basin in Canada in
relation to hydrodynamics and hydrocarbon occurrences: Geophysics, v. 51, no. 3, p. 767-779.
Smith, L., and Chapman, D. S., 1983, On the thermal effects of groundwater flow: Jour. Geophys. Res., v. B88,
no. I, p. 593-608.
THERMAL INSULATION BY LOW THERMAL
CONDUCTIVITY SHALES: IMPLICATIONS FOR
BASIN-SCALE FLUID FLOW AND HEAT TRANSPORT

Jeffrey A. Nunn, Guichang Lin, and Libo Zhang

Louisiana State University


Baton Rouge, Louisiana, USA

ABSTRACT

Organic-rich, fine-grained sediments, such as shale and mudstone, are deposited in


many sedimentary basins. These sediments are characterized by thermal conductivities in the
range of 0.8 to 1.45 W/moC, lower by a factor of2 than other abundant sedimentary rocks.
Thus, they act as a thermal blanket retaining heat energy within the sedimentary basin. A two-
dimensional finite-element model of fluid flow and heat transport has been used to study the
insulating effect oflow thermal conductivity shales on an uplifted foreland basin. Our model
section cuts through the Arkoma Basin to Ozark Dome and terminates near the Missouri River,
west of St. Louis. Topographically driven recharge is assumed to be the major driving force
for regional groundwater flow. Our model results show that the thermal gradient in the top
layer can be up to 0.08°Clm, higher than normal geothermal gradients of 0.02-0.04 °Clm.
High temperatures in the discharge region of the basin are caused by thermal insulation and
high rates of advective heat transport associated with basin-scale fluid flow along basal
aquifers (Darcy velocities greater than 6 m/yr.). Our model results show high temperatures
(-90°C) in the groundwater discharge region at shallow depths «1 km), even with a typical
continental basal heat flow of60 mW/m2 . A modest 34% decrease in thermal conductivity of
the top layer of sediment (2.0 to 1.33 W/m 0c) elevates temperatures in underlying basin
sediments by 15-20°C for millions of years. This change in temperature history is roughly
equivalent to the thermal effect of fluid flow by topographic recharge for the model basin
considered here. Initial steady-state conduction temperatures range from 45 ° C to 7.5 °C higher
depending on the thickness of the insulating shale layer. Elevated temperatures will persist for
as long as the low thermal conductivity shales cover the basin.

117
118 NUNN, LIN, AND ZHANG

INTRODUCTION

Numerical simulations of basin-scale heat transport by fluid flow have become a


valuable tool for unraveling the evolution of sedimentary basins because chemical reactions
associated with diagenesis and ore deposition are dependent on temperature (Nunn and
Deming,1991). For example, the relatively high temperatures (100-150 0c) at shallow depths
«1.5 Ian) inferred from studies of fluid inclusions associated with Late Paleozoic Mississippi
Valley-type (MYT) lead-zinc deposits (SveIjensky, 1986) places a restrictive constraint on ore
formation by regional fluid flow (Bethke, 1986). Modeling studies by Cathles and Smith
(1983), Bethke (1985), Ge and Garven (1989), and Deming, Nunn, and Evans (1990) have
demonstrated that compaction driven flow rates are too low to generate the high temperatures
at shallow depths required for MYT ore deposits unless flow is spatially focused or episodic.
Although topographically driven recharge can produce the required high temperatures at
shallow depths in the regional discharge zone (Garven and Freeze, 1984; Bethke and Marshak,
1990; Garven, and others, 1993), it can do so only under a limited range of geological
conditions, such as high basal heat flow, the presence of a high permeability aquifer at depth,
or a narrow, deep basin geometry, that may not be representative of the geology or conditions
at the sites of ore deposition (Deming and Nunn, 1991).
The temperature history of a sedimentary basin also has important implications for the
maturation and migration of hydrocarbons (Nunn and Sassen, 1986). Thermal degradation of
kerogen to residual, oil, methane, and carbon dioxide is a set of first-order rate kinetic reactions
which are dependent on temperature. The reaction rate approximately doubles for every 10°C
increase in temperature (Lopatin, 1971). More recent descriptions of the thermal maturity of
organic matter represent kerogen as a heterogeneous mixture of organic molecules that consist
of a variety of C-O, H-O, and H-C bonds (Tissot, Pelet, and Ungerer, 1987; Burnham and
Sweeney, 1989). Each class of organic compounds has different activation energies (or
threshold temperatures) for thermal degradation. Thus, organic maturation depends on
maximum temperatures as well as temperature history.
Low thermal conductivity sediments, such as shale and mudstone, prevail in many
sedimentary basins. These rocks are characterized by thermal conductivities in the range of
0.8 to 1.45 W/m °c, lower than that of other sedimentary rocks (Blackwell and Steele, 1989;
Pollack and Cercone, 1994; Herrin and Deming, 1996). Several hundred meters of low
thermal conductivity sediments at the surface substantially increases heat retention generating
elevated temperatures within underlying basin sediments. In this study, we use a two-
dimensional finite element model of fluid flow and heat transport to examine the insulating
effect of low thermal conductivity sediments on the thermal evolution of an uplifted foreland
basin.

NUMERICAL MODEL
Mathematical Representation

Fluid movement in response to pressure and density gradients in a two-dimensional


fixed reference frame is described by

(1)
INSULATION BY LOW THERMAL CONDUCTNITY SHALES 119

where 4> is porosity, P is fluid density, Po is an arbitrary reference density (1000 kglm\ g is
the acceleration resulting from gravity, Il is dynamic viscosity, t is time, z is elevation, x is a
horizontal coordinate perpendicular to z, kx and kz are permeability in the x and z directions,
and h is an equivalent freshwater head. All symbo'Is and values used in this study are defined
in Table 1.
Equation (1) is derived by assuming conservation of mass and the general form of
Darcy's law (Smith and Chapman, 1983; Garven and Freeze, 1984). Equivalent freshwater
head is (Smith and Chapman, 1983)

P (2)
h=-+z
Pog

Table 1. Nomenclature.

Symbol Description SI Values


Cw fluid specific heat J/kgOC 4200
Cr bulk specific heat J/kgOC 837
g Gravitational acceleration mls2 9.8
kx horizontal sediment permeability m2 see Table 2
kz vertical sediment permeability m2 see Table 2
A.x horizontal sediment thermal conductivity W/moC see Table 2
A. z vertical sediment thermal conductivity W/moC see Table 2
Il Dynamic viscosity of pore water Pas computed
cjl Sediment porosity % see Table 2
p density of water kglm 3 1000
po reference density of pore water kglm 3 1000
pr bulk density of rock kglm 3 2500
P fluid pressure MPa computed
T Temperature DC computed
Vx horizontal Darcy flux mls computed
Vz vertical Darcy flux mls computed

Heat transport by conduction and advection in a two-dimensional fixed reference frame


is described by

where T is temperature, C w is fluid specific heat, Cr is bulk (grains and pore fluids) specific
heat of the rock, Pr is bulk density of the rock, Ax and A..z are bulk thermal conductivities in the
x and z directions, Vx and Vz are specific discharges or Darcy velocities in the x and z
directions. Radioactive heat generation is ignored in our calculations. Equation (3) is derived
through an assumption of conservation of energy and Fourier's law of heat conduction (Smith
and Chapman, 1983; Garven and Freeze, 1984).
Equations (l) and (3) are solved numerically using a Galerkin finite-element technique
with linear basis functions applied over quadrilateral elements (Baker and Pepper, 1991).
120 NUNN, LIN, AND ZHANG

Transient solutions to the resulting system of equations are obtained using the 2-D finite-
element Akcess.Basin™ software package (Cathles and others, 1995). In the present study,
the solution to each nonlinear equation is computed iteratively using a fully implicit, Newton-
Raphson technique.
Model results were tested a posteriori for heat conservation by comparing the net
change in heat content to net heat flux across the boundaries between successive model time
steps. Individual time step and cumulative errors were 5% or less which is the accuracy to
which we can estimate surface heat flux from model temperatures. Steady-state temperatures
were calibrated against Smith and Chapman (1983).

Model Basin

Our model results are constructed to simulate fluid flow and heat transport in the
eastern half of the Arkoma Basin during the Late Paleozoic along a transect that begins in the
undeformed foreland north of Little Rock, Arkansas, extends northward over the Ozark Dome,
and terminates near the Missouri River, west of St. Louis (Fig. 1). The model basin geometry
and hydrostratigraphy (Figs. 2 and 3) are taken from Garven and others (1993). Stratigraphic
units C and D are basal aquifers with high permeabilities (Figs. 2 and 3, and Table 2).
Following Garven, Persan, and SveIjensky (1993), we use a permeability anisotropy of 100
Oexlkz) as appropriate for large-scale flow through layered sequences of sedimentary rocks of
contrasting permeabilities (Deming, 1993). See Garven and others (1993) for additional
details regarding location, stratigraphy, and lithology of the model profile (their profile A-A').
Physical parameters used in our model are based on lithology and hydrostratigraphy of the
Ozark region (Fig. 3) and are listed in Table 2. We assume that basement rocks are
impermeable. Freeze and Witherspoon (1967) have shown that when a high permeability layer
(sediments) overlies a low permeability layer (basement) there is negligible penetration offluid
into the low permeability layer.

200 kilometers
'----1._....1,
200 Miles .; Illinois
'--_--1._ _--',
.........
Iowa
••.•.•• Indiana Ohio
Nebraska --"-"-"-"-"-"-"-'-'.'" : \....
...
!., ~
'
,
:",i
Missouri
.... - .... -......
~;"\'"
~I
Illinois B~sin""
I .'
... ,~ , .. , ..">.!
Kansas
':'-····.'·./T
,-- \ I. ' .. I' ""

,: Ozark - ... .;~~, Kentucky ,-.'r.,


. -__ Uplift }. . ' .. ' ...-
'''' ,. .. ~. : -
_,- _ . ..".;' -' ,.
';_ . _'- _. _. _. _. '··_··_··_··-l ~.:~;::~ . -·-··-~/~<k·~~~~~~:s;··-·· Tenness~~ ,.-J.J
;
:
Oklahoma
..... i
.i·········:-·"':"::"·. ... ,,'
i MISSOURI DISTRICT'-- .. _ .. __ \ .. - .... - .. - .. ~:
,
"-Arkoma Basin : ,:._ .. -.. -.. -.- ._ .. _.. - . , "- ..
I :: . . . . . . .\ . . . . . . . . . . . . . . . . . ~ .:t: \"
".-... _.. ,... : .•••• OUACHITA IIII'Ns ":;/ , Georgia
~~:~:··'··,/·-··'··, . ,l Arkansas' \ Mississippi Alabama' ,
.. _.-----_._ .. - .. _.. ..; \

Figure 1. Location map for hydrogeologic section model for Arkoma Basin and Ozark Uplift.
INSULATION BY LOW THERMAL CONDUCTIVITY SHALES 121

In order to minimize differences in the flow fields between model basins with different
thermal conductivity distributions and to demonstrate that thermal insulation by low thermal
conductivity shales is a first-order effect, fluid density is assumed to be constant. As a result,
the left-hand side of Equation (1), the last term on the right-hand side in Equation (1), and the
second term on the left-hand side of Equation (3) are all set to zero. The fluid-flow field is
stationary, independent of time, except for viscosity variations resulting from changes in the

South Model Hydrostratigraphy North


-1
0 A
neous
S 1
~
D
2
....
.c~

~ 3
Q,I
Q 4 Basement
C
5
6
0 50 100 150 200 250 300 350
Distance, km
Figure 2. Hydrostratigraphic model of eastern half of Arkoma Basin used in our study. See Figure I for location.
Geometry and hydrostratigraphy are taken from Garven and others (1993). Lithology and hydrostratigraphy are
shown in Figure 3. Physical parameters for each unit are listed in Table 2. Thermal conductivity of Unit A
(Pennsylvanian shales) is principal experimental variable examined in this paper.

temperature history. In our simulations in which fluid density was allowed to vary as a linear
function of temperature (not shown), the temperature-induced fluid density inversion produces
some free convective overturn in the deepest part of the basin. In a sedimentary basin, where
the permeability structure is more heterogeneous than modeled here, it is unclear whether free
convection would actually take place (Bjorlykke, Mo, and Palm, 1988; Lopez and Smith,
1995). Moreover, increasing pore water salinity with depth would counteract thermal
buoyancy effects (Hanor, 1988; Evans and Nunn, 1989; Deming and Nunn, 1991; Garven and
others, 1993). In all simulations, buoyancy driven fluid flow has only a secondary effect on
the temperature distribution. Dynamic viscosity, ~,is calculated as a function of temperature
and pressure using an empirical formula (Kestin, Ezzat Khalifa, and Correia, 1981).
The fluid-flow region within the basin is discretized with a finite-element mesh
containing 33 rows and 81 columns with variable element sizes. For the basement, we use a
coarse finite-element mesh of 12 rows and 81 columns to compute heat conduction in the
basement down to 15 km depth.

Boundary and Initial Conditions

Boundary conditions for both models are similar to those used by Garven and others
(1993). The upper boundary, which represents the watertable, is idealized to have an uniform
slope of 1.2 km for 360 km (Fig. 2). Hydraulic head along the top boundary is fixed at
elevations corresponding to the watertable [see Eq. (2)]. Temperature along the top was
122 NUNN, LIN, AND ZHANG

SYSTEM
1
LlTIIOLOGY
I/YDROSTRATIGRAPIlIC
UNIT

~ • 'c
III
o
«
PE NSYL V A IAN

1---- - - 3 1 2 .', : '~ "

MISSISSIPPIAN SPRINGFIELD
AQUIFER
1-----345
DEVONIAN
1 - - - -- - 395
SILURIAN
f -- - -- 435 ......... .

ORDOVICIAN

OZARK AQUIFER

1--- - - - 500 .

CA 1I3RIAN
AQUITARD,
S, FRANCOIS
AOUIFER

f - -- --580 ~::::::~~11. SIMON AQUWER


PRECAMBRIAN :::::::::: AQUITARD

L:::::\ SIIALE [i.·:.< 1 SA DSTO E

II CARBO ATE
,',
mill
»»,","
,,,, , BASEME T

Figure 3. Lithologic column and hydrostratigraphy for Ozark region. Modified from Garven and others (1993).

fixed at 20°e. Constant heat flow of 60 mW/m 2 , which is characteristic for Paleozoic
sedimentary/metamorphic terrains (pollack, Hurter, and Johnson, 1993), was maintained along
the lower boundary. Garven and others (1993) used a basal heat flow of80 mW/m2 . Observed
heat flow does vary spatially in the study area (pollack, Hurter, and Johnson, 1993). We have
opted to use a constant heat-flow value in order to isolate the thermal effects of an insulating
layer of variable thickness and to be consistent with the earlier study by Garven and others
(1993). There is no heat transport or fluid flow across the side boundaries. Initial conditions
are no fluid flow and steady-state heat conduction.
INSULATION BY LOW THERMAL CONDUCTIVITY SHALES 123

Table 2. Model parameters constant in simulations.

kx kz Ax,Az <l>

Unit A 3.2 x 10- 14 m2 3.2 x 10- 16 m2 1.33-2 W/moC 10


UnitB 3.2 x 10- 14 m2 3.2 x 10- 16 m2 2 W/moC 10
Unit C 9.5 x 10- 13 m2 9.5 x 10- 15 m2 2 W/moC 20
UnitD 1.9 x 10- 12 m2 1.9 x 10- 14 m2 2 W/moC 25
Igneous Rocks 3.2 x 10- 20 m2 3.2 x 10- 22 m2 2.9W/moC
Basement 0.Om 2 0.Om 2 2.9 W/moC

MODEL RESULTS

In this study, we consider two situations: a reference model where the uppermost
stratigraphic layer (unit A in Fig. 2) has the same thermal conductivity (2.0 W/m 0c) as other
basin sediments and thus is not insulating; and a low thermal conductivity shale (LTCS) model
where the top unit has a thermal conductivity (1.33 W/m 0c) typical of shales (Blackwell and
Steele, 1989) which acts as an insulating layer trapping heat in basin sediments. We use
thermal conductivities of2.0 W/m °C for all other sediments (units B, C, and D in Fig. 2) and
2.9 W/m °C for igneous rocks in the st. Francois Mountains and basement rocks below the
basin. Units B, C, and D are dominantly carbonate and sandstone whereas unit A is
dominantly shale (Figs. 2 and 3). These values are typical for our study area (Meert, Smith,
and Fishkin, 1991; Garven and Freeze, 1984).

Initial Temperature Distribution

For the reference and LTCS models, the conductive thermal gradient is 0.03°C/m and
0.045 °C/m across the top layer, respectively (Fig. 4). A higher temperature gradient in the top
layer leads to a significant increment in temperature for the underlying sediments which ranges
from 45°C in the south to 7.5°C in the north (Fig. 5). Differences in the amount of
temperature increase in underlying sediments across the basin is primarily the result of
variations in thickness ofthe insulating shales (unit A in Fig. 2). There is some heat refraction
(bend in the isotherms at the LTCS-sediment and sediment-basement interfaces in Fig. 4) as
heat energy tends to follow preferentially high thermal conductivity pathways. But this effect
is extremely minor because of the small dip angles of the stratigraphic layers (note that Fig.
4 has a vertical exaggeration of 20 to 1). The large, but uniform increment in temperature
below the insulating layer (Fig. 5) contrasts with a depth dependent temperature increment
resulting from an increase in basement heat flow (Pollack and Cercone, 1994). Thus, without
increasing basal heat flow, the insulating layer causes additional storage of thermal energy in
the model basin. Elevated temperatures will persist for as long as the insulating shales remain.
Evidence for present-day thermal insulation by low thermal conductivity sediments
include: Karoo coal (0.1 °C/m) of eastern Bostswana, Africa (Ballard, Pollack, and Skinner,
1987), Jurassic coal deposits (0.14°C/m) in Australia (Beck, 1976), Gondwana coals (0.04 to
0.25°C/m) in India (Kayal and Christoffel, 1982), carbonaceous rocks in a borehole
(0.06°C/m) in northwestern Colorado (Blackwell and Steele, 1989), and coals and black shales
(0.06 to 0.3 °C/m) in western Pennsylvania (Cercone, Deming, and Pollack, 1996).
124 NUNN, LIN, AND ZHANG

- Ii'" --- • - - •• ~., ~-.- 0 20

s
..::s:::
:§?F~'
-----=
--
......
~-~
- .
'ow

100 ....
•• _.~~~
::::::----~~- .• _._
_~ •
....... - •. __ _
-=fr .< ------____
------.-- .-. -'-__
. . . . -. -. -
~ 2 . ~- .---~_
~
3~
140

-.---==-.------1'
~ . ... ............... ...~"""- ~

. = 20 I
4 ..... ..........-_ .....-------~

5r-:_~8<~-
I __. · .--.__ -.-.-___ ~ t
=
0 Ma .
6r I I ........ , I i ~- I I
o 50 100 150 200 250 300 350
Di tance, km
-11--
...~ -. .-::.:. . .-~.~~::2---. 20
0j"-' ---60 ............'- ... _ _ _
- __ .~_
· r.~ • - •_ _ _••

S
..::s:::~
l l--;;~::i;~~~~=-:
i'
140-. - -'-' .-:::::==-=____ ------_. -----
-=.... 2 ---.-.
I.
_ ___ . .
--.........,
~ ~.

-.::::.<,'" -. - - - - -'__-----..,. ". ' - -


.~_ -.~.-.

-
o. ,.....---........ • ___ •
180 ""'-" '-..... -'. -
c. 3
Q 4 -
/' ......" '. --.- -- . ..
5
.. / '
' 220 ,
"' '-... _. .. •
.--.-"." ~ ~----
A. = 1.33
-. - . t= 0 Ma
6 ----, - - - -,,---'- , I - " - - - - - "-
o 50 100 150 200 250 300 350
Distance, km
Figure 4. Initial temperature distribution assuming steady-state heat conduction for reference model (top) and
low thermal conductivity shale model (bottom). Thin black lines outline basin stratigraphy. Thick gray lines
are temperature contours in QC.

0'\\
02~
. onh lscharge
ID
-1 SouthlRecharge
0
\ 20 1.33 A
E 0.2 \ E A
..!<::
~
...c 0.4
0.. \ ~33 B ...c 2
, \ 0..

\\
(1,)
(1,)
Q 0.6 Q 3 B

0.8 4
C&D 2.0

1
\ '
5
20 40 60 80 100 20 60 100 140 180 220 260
Temperature, °C Temperature, °C

Figure 5. Temperature versus depth at two locations in model basin: (left) NorthlDischarge area; and (right)
South/Recharge area. Thick gray lines are initial temperatures (Fig. 4) and thin black lines are model
temperatures after 0.2 Ma of fluid flow (Figs. 7 and 8). Numbers on lines are thermal conductivity in W/mQC
for top layer of sediments (unit A) in model basin. Thin horizontal lines indicate stratigraphic boundaries.
INSULATION BY LOW THERMAL CONDUCTIVITY SHALES 125

Fluid Flow

Figure 6 shows specific discharge from the LTCS model at an elapsed time of 0.2 Ma.
The computed velocity fields from both models are similar at this time step. The velocity
arrows are shown with the same vertical exaggeration as the physical dimensions ofthe model
and thus show the true direction of fluid flow in the model basin. On the left side ofthe model,
fluids are driven downward by the topographic gradient to recharge the basal aquifers with
vertical Darcy velocities of a few centimeters per year. In the discharge area on the right side
ofthe model, flow is upward from the basal aquifers. In the central region ofthe model basin,
fluid flow in the aquifers is mainly horizontal (note vertical exaggeration in Fig. 6 is 20 to 1).

-
-1 '
0

e 1

-
"----~-~
~
2
~
Q
QJ
.3
~

~ 4

5 ~L 3 m/yr
6
0 50 100 150 200 250 300 350
Distance, km
Figure 6. Specific discharge at elapsed time of 0.2 My for low thermal conductivity shale (J.. = 1.33 W/m°C)
model. Discharge vectors are shown with same vertical exaggeration as physical dimensions (20: 1) and thus
show true direction of fluid flow in model.

Computed maximum horizontal Darcy velocities in the basal aquifers are


approximately 5 m/yr for the reference model and more than 6 m/yr for the LTCS model.
Darcy velocities from our models are similar to those obtained by Garven and others (1993)
and Deming and Nunn (1991). Higher Darcy velocities in the LTCS model result from higher
basinal temperatures caused by the insulating shales (Fig. 4). Higher temperatures reduce fluid
viscosities (Kestin, Ezzat Khalifa, and Correia, 1981) and thus increases flow rates. Based on
our calculations using empirical equations developed by Kestin, Ezzat Khalifa, and Correia
(1981), an increase of35 °C can cause at least a 45% decrease in fluid viscosity at temperatures
between 20 and 55°C, and about a 30% decrease at temperatures between 90 and 120°C.
Thus, thermal insulation by overlying low thermal conductivity shales increases the efficiency
of advective heat transport.
Because our model does not consider changes in watertable topography and sediment
properties with time or the effect of temperature on fluid density, fluid flow at elapsed times
other than 0.2 Ma is similar to Figure 6. The primary change in the flow field with time is that
as the recharge area cools, fluid viscosity increases and thus fluid velocities decrease.

Temperature Distributions

Figures 7 and 8 show the temperature conditions at an elapsed time of 0.2 Ma for the
reference and LTCS models, respectively. In both models, the temperature field has been
changed dramatically by fluid advection. Most of the left side of the model basin has been
126 NUNN, LIN, AND ZHANG

cooled by the downward flow of freshwater. On the left side ofthe model basin, the extent of
cooling is more than 60°C. More rapid cooling of the recharge area in the LTCS model is
because of higher flow velocities. Conversely, temperature in the discharge area (upper, right
side of the model basin) has increased. Changes in the thermal regime with time are most
apparent in the temperature anomaly (Figs. 7 and 8). The temperature anomaly is the
temperature field minus the initial thermal regime. Maximum warming is about 20 ° C in the
reference model and about 40°C in the LTCS model (Figs. 7 and 8). In both examples, the
maximum warming occurs in the basal aquifer at a depth ofless than 1 km where the aquifer
terminates against low permeability igneous rocks in the St. Francois Mountains. Our results
for the reference model are consistent with the results of Garven and others (1993) when
adjusted for the higher basal heat flow, 80 mW/m2, and different sediment thermal
conductivities used in their study.

i i i
50 ]00 150 200 250 350
Distance, km

i r I i
]00 150 200 250
Distance, km
Figure 7. Temperature (top) and temperature anomaly (bottom) for reference model simulation at elapsed time
of 0.2 Ma. Temperature anomaly is temperature minus initial temperature. Thin black lines outline basin
stratigraphy. Thick gray lines are temperature or temperature anomaly contours in DC.

Thermal insulation by a low thermal conductivity layer also retains heat during
discharge of hot fluids by reducing surface heat loss. This is indicated by the large thermal
gradient (O.077°C/m) in the insulating shales (unit A in Fig. 2) and the low thermal gradient
(0.019°C/m) in the underlying sediments (Fig. 5). Thermal gradients taken at the same depths
in the reference model are 0.056 and O.027°C/m, respectively.
In the discharge region, thermal insulation by low thermal conductivity shale has as
dramatic an effect on sediment temperatures (-20°C increase) as fluid flow itself (Fig. 5). In
INSULATION BY LOW THERMAL CONDUCTIVITY SHALES 127

contrast, cooling of basin sediments by downward movement of cold meteoric water is clearly
the dominant thermal effect in the recharge zone (Fig. 5).
Figure 9 shows temperature versus time at points in the recharge and discharge regions
of the basal aquifer for both models. Temperatures in the recharge area slowly cool with time
as the model basin reaches thermal equilibrium during a few million years. However,
isotherms in basin sediments in the discharge region remain almost unchanged between 0.2
to 1.0 Ma (Fig. 9). Thus, the LTCS model maintains elevated temperatures of 15 to 20°C
above the reference model on time scales of millions of years.

--_.. - - .. 20 ..

-O1--1J -..-.~:- -----~-- . . -......


..-•.• ~ ....- - - - -.. _ •
F------40~-.... -'/r~L
\,.
''''1 ..,-.r

- 20 0 -~.. - - _ ..._ ..........


-----~-;;- . r - -
~ 1 ~20 .c:·;----
~~ 2 ~--r)j.!-~;-- ,
-= 160
Q. 3 ~ - . I /
.~,,?-~
/'.'
..... ~.-- -- -.'-

o 4{~1 /
5V/ I I 1..=2.0 I
6 4~--',1------,'-------r1______'I------'I------"~t-=~0~.2~M~la
o 50 100 150 200 250 300 350
Distance, km
Figure 8. Temperature (top) and temperature anomaly (bottom) for low thermal conductivity shale (/.. = 1.33
W/m°C) model simulation at elapsed time of 0.2 Ma. Temperature anomaly is temperature minus initial
temperature. Thin black lines outline basin stratigraphy. Thick gray lines are temperature or temperature
anomaly contours in °C.

DISCUSSION AND CONCLUSIONS

The low thermal conductivity shales used in the LTCS model causes the topmost
sedimentary layer to have a high thermal gradient, which in turn elevates temperatures in
underlying sediments, even for an average basal heat flow (60 mW/m2). Thermal insulation
causes additional storage of thermal energy in the model basin even in the absence of fluid
flow. Depending on the thickness ofthe insulating layer, steady-state conduction temperatures
in underlying sediments were 7.5 to 45°C higher for the LTCS model. Elevated temperatures
persist for as long as the low thermal conductivity shales reside on top ofthe basin.
128 NUNN, LIN, AND ZHANG

Higher average temperatures in the model basin significantly decreases fluid viscosity,
and thus increases fluid flow. Our results indicate that the viscosity difference produces an
increase in maximum Darcy velocity of more than 1 mlyr in the early stages of fluid flow «0.2
Ma). Thus, heat energy is more efficiently transported by groundwater flow.

230 \
210 \
U 190
o ~ 170
1.33
~
.aro
ISO ,~
'-.....·2.0 .........-~
"-------
1i:5 130
0..
E 110
~ 90 1 . 3 3 - - - - -_
_ - - - 2.0 - - - - -
70
50~--~~--~~~~r---~
o 0.25 0.5 0.75
Time, Ma
Figure 9. Temperature versus time at two locations in unit C of model basin (Fig. 2): (thin black lines)
Discharge area and (thick gray lines) Recharge area. Numbers on lines are thermal conductivity in W1m 0 C for
top layer of sediments (unit A) in model basin.

Low thermal conductivity shales also trap heat energy transported by fluid flow. This
is indicated by a thermal gradient in the low thermal conductivity shale layer in the discharge
area up to 0.07rC/m about 2.5 times the reference model steady-state conduction thermal
gradient. Compared to the reference model, surface heat flow at 0.2 Ma is almost 10% lower.
In contrast to model results with an elevated basal heat flow, thermal insulation does
not produce a large thermal gradient in underlying sediments. Most sediments in the discharge
area have thermal gradients less than 0.02°C/m.
For the model basin considered here, a modest 34% decrease in the thermal
conductivity (2.0 to 1.33 W/m°C) of the topmost stratigraphic layer elevates temperatures in
underlying sediments in the discharge zone by 15-20°C for millions of years. This is
equivalent to the thermal effect of basin-scale fluid flow in the reference model. Thus, thermal
insulation by low thermal conductivity sediments should be considered in modeling of organic
maturation, diagenesis, and ore formation.
The magnitude and distribution of elevated temperatures in basin sediments caused by
thermal insulation depends on the thickness and thermal conductivity ofthe insulating layer
as well as the permeability structure, aspect ratio, and basal heat flow ofthe sedimentary basin.

ACKNOWLEDGMENTS

We would like to thank Henry Pollack and an anonymous reviewer for helpful
comments on this manuscript. Grant Garven provided additional information about model
parameters used in Garven and others (1993). This research was supported by grants EAR
9316627 and EAR 9421065 from the National Science Foundation.
INSULATION BY LOW THERMAL CONDUCTIVITY SHALES 129

REFERENCES

Baker, A. J., and Pepper, D. W., 1991, Finite elements 1-2-3: McGraw-Hill Book Co., New York, 341 p.
Ballard, S., Pollack, N. H., and Skinner, N. J., 1987, Terrestrial heat flow in Botswana and Namibia: Jour.
Geophys. Res., v. B92, no. 7, p. 6291-6300.
Beck, A. E., 1976, The use of thermal resistivity logs in stratigraphic correlation: Geophysics, v. 41, no. 2, p.
300-309.
Bethke, C. M., 1985, A numerical model of compaction-driven groundwater flow and heat transfer and its
application to the paleohydrology of intracratonic basins: Jour. Geophys. Res., v. B90, no. 8, p. 6817-
6828.
Bethke, C. M., 1986, Hydrologic constraints on the genesis of the Upper Mississippi Valley district from Illinois
Basin brines: Econ. Geology, v. 81, no. 2,p. 233-249.
Bethke, C. M., and Marshak:, S., 1990, Brine migrations across North America - the plate tectonics of
groundwater: Ann. Rev. Earth Planet. Sci., v. 18, p. 287-315.
Bjorlykke, K., Mo, A., and Palm, E., 1988, Modeling of thermal convection in sedimentary basins and its
relevance to diagenetic reactions: Marine and Petroleum Geology, v. 5, no. 4, p. 338-350.
Blackwell, D. D., and Steele, J. L., 1989, Thermal conductivity of sedimentary rocks: measurement and
significance, in Naeser, N. D., and McCulloh, Th. H., eds., Thermal History of Sedimentary Basins:
Methods and Case Histories: Springer-Verlag, New York, p. 13-36.
Burnham, A. K., and Sweeney, J. J., 1989, A chemical kinetic model of vitrinite maturation and reflectance:
Geochim. Cosmochim. Acta, v. 53, no. 2, p. 2649-2657.
Cathles, L. M., and others, 1995, The Akcess.Basin finite element simulation code: U.S. Department of Energy
Final Report DE-FC22-93BC 14961, Washington, D. C.
Cathles, L. M., and Smith, A. T., 1983, Thermal constraints on the formation of Mississippi Valley-Type lead-
zinc deposits and their implications for episodic basin dewatering and deposit genesis: Econ. Geology,
v. 78,no. 5,p. 983-1002.
Cercone, K. R., Deming, D., and Pollack, N. H., 1996, Insulating effects of coals and black shales in the
Appalachian basin, western Pennsylvania, in Stanton, R., ed., Collected Papers from Twelfth Annual
Meeting of Society for Organic Petrology, Organic Geochemistry, v. 24, no. 2, p. 243-249.
Deming, D., 1993, Regional permeability estimates from investigations of coupled heat and groundwater flow,
North Slope of Alaska: Jour. Geophys. Res., v. B98, no. 9, p. 16,271-16,286.
Deming D. and Nunn, J. A., 1991, Numerical models of brine migration by topographically driven recharge:
Jour. Geophys. Res., v. B96, no. 2, p. 2485-2499.
Deming, D., Nunn, J. A., and Evans, D. G., 1990, Thermal effects of compaction driven groundwater flow from
overthrust belts: Jour. Geophys. Res., v. 95, no. 5, p. 6669-6684.
Evans, D. G., and Nunn, J. A., 1989, Free thermohaline convection in sediments surrounding a salt column: Jour.
Geophys. Res., v. B94, no. 9, p. 12,413-12,422.
Freeze, R. A., and Witherspoon, P. A., 1967, Theoretical analysis of regional groundwater flow: 2. Effect of
watertable configuration and subsurface permeability variation: Water Resources Res., v. 3, no. 2, p.
623-634.
Garven, G., and Freeze, R. A., 1984, Theoretical analysis of the role of groundwater flow in the genesis of
stratabound ore deposits, 2, Quantitative results: Am. Jour. Sci., v. 284, no. 10, p. 1125-1174.
Garven, G., Ge, S., Person, M. A., and Sverjensky, D. A., 1993, Genesis of stratabound ore deposits in the
Midcontinent basins of North America. 1. The role of regional groundwater flow: Am. Jour. Sci., v. 293,
no. 6, p. 497-568.
Ge, S., and Garven, G., 1989, Tectonically induced transient groundwater flow in foreland basins, in Price, R.
A., ed., Origin and Evolution of Sedimentary Basins and Their Energy and Mineral Resources:
Geophys. Monogr. Ser., v. 48, Am. Geophys. Union, Washington, D.C., p. 145-157.
Hanor, J. S., 1988, Origin and migration of subsurface sedimentary brines: Soc. Econ. Paleon. Mineral., Short
Course 21, 247 p.
Herrin, J. M., and Deming, D., 1996, Thermal conductivity of U.S. coals: Jour. Geophys. Res., v. BIOI, no. 11,
p.25,381-25,386.
Kayal, J. R and Christoffel, D. A., 1982, Relationship between electrical and thermal resistivities for differing
grades of coal: Geophysics, v. 47, no. 10, p. 121-129.
130 NUNN, LIN, AND ZHANG

Kestin, J., Ezzat Khalifa, H., and Correia, R. J., 1981, Tables of the dynamic and kinematic viscosity of aqueous
NaCI solutions in the temperature range 20-150°C and the pressure range 0.1-35 MPa: Jour. Phys.
Chern. Ref. Data, v. 10, no. 1, p. 71-77.
Lopatin, N. V., 1971, Temperature and geologic time as factors in coalification (in Russian): Akademiya Nauk
SSSR Izvestiya, Seriya Geologicheskaya, no. 3, p. 95-106.
Lopez, D. L., and Smith, L., 1995, Fluid flow in fault zones: analysis of the interplay of convective circulation
and topographically driven groundwater flow: Water Resources Res., v. 31, no. 6, p. 1489-1503.
Meert, J. G., Smith, D. L., and Fishkin, L., 1991, Heat flow in the Ozark Plateau, Arkansas and Missouri:
relationship to groundwater flow: Jour. Volcanology Geothermal Res., v. 47, no. 3-4, p. 337-347.
Nunn, J. A., and Deming, D., 1991, Thermal constraints on basin-scale flow systems: Geophys. Res. Lett., v. 18,
no. 5, p. 967-970.
Nunn, J. A., and Sassen, R., 1986, The framework of hydrocarbon generation and migration, Gulf of Mexico
continental slope: Gulf Coast Assoc. Geol. Soc. Trans., v. 35, p. 257-262.
Pollack, H. N., and Cercone, K. R., 1994, Anomalous thermal maturities caused by carbonaceous sediments:
Basin Research, v. 6, no. 1, p. 47-51.
Pollack, H. N., Hurter, S. J., and Johnson, J. R., 1993, Heat flow from the Earth's interior: analysis of the global
data set: Rev. Geophys., v. 31, no. 3, p. 267-280.
Smith, L., and Chapman, D. S., 1983, On the thermal effects of groundwater flow 1. Regional scale systems:
Jour. Geophys. Res., v. B88, no. 1, p. 593-608.
Sverjensky, D. A., 1986, Genesis of Mississippi Valley-type lead-zinc deposits: Ann. Rev. Earth Planet. Sci., v.
14, p. 177-199.
Tissot, B. P., Pelet, R., and Ungerer, P., 1987, Thermal kinetics of oil and gas generation: Am. Assoc. Petroleum
Geologists Bull., v. 71, no. 12, p. 1445-1466, 1987.
Waples, D. W., 1980, Time and temperature in petroleum formation: application of Lopatin's method to
petroleum exploration: Am. Assoc. Petroleum Geologists Bull., v. 64, no. 7, p. 916-926.
TEMPERATURE AND MATURITY EFFECTS
OF MAGMATIC UNDERPLATING IN THE
GJALLAR RIDGE, NORWEGIAN SEA

W. Fjeldskaar1 ,H. Johansen2 , T.A. Dodd3.4, and M. Thompson3,5

1 RF - Rogaland Research, Stavanger, Norway


2 Geologica a.s, Stavanger, Norway.
3 BP Norge UA, Forus, Norway
4 Present address: BP Exploration pIc, Houston, Texas, USA

5 Present address: BP Exploration Operating Company Ltd.,

Middlesex, United Kingdom.

ABSTRACT

The Gj aHar Ridge is. an area of complex geology situated in the west of the V",ring
Basin. Regional structural work in the area suggests multiple phases of rifting during basin
development. Seismic refraction data indicate the existence of a "low velocity layer" which can
be interpreted as magmatic underplating.
The aim of this study was to evaluate the effects of a possible magmatic underplating
over the GjaHar Ridge. More precisely, we have estimated the temperature effects and the
isostatic effects associated with the emplacement of a lower density 5-km thick magmatic body
at a depth of 15 - 20 km.
To model the basin evolution a series of stretching events were assumed: the opening
of the basin during the Permo-Triassic, an event in the Middle Jurassic to Early Cretaceous,
an event during the Middle to Late Cretaceous and an event in the Paleocene. Models were run
with and without underplating by a magmatic body emplaced during the Early Tertiary. The
effect of underplating has dramatic consequences on the interpretation of the Early Tertiary
stretching event. In the model with no underplating low stretching (p =1.2) is required to
match subsidence. If underplating is considered, increased stretching is required, with p -
factors up to 1.5.
The magmatic underplating has two main effects. One is a short lived (less then 5my)
increase in heat flow by 100 mW/m2 related to dissipation of heat. The other is a longer term
effect associated with increased Early Tertiary stretching. The combined effect results in an
increased temperature of 40°C for the Upper Jurassic source rock interval over the GjaHar
Ridge. Maturity effects of the magmatic underplating is significant, partiCUlarly because the

131
132 FJELD SKAAR, JOHANSEN, DODD, AND THOMPSON

heat pulse give a kick in the generation 5-10 mill. years earlier, which could be important for
oil and gas available to traps formed in late Cretaceous-Paleocene.
Models were calibrated to the observed present-day crustal thicknesses derived from
seismic refraction profiles. The model with underplating calibrates best to these data.

INTRODUCTION

The V"ring Basin is a large sedimentary basin province with grabens, subbasins, and
structural highs. The basin is terminated to the west by the V"ring Escarpment (Fig. 1), the
eastern boundary of the V"ring Marginal High. The formation of this high, during early
Tertiary breakup, was associated with massive extrusive and intrusive magmatic activity. Sills
and dikes are observed mostly as intrusives in Mesozoic sediments over a wide area (100-200
km) landward of the continent-ocean boundary. A lower crustal velocity body has been
mapped at the base of the crust beneath the outer margin, a body that is regarded as a magmatic
underplating, a part of the total amount of melt generated across the rift during breakup
(Skogseid, 1994).

6'0'


Figure 1. Map showing location of modeled cross section (solid line offshore mid Norway). Black areas show
early Tertiary igneous rocks, and dark-shaded areas shows sill and flow complexes (from Skogseid, Pedersen,
Eldholm, and Larsen, 1992).

This paper reports a quantification of the effect of magmatic underplating on the


paleotemperature regime in the Gjallar Ridge, mid-Norway. Magmatic underplating is
important because of the potential of affecting the thermal regime of the sedimentary basin.
The most obvious effect is increased heat flow associated with the injection of molten rock.
The underplating, however, will affect the basin subsidence and thereby the reconstructed
paleo heat flow of the basin, by causing regional uplift by decreasing the density of basement
materials.
MAGMATIC UNDERPLATING IN THE GJALLAR RIDGE 133

The purpose of this study is to analyze and model basin dynamics and factors controlling
basin formation and palaeo temperature regime. The study has focused on the deep structure
of the Y",ring Basin, and the implication for the temperature regime and maturation history if
the observed "low velocity layer" on seismic refraction data really is a magmatic underplating.

BASIN DEYELOPMENT

The seismic section that was modeled is based on deep seismic refraction data
interpreted by Skogseid, Pederson and Larsen (1992). The section extends from the middle of
the Yigrid Syncline to the eastern edge of the Y",ring Escarpment. The geological section used
for the BMTTM (Basin Modelling Toolbox; from Geologica a.s; Lander and others, 1994)
modeling also includes the crustal structure over the Gjallar Ridge (Fig. 2). The section shows:

Figure 2. Cross section of present-day geology showing faults, depositional surfaces, underplated body, and
assumed crustal thickness (partly based on Skogseid, Pedersen, and Larsen, 1992).

• A postulated underplated magmatic body at 15 to 20 km depth, extending from the


Y",ring Escarpment to the axis of the Yigrid Syncline. Seismic refraction data show
low velocities and densities in this anomalous zone. The magmatic body is postulated
to have been emplaced during the Paleocene, based on the age of the basalts drilled
on the Y",ring Marginal High.
• Marked topography on the top of this low-velocity layer; and
• Extensive Late Cretaceous to Early Tertiary faulting over the Gjallar Ridge.
The cross section was reconstructed through geological time by decompaction of
sediments (backstripping) and restoration of faults. The structural restoration was conducted
using simple vertical shear for the normal fault segments. Additional information on BMT's
technique for automatic restoration offaults is given in Lander and others (1994).
Important input data defining the reconstructed geological section included: (1) porosity
depth trends for each lithology for use in sediment decompaction; (2) definition of the
magnitude and distribution of any erosional event; and (3) definition of paleowater-depths.
Lithology properties are presented in Table 1.
134 FJELD SKAAR, JOHANSEN, DODD, AND THOMPSON

Table l. Lithological parameters used in modeling.

Units Age eSZ[2§1~lgSb 1[IDsll Densltl! CQlldUI<lillilll kll CQlldu!:Iillilll kll i:teal ~aQaQit)l
EQ I< lsa/mJ t1QrQ12itll Q 1J t1QrQ12illl Q,Q JLkg ~
Late Tertiary 0-29 0.55 0.62 2700 2.1 1.2 1090
Oligocene 29-50 0.55 0.62 2700 2.1 1.2 1090
Eocene 50-56 0.60 0.26 2700 2.1 1.2 1090
Paleocene 56-65 0.60 0.23 2700 2.1 1.2 1090
Late Cretaceous 65-95 0.60 0.23 2700 2.1 1.2 1090
Early Cretaceous 95-140 0.53 0.23 2720 2.1 1.2 1090
Jurassic 140-208 0.53 0.23 2680 2.1 1.2 1090
Permo-Triassic 208-290 0.53 0.23 2680 2.1 1.2 1090
Basement 0.04 0.05 2800 3.1 3.1 1090
Intrusion 0.02 0.05 2650 3.1 3.1 1090
Asthenose!:!ere 3300 3.5 3.5 1100

(1) Porosity-depth. -- Porosity/depth functions (Sclater and Christie, 1980) affect


modeling results in three ways, they affect: (1) reconstructed thicknesses which in tum affect
the reconstructed geometry of the cross section: (2) reconstructed temperatures by changing
thermal conductivity values and burial depths; and (3) isostatic subsidence simulations because
they are linked to the modeled sedimentary load through time. The porosity/depth parameters
used in this study are shown in Table I.
(2) Erosion. -- The crests of several tilted fault blocks caused by Late Cretaceous to
Paleocene extensional faulting have been eroded during the Paleocene. The erosion surfaces
are relatively planar, suggesting exposure above sea level and the subsequent development of
wave-cut platforms. This erosional event is assumed to be limited to the crests of the fault
blocks and not to be regional in extent. The maximum erosion is assumed to be some several
hundreds of meters (see Fig. 3).

Figure 3. Reconstructed snapshot showing distribution of eroded interval associated with erosional events in
Paleocene time.
MAGMATIC UNDERPLATING IN THE GJALLAR RIDGE 135

(3) Paleo water-depths. -- Paleo water-depths affect model results in three ways: (l) they
affect the geometrical reconstruction of the cross section, (2) they affect the modeled isostasy
resulting from the load of water, and (3) they affect the "observed" tectonic subsidence
associated with rifting events.
Paleo water-depth profiles for the reconstructed timesteps are based on backstripping
studies and on the sequence-stratigraphic interpretation of depositional patterns in the cross
section. Representative curves of the paleo water-depths through time for the cross section are
shown in Figure 4. Several notable features are shown by the plot:
• The water-depth during the Paleocene is close to zero over the Gjallar Ridge. Deep
marine conditions are modeled to the east of Gjallar Ridge (in the Vigrid Syncline).
• Increasing paleo water-depths from the Eocene to the present day.
• The paleo water-depths in pre-Tertary time are uncertain.

,,
500

..... ....
--70M. 1000
--78Ma
- - 95M.
- - - - -140 Ma
........ ·208 Ma

o 20 40 60 80 100 120

Distance (km) A
O~Te~rt~ia~~~_~_~_~_~__~_~_~.r-~~~*,.-~..~
.. _7~~~~~~-t0
a:-:.:- ...

- - - -- - -- - - - - - - ~'~:~-.,=:.-........
I 500 500
~ ----------------------
..,ii
OJ
~
--OMa
~ - - 29Ma
1000
;:1000 - - 50Ma
-----56Ma
·········65Ma

1500 +-...--.. . . .-,-..,.............;~::::;::::::;....'F"'T=;::::;::::;:::;:::..,.-.,--r-.,..,r-r-.-+


120
1500
o 20 40 60 80 100

Distance (km) B
Figure 4. Assumed paleo water-depth change over profile in pre-Tertiary, A; and Tertiary time, B.

Based on data on the porosity-depth relations, paleo water-depth and erosion, the basin
development was reconstructed. An example of one reconstructed time slice (paleocene) is
given in Figure 3. The resulting movements of the top of the basement is. termed "total
observed subsidence." In the following subsidence analysis the total observed subsidence will
be the basis for defining the isostatic and tectonic subsidence, and thus for estimating the
magnitude of stretching over the profile.
136 FJELD SKAAR, JOHANSEN, DODD, AND THOMPSON

SUBSIDENCE ANALYSIS

Subsidence analysis provides the basis for understanding the controls on the
depositional, erosional, and structural evolution of the area as well as for quantifying heat-flow
variations through time. BMT provides forward models of both isostatic and tectonic
subsidence that can be used to generate overall subsidence models for the modeled cross
section. These forward subsidence simulations are constrained to match the subsidence history
generated from the geohistory reconstruction as discussed previously.

Isostatic Subsidence

The isostatic subsidence was calculated using the Airy approximation where the
subsidence is related only to buoyancy effects. Thus lithospheric elastic properties are not
taken into account in the calculations. Lithospheric properties could have important
consequences in the modeling because the elastic lithospheric thickness affects the two-
dimensional isostatic response of the lithosphere to loading and unloading (Fjeldskaar and
others, 1993). Therefore, there is a possibility that the stretching factors are underestimated.
However, sensitivity analyses not presented here show that the effect was not significant in this
study.
The density of the rock intervals must be defined in order to calculate the total load of
sediments used as input for the isostatic model. The values used for the units are (given in
Table 1) are partly based on standard values from the literature. The values for the thermal
conductivity (for 13% and 50% porosity, respectively) are based on inhouse measurements
from the area.

Tectonic Subsidence

Tectonic subsidence is modeled with standard McKenzie (1978) extension, where


crustal extension is equal to subcrustal extension, and based on Airy isostasy. The magnitude
of the stretching was determined empirically by matching the combined subsidence resulting
from the isostatic and tectonic modeling with the "observed" subsidence derived from the
geohistory reconstruction (Fig. 5).
The seismic section shows extensive Late Cretaceous to Paleocene extensional faulting.
This is the precursor to the eventual separation of Norway from North America/Greenland, and
ocean-floor spreading during the earliest Eocene. The thermal subsidence associated with these
extensional events results in the subsidence pattern observed, with the development of a deep
water starved basing throughout Tertiary. This has important heat-flow consequences.

Stretching Factors

To obtain a good correspondence between the "observed" subsidence and that calculated
in the isostatic and tectonic modeling we defined four rifting episodes based on a regional
geological interpretation (Bjmnseth and others, 1997):
• A uniform Permo-Triassic stretching event responsible for the opening of the basin.
This is highly speCUlative because of poor seismic imaging at depth;
• A Middle Jurassic to Early Cretaceous stretching event. This also is speculative
because of poor seismic and stratigraphic resolution in the pre-Cretaceous section.
MAGMATIC UNDERPLATING IN THE GJALLAR RIDGE 137

One rifting episode is used for the Middle to Late Jurassic and latest Jurassic to
earliest Cretaceous rifting episodes;
• A stretching event in the Middle to Late Cretaceous. This can be inferred from
seismic and stratigraphic evidence in the Vruing Basin. It models rifting during the
Early Cenomanian to Late Campanian; and
• A Paleocene rifting episode. This affected the entire North Atlantic margin.
These episodes were modeled as single events at 290 Ma, 140 Ma, 95 Ma, and 65 Ma.

-2 C-,--,---r---r--r ' iii Iii i i r--.----T--.-r--r-T'---T--,--,--"j -.--.-.-r-


t
o f--
2 t--..-.: ; ;: : -----
"Observed" and theoretical subsidence
Gjallar Ridge

E Tectonic subsidence

ioslt
4 f-

6
--- - -- --- .....
-- ~
Isostatic'subsidence ~ - ~ --

~
Total theoretical subsidenc
10

12 lL-~~~~L-~~~~~~~~~~~~~~~~~~~~~~
300 250 200 150 100 50 o
Time (Ma)
Figure 5. Example of observed and theoretical (isostatic + tectonic) subsidence over Gjallar Ridge.

The stretching factors tuned to give a fit were: p= 1.2 for the Permo-Triassic stretching
event, P=1.3 - 1.6 for the Middle Jurassic to Early Cretaceous stretching event, with stretching
factors for the Middle to Late Cretaceous event ranging from 1.0 in the west up to 1.5 in the
easternmost part of the section, and Paleocene stretching factors of 1.1 with an increase up to
1.4 towards east (Fig. 6).
In the calculations so far we have neglected the possible magmatic underplating in the
area. On the basis presented previously, we will study in the following the effects a magmatic
underplating would have on the temperature regime and hydrocarbon maturation.

MAGMATIC UNDERPLATING

Skogseid, Pedersen, and Larsen (1992) have been able to construct a model of the
magmatic features in the Vruing Basin based on velocity-depth curves from seismic refraction
data and gravity data. A large magmatic body is interpreted underplating the continent-ocean
transition region (Fig. 2). The underplated body is here assumed to have been emplaced at 65
Ma at 1300 °C.
Emplacement of a less dense underplated body will result in uplift, as previously
mentioned. Present-day water-depth in the Vruing Basin is of the order of 1400 m over the
section with 1-2 km of post-uplift sediments deposited from Eocene to the present day. It thus
138 FJELDSKAAR, JOHANSEN, DODD, AND THOMPSON

1,7

Stretching factors
1,6
Without underplating - -- 1,6

--
--
1,5 1,5
...0
0
.", ..... -Middle Jurassic to Early Cretaceous /
I'll 1,4 1,4
/ /
Cl

--...
r:: /
.r:: 1,3 1,3
0
<11 _..... /
f
( J)
1,2 ______________ ______________~~~~~_T~~~~_~}_C?__________________________ _L_________ ____ ___ 1,2

~
1,1 1,1

20 30 40 50 60 70 80 90 100

Distance (km)

Figure 6. Magnitude of crustal and subcrustal stretching used to achieve correspondence between subsidence
history implied by geohistory reconstruction and forward models of isostatic and tectonic subsidence.

seems to have been extension in post-Tertiary time that increased the water-depth. This is
supported by extensional faulting of this age over the Gjallar Ridge and the presence of ocean-
floor spreading from the earliest Eocene in the area to the west.
The underplating is treated in BMT as an instantaneous intrusion of hot material. The
additional heat flow associated with underplating is related to two factors; the implied
Paleocene rifting needed to overcome the uplift effects of the intrusive magmatic body, and
the additional heat caused by the hot mass intruded into the Earth's crust (see next section).

Thermal Model Overview

The temperature calculation is divided into two steps. (1) The palaeo heat flow below
the sedimentary cover is calculated based on stretching events and magmatic underplating. (2)
The palaeo temperature regime of the sediments is calculated using the calculated palaeo heat
flow as lower boundary conditions. In both steps a two-dimensional transient finite difference
model is used (see the Appendix). Step I uses a rather low grid resolution whereas step 2 uses
a high resolution grid.

The Underplating Model

In the modeling it is assumed that the entire magmatic body (Fig. 2) is intruded at 65
Ma. The geometries of the magmatic bodies are defined in the present-day geometry of the
modeled cross section. The input parameters for the magmatic body are given in Table 2.
The thermal development in the subsurface down to the asthenosphere is calculated
using a two-dimensional transient finite difference model (see the Appendix). The width of the
modeling domain is defined by the length of the cross section, the depth is defined by the
lithospheric thickness (depth to 1300 °C isotherm) and the palaeo thickness of the sedimentary
MAGMATIC UNDERPLATING IN THE GJALLAR RIDGE 139

Table 2. Intrusion parameters.

Density Pu (Kg/m3): 2650


Thermal conductivity k (W/mK): 3.1
Heat capacity cp (J/kg K): 1090
Latent heat L (J/kg): 420000
Start melting temperature T min caC): 800
End melting temperature Tmax caC): 1100
Reference temperature T ref caC): 600
Initial temperature To caC): 1300
Expansion coefficient a (10-5 °C- I ): 3.3

cover. The modeling domain is divided into three major units, the sediments, the crust, and the
subcrust (Fig. 7). The lateral grid resolution (number of columns) is 18 for all three regions.
The vertical grid resolution (number of rows) differs between the regions. The sediments
(approximately 16 km thick) have a vertical grid resolution of 40 rows, the crust (initially 37
km thick) has a vertical grid resolution of 80 rows, and the subcrust (23 km thick) has a
vertical grid resolution of 40 rows. This grid resolution is sufficient for taking the geometry
and the temperature of the magmatic body into account and to calculate the paleo heat flow.

Surface temperature

Sediments ~~~~~___ Heat flow calculation depth


i=l 100m below bOllom of the basin

Crust

Lithospheric thickness

Sub-crust

Astenospheric temperature

Figure 7. Sketch of tectonic model fInite-difference grid with modeling regions and boundary conditions.
140 FJELDSKAAR, JOHANSEN, DODD, AND THOMPSON

The upper and lower boundary conditions are specified as surface temperature and
asthenospheric temperature, respectively. No flow conditions are defined on the lateral (left
and right) boundaries. The vertical heat flow is calculated along a horizontal line 100 meters
below the bottom of the basin at each time step (Fig. 7). In this situation the heat flow is
calculated at a depth approximately 16 km below sea level for present day. The vertical paleo
heat flow is determined by using the calculated geothermal gradient and the thermal
conductivity in the finite difference grid at the heat flow calculation depth shown in Figure 7.
Both crustallsubcrustal extension and intrusion of magmatic bodies cause a change of the
thermal regime of the crust and subcrust instantaneously. A stretching event will increases the
temperature gradient compared to the pre-stretching temperature gradient. This will result in
an increased heat flow into the base of the sedimentary cover. After a stretching event the
temperature and heat flow will decrease gradually towards thermal equilibrium. Because of
insulation effect of sediments the equilibrium heat flow below the sediments will be lower
after sedimentation compared to before sedimentation.
The intrusion of a magmatic body into the Earth's crust also will instantaneously change
the thermal regime of the crust and subcrust. Once emplaced into the modeling domain, the
magmatic body will start to cool. The mass above and below the magmatic body will
experience a heat pulse which results in a short lived increase of temperature. The heat pulse
after some time (depending of the depth of the underplating), will rise upward to the depth
where the paleo heat flow is calculated (see Fig. 7) and a short lived increase in heat flow will
be observed.
A magmatic intrusion consists of several minerals with different melting/solidification
temperatures. Therefore, the solidification of a magmatic melt will take place through a
temperature interval, depending on the mineralogical constituents. The model assumes that the
latent heat is liberated equally for this melting/solidification interval. The calculation of the
latent heat is implemented as an increased specific heat for the melting/solidification interval
given by cp ' = c.,+ L/{Tmax -Tmin) where L is latent heat (J/kg), Tmin is start melting temperature
eC), Tmax is end melting temperature eC), and cp is the specific heat.
The uplift effect of the underplating also is modeled. The uplift U caused by a magmatic
body with thickness Z, assuming Airy compensation, is U = Z(l - PulPm)' where Pu is the
density of the underplating body and Pm the density of the mantle. Both densities are functions
oftemperature (p=poI(l - a*(T - Tref)), where P is the actual density, Po is the density at Tref,
the reference temperature, a is the thermal expansion coefficient, and T is the actual
temperature). With a density of the underplating body of 2700 kg/m3, and a density of 3300
kg/m3 for the mantle, the uplift of top basement will be 20% of the thickness of the
underplating. A 5-km thick magmatic underplating thus gives a l-km uplift of the surface. This
uplift however, will be modified during the cooling of the magma because of thermal
contraction.
As mentioned earlier, the uplift response will require increased thinning in order to get
a match between observed and theoretical subsidence. The underplated body, therefore, will
not only increase heat flow because of its own heat, but also, indirectly, give increased heat
flow because of the additional stretching needed to match the geohistory subsidence.

Effect on Paleocene Stretching

The model with magmatic underplating gives the same stretching factors as the model
without underplating, except for the Paleocene event. The Paleocene stretching, however, is
affected significantly by the underplating if theoretical subsidence is to be matched with
MAGMATIC UNDERPLATING IN THE GJALLAR RIDGE 141

observed subsidence. The stretching factors are higher (average of 1.5) and show a significant
increase towards west (Fig. 8).

1,7

Stretching factors
1,6 With underplating --- - -- 1,6

--
...0 1,5
...... -- MickUe Ju';ssic to Early Cretaceous
1,5

U
1'0 1,4 / 1,4
Cl /
Paleocene
,/
--...
c:
:cu /

-
1,3 1,3
/
CI)
..... /
en 1,2
• __________ •• _______.___ • __ •• ______ •• ________________ • ______________ • __________ _..l ________.___ • __ . __ _
1,2

._-
Permo Triassic
1,1
,/ 1,1

1
-
-.-Middle to Late Cretaceous ~/
- - - - --.----
20 30 40 50 60 70 80 90 100

Distance (km)

Figure 8. Magnitude of crustal and subcrustal stretching used to achieve correspondence between subsidence
history implied by geohistory reconstruction and forward models of isostatic and tectonic subsidence, based on
underplating option.

Effect on Paleo Heat Flow

The present surface heat flow of the area is assumed to be approximately 62 rnW/m2,
based on figures from Sundvor and Eldholm (1992). Theoretically the heat flow decreases
from a calculated starting value of ca. 80 mW/m2 in Permo-Triassic to the present-day value
of approximately 62 mW/m2. The decrease is caused by sedimentation, because of the
relatively low thermal conductivities of the sediments compared to the thermal conductivity
of the basement.
The modeled tectonic crustal thinning gives high heat flow related to the Middle Jurassic
to Early Cretaceous rifting both on the Gjallar Ridge and in the Vigrid Syncline. The heat pulse
decayed rapidly through the Cretaceous. The shallow depth of burial of any Late Jurassic
source rocks at this time probably meant this increased heat flow had little or no effect on
source rock maturation.
Another pulse of high heat flow was associated with Middle to Late Cretaceous
stretching events. These events, coupled with the rapid sedimentation in the Vigrid Syncline,
resulted in rapid temperature increase and maturation of any Late Jurassic source rocks in the
deep part of syncline, thus expelling all their oil and gas during these events.
The modeled heat flow shows the importance of underplating over the GjaIIar Ridge.
Figure 9 gives the heat flow for the two main modeling options, with and without underplating,
for three locations over the profile. In the model assuming no underplating, only small amounts
of stretching are required for explaining the subsidence over the GjaIIar Ridge (Fig. 9A, 9B),
and consequently there was little increase in heat flow in the Paleocene. In the underplated
142 FJELD SKAAR, JOHANSEN, DODD, AND THOMPSON

220

200

180 Underplating 200 Underplating

... 160

~
g 140

~ 120

i
·COld" underplating
V·COld" underplating
X 100
100

10 ~
10 j---J -. '~~~~derplating" -....................._-................... ~ -·N~-~~d;;rpiaiiiig··--··-······-·········-····-·-····

·1. ·20 ·1.


Age (Ma)
A Age (Ma)
B
Figure 9. Theoretical paleo heat flow at depth of 16
\an below surface at different locations over profile
.. Underplating
(22.5 \an, Fig. 9A; 60 \an, Fig. 9B, and 97.5 \an, Fig .
9C). Three different examples are no underplating

. (dotted line), with "cold" underplating (stippled line)


and with underplating and latent heat (solid line) .
Difference between stippled line and solid line gives
7. "Cold" underplating
direct heating effect of underplated magmatic body.

........................
60

·7. -60 -so --40 ·30 -10 0

Age (Ma) C

model Paleocene heat flows show a pronounced spike associated with the emplacement ofthe
underplated body, superimposed on a longer lived thermal event associated with the rifting.
There is no effect in the axis of the Vigrid Syncline (Fig. 9C) where the underplating body is
modeled as pinched out.
The Tertiary heat-flow peak will be significantly different in the two examples, partly
because of the more pronounced Tertiary stretching for the underplating option. This model
gives a heat flow peak of more than 200 mW/m2, a really significant increase from the "no
underplating" option, which peaks at approximately 70 mW/m2. The effect of the latent heat
is 30-50 mW/m2 (Fig. 10).
A quantification of the direct heat effect also is given in Figure 9, which shows the heat
flow difference for a "cold" vs. "hot" intrusion. The "cold" intrusion model assumes an
underplating temperature similar to the surrounding sediments. The "hot" underplating model
assumes an initial underplating temperature of 1300 0c, The effect is significant, temporarily
increasing the heat flow by ca. 100 mW/m2. The effect is highest were the magma is thickest,
and short lived, it has almost died out after 5 million years.
The two effects of the magmatic underplating in Paleocene time is the direct heating
effect, which contributes to a heat flow effect of up to 100 mW/m2 and increased crustal
thinning, which contributes to a heat flow effect of up to 30-40 mW/m2.
MAGMATIC UNDERPLATING IN THE GJALLAR RIDGE 143

200
I, Underplating, with latent heat I
Underplating. with latent heal

180 ~ 200

Underplating, no latent heat

1.s 160

140
"E
~
.s 150
Underplating, no latent heat

~
~ 0

~
r
120
~r
100
100

80

60
50
·70 ·50 -40 -30 ·'0 ·10 ·70

Age (Mal A Age (Mal 8


100

Figure 10. Theoretical paleo heat flow at depth of 16


90 km below surface at different locations over profile
(22.5 km, Fig. lOA; 60 km, Fig. lOB; and 97.5 km,
Fig. 10C). Two examples are both for underplating
Underplating, with latent heat option, one with latent heat (dotted line), other without

/
latent heat (solid line) .

...--""
Underplating, no latent heat
60

Age (Mal c
Temperature Effects

When the paleo heat flow was calculated, the temperature history of the sediments could
be calculated using a high-resolution two-dimensional transient finite-difference model. In our
modeling a grid resolution of approximately 250 x 250 grid cells (with highest resolution in
geometrically complex areas) is used. The surface temperature (upper boundary condition) is
kept constant at 7°e. The thermal properties of the sediments are shown in Table 1.
Figure 11 shows the maximum total temperature effect of the underplated body at three
time steps, 64.5 Ma, 62.5 Ma, and 56 Ma. The temperature is calculated for the entire profile,
but only the sediment temperature is displayed. 500,000 years after the underplating took
place, the temperature at the deepest point of the sediments is raised by about 50°C (Fig. IlA).
Maximum temperature difference occurs 3.5 my after the intrusion, the temperature difference
at the deepest point now is close to 100°C, and the temperature anomaly reaches closer to the
surface (Fig. lIB). This is the time close to maximum effect of the hot magma. The increased
thinning, however, gives a longer lasting temperature anomaly, and at 56 Ma (Fig. llC) it is
a pronounced temperature anomaly in the basin, close to 80°C at top basement.
The temperature effects also are shown for four different depths in the pseudo well
(location, Fig. 2). The maximum temperature difference at 4 km depth below ocean bottom is
ca. 40°C (Fig. 12).
144 FJELDSKAAR, JOHANSEN, DODD, AND THOMPSON

'J

lDJ

"J

uea Km

L Temperature

• woo u)O
. 8!tte 1000

• 1~0II '~-GC
• t1K .Cl~
it: .~c.;- ,~~
,,!OO iJ~M

,=- 00 MI (10
1:3.Cf: 2500<:1
noo

I. :: :::
)0 00

• '000 , ~oo
• '.00 Soooo
• )0 co nOO
• S~)G tol)l

---
..- - - -.....---...,..-;.--------'u . - - t40g---"'i"ii"a - r t o · o K m

B
Transformation Ratio

The effect on the hydrocarbon maturation is demonstrated by the transformation ratio,


which is the ratio of the organic matter that is transformed to hydrocarbon (Kerogen type II is
assumed; Fig. 13). It is interesting to note that the present-day maturity in the pseudo well is
MAGMATIC UNDERPLATING IN THE GJALLAR RIDGE 145

.,
Km .

n.

~oa- faa 0 ilu Km

c
Figure II. Temperature differences at 64.5 (Fig. IIA), 62.5 (Fig. liB) and 56 Ma (Fig. IIC) over profile,
between models with and without underplating. Plots show combined effect of hot magmatic body and increased
Paleocene thinning.

Paleo temperat ure difference - underplating vs. no underplating

- - underplating 2'600 m
180
- - underplating 3400 m

- - underplating 4200 m
160
- - unaerpla ting 5 100 m

• • • • • •. no underplaong 2600 m
140
.. ....... ... no vnderplating 3400 m

. ...... ' no underplallng 4200 m


120
......... . no underplabng 5100 m

£'00
E
""
80

60

40

20

- 120 - 110 - 100 - 90 -80 - 70 -60 -50 -40 -30 -20 - 10


Age IMa)

Figure 12. Total temperature difference for different depths in pseudo well (see Fig. 2 for location) based on
palaeo heat flow for options with (solid lines) and without (dotted lines) underplating.
146 FJELD SKAAR, JOHANSEN, DODD, AND THOMPSON

the same, whether there was an underplating or not. Significant effect on the hydrocarbon
maturity is predicted only in a narrow zone in the pseudo well, close to 4 km depth.

Transformation Ratio - underplating vs. no underplating

1.0 r---------;::::;::;::===============l
!
- - underplating 2600 m

- - underplating 3400 m
0.9
- - underplating 4200 m

0.8
. - - underplating 5100 m

no underplabng 2600 m
i
• .... .. oo

0.7 • ••• • •• no underplating 3400 m


I
• • • • • •• no underplating 4200 m
0.6 , ••••••• no underplating 5100 m
2
~ 0.5

0.4

-120 -110 -100 -90 -80 -70 -60 -SO -40 -30 -20 -10 0
Age (Ma)

Figure 13. Calculated transfonnation ratio for different depths in pseudo well (see Fig. 2 for location) based on
paleo heat flow for options with (solid lines) and without underplating (dotted lines).

At this depth, the difference is occurring at 65 Ma, effected by the heat pulse of the hot
magma. Even if the temperature effect is more pronounced deeper in the basin, there is no
effect on the transformation ratio, because the temperature was above the oil window when the
underplating took place. On more shallow location the temperature is below the oil window,
and the effect of the magmatic underplating is not enough to bring the temperature above the
oil window.
The maturity difference at about 4 km depth, however, could be of significant
importance. The underplating produces a pronounced kick in the maturation 5-10 million years
earlier than the option without underplating, which could be important for the oil and gas
available to trap formed in Late Cretaceous-Paleocene. It could have marked effect on
predictions of oil and gas accumulations.

Crustal Thinning

The predicted present crustal thickness by the cumulative thinning (assuming an original
thickness of 37 km) differs between the modeling options (Fig. 14). The model without
underplating gives a poor fit.
The presence of significant faulting over the Gjallar Ridge and the nearby Tertiary
oceanic crust implies that Paleocene stretching has been significant. The high stretching factors
implied in the underplated models provide better fits to these observations.
MAGMATIC UNDERPLATING IN THE GJALLAR RIDGE 147

Crustal thickness
10 10

... . . .---
~-
... _---- .. "observed" ~~- ... -.
....... -'-.. -~ ...
... ".
---~--~-~-~-~--::--:-:-:-:-~----------------..-~~-~~---~
E
......
~
...
III
III 15 Prediction, with underplating / 15
Q)
C I
~
0
:c ~/
I-
20 -''-- ---- --- -- -----
Prediction, no underplating 20

20 30 40 50 60 70 80 90 100

Distance (km)

Figure 14. Observed and predicted present crustal thickness (assuming initial thickness of37 kIn) for modeling
options.

CONCLUSIONS

This study has shown that underplating has two important effects: (1) Additional heat
resulting from intrusion of a hot body; and (2) uplift that is the result of emplacement by a light
body, which require additional stretching to explain basin subsidence.
The following conclusion can be drawn from these results:
There was a significant heat-flow effect associated with the Paleocene stretching,
which increased heat flow by 30-40 mW/m2 •
The additional effect of the heat associated with the underplated body itself was
short lived, but pronounced, up to 100 mW/m2 increased heat flow.
The combined effect results in an increased temperature of 40°C for the Upper Jurassic
source rock interval over the Gjallar Ridge.
Maturity effects of the magmatic underplating is significant, particularly because the
heat pulse caused by the underplating gives a kick in the generation 5-10 million years earlier
than without underplating, which could be important for oil and gas available to traps formed
in Late Cretaceous-Paleocene.
We have done sensitivity studies on flexure, paleo bathymetry, and thickness and depth
to intrusion. The story presented here gives one single solution, but it shows the most likely
difference between the options, underplating or no underplating.

ACKNOWLEDGMENT

We thank BP Norge UA for support and permission to present and publish this work.
We also thank two anonymous referees for constructive comments on an earlier version of this
paper.
148 FJELD SKAAR, JOHANSEN, DODD, AND THOMPSON

REFERENCES

Appleyard, R. N., and Cheshire, I. M., 1983, Nested factorization: presented at Seventh Symposium on Reservoir
Simulation, SPE 12264, unpaginated.
Bjmnseth, H.M., Grant, S., Hansen, E.K., Hossack, J.R., Roberts, D.G., and Thompson, M., 1997, Structural
evolution of the Vering Basin, Norway, during the Late Cretaceous and Paleogene: Jour Geo!. Soc.
London,v.154,pt3,p.559-563.
Fjeldskaar, W., Prestholm, E., Guargena, c., and Gravdal, N., 1993, Isostatic and tectonic development of the
Egersund Basin, in Don:, T. and others, eds, Basin Modelling: Advances and Applications: Elsevier,
Amsterdam, p. 549-562.
Hageman, L.A., and Young, D.M., 1981, Applied iterative methods: Academic Press, New York, 386 p.
Lander, R. H., Langfeld, M., Bonnell, L., and Fjeldskaar, W., 1994, BMT User's Guide, version 3.2: Geologica
proprietary publication.
McKenzie, D., 1978, Some remarks on the development of sedimentary basins: Earth Planet. Sci. Letters, v. 40,
no. 1, p. 25-32.
Sc1ater 1.G., and Christie, P.A.F., 1980, Continental stretching: an explanation of the post-mid-Cretaceous
subsidence of the cenrtal North Sea basin: Jour. Geophys. Res., v. 85, no. B7, p. 3711-3739.
Skogseid, J., 1994, Dimensions of the Late Cretaceous-Paleocene Northeast Atlantic rift derived from Cenozoic
subsidence: Tectonophysics, v. 240, no. 1-4, p. 225-247.
Skogseid, J., Pedersen, T., and Larsen, V.B., 1992, Vering Basin: subsidence and tectonic evolution, in Larsen,
V.B., and others, eds., Structural and Tectonic Modelling and its Application to Petroleum Geology: NPF
Spec. Pub!. 1, Elsevier, Amsterdam, p. 55-82.
Skogseid, 1., Pedersen, T., Eldholm, 0., and Larsen, B.T., 1992, Tectonism and magmatism during NE Atlantic
continental breakup: The Vering Margin, in Sorey, B.C., and others, eds., Magmatism and the Causes
of Continental Break-up: Geo!. Soc. Spec. Pub!. No. 68, p. 305-320.
Sundvor, E., and Eldholm, 0., 1992, Norway: off-shore and Northeast Atlantic, in Hurtig, E., and others, eds.,
Geothermal Atlas of Europe, Explanatory Text: Herman Haack Verlag, Geo. Forsh. Zentr., Potsdam,
Pub!. No.1, p. 63-66.
Thomas, G.W., 1982. Principles of hydrocarbon reservoir simulation: IHRDC, Boston, 207 p.
Young, D.M., and Jea, K.C., 1980-81. Generalized conjugate acceleration of iterative methods part 2: The
nonsynnnetrizable case. Center For Numerical Analyses, The University of Texas at Austin, CNA-163,
255 p.
MAGMATIC UNDERPLATING IN THE GJALLAR RIDGE 149

APPENDIX

Temperature Calculations

The following equation is discretized:


a ara
-Kh-+-Kv-=-(cT)
ara
ax ax (k (k at
where T is the temperature, Kh is the horizontal conductivity, Kv is the vertical
conductivity, and c is the heat capacity. Finite differences and a block-centered grid are

used. In the block with indices (ij) the expression ~ Kh dT is evaluated by the following
equation (Thomas, 1982): ax ax
[~Kh ar]ii = -'!"[Kh. J.(2(T T;,j»)_ Kh. } .(2(T,J -T;-Jj»)~l
i+},j -
ax ax Ox, '~J Ox, + Ox'+1 'z') Ox/-l - + Ox, 1

Kh. } . is the value ofKh at the boundary between the blocks (iJ) and (i+ 1J). It is
,~,}

a dT
computed as the harmonic mean ofKhiJ and KhH1J. The expression (k Kva;' is treated
analogously.

This gives M x N equations to determine the TiJ' unknowns, where i = 1,2, ... , M and j =
1,2, ... , N. Here M and N are the number of blocks in x-direction and z-direction,
respectively.

We use both Dirichlet and Neumann boundary conditions for the temperature model. For
Dirichlet boundary conditions the temperature, T, at the boundary is given whereas

for Neumann conditions the heat flux Kh ar and Kv ar , is given. A Neumann boundary
ax (k
condition with a heat flux of zero is used for the basin edges.

An iterative method is used to solve the linear system. Conjugate gradients are used
as an acceleration method (Hageman and Young, 1981; Young and Jea, 1980-1981). The
conjugate gradient method is preconditioned by nested factorization (Appleyard and
Cheshire, 1983). Nested factorization was originally developed to solve the linear systems
that arise in oil-reservoir simulation. According to our experience it perfonns better on heat
conduction problems than on reservoir simulation problems.
COMBINING TECTONICS AND THERMAL FIELDS
IN TARANAKI BASIN, NEW ZEALAND

Phillip A. Annstrong and David S. Chapman

Department of Geology and Geophysics


University of Utah
Salt Lake City, Utah, U.S.A.

ABSTRACT

The Taranaki Basin, New Zealand, is used to illustrate the relationships between tectonics
and thermal fields in an active sedimentary basin. The Cenozoic basin lies 400 km west of the
Pacific-Australian subduction plate boundary and 200 km above the subducting Pacific plate.
The western half of the basin (Western Platform) has not been affected by Miocene to Recent
plate boundary deformation whereas the eastern half (Eastern Mobile Belt) has experienced
at various times rapid burial, uplift and denudation, structural inversion, and igneous intrusion.
This regional separation permits calibration of the various tectonic and thermal interplays, a
feature not afforded most tectonically active sedimentary basins.
Present-day surface heat-flow values average 57 mWm·2 in both the western and eastern
regions. Surface heat flow is remarkably constant in the Western Platform, but varies by more
than 20 m Wm- 2 in the Eastern Mobile Belt. Much of this surface heat-flow variation is
reconciled with detailed basin analysis and thermal modeling that accounts for burial and
exhumation, both of which locally exceed 2 km in time spans of a few million years. The
largest lateral surface heat-flow variation in the Eastern Mobile Belt is a south to north increase
from 50 to 75 mWm-2 for a lateral distance ofless than 50 km. The high heat flow is spatially
coincident with the onshore Taranaki volcanoes suggesting a causal relationship. Apatite
fission track and vitrinite reflectance data and modeling, in combination with detailed basin
analysis, show that the heat flow has been elevated for <1 My and probably is associated with
upper crustal intrusion.
Even though we can account for most ofthe obvious anomalies in the surface heat-flow
distribution, hidden thermal anomalies exist that are not noticeable at the surface. After
accounting for the obvious burial, exhumation, igneous, and thrust-related transient thermal
effects, analysis of modeled basal heat flow (applied at 40 km depth) reveals significant
anomalies in the thermal budget. Basal heat flow varies by more than a factor of two between
wells in the Eastern Mobile Belt and Western Platform regions, even where surface heat flow
is the same. Our modeling shows that a combination oflow heat-producing crust and the heat

151
152 ARMSTRONG AND CHAPMAN

sink effects of crustal thickening below the basin in the Eastern Mobile Belt can account for
the thermal budget anomalies, which is consistent with gravity and magnetic data. A
significant aspect of this study is that some thermal anomalies may be revealed only after
combining geothermics and detailed basin analysis.

INTRODUCTION

Sedimentary basin development and the maturation of hydrocarbons within basin


sediments are both inexorably linked to the local thermal state of Earth's crust. But crustal
thermal regimes, in turn, are modified by the existence of a basin and by tectonic events that
form and modify basins. By tectonic events we infer vertical movements of the crust and
associated exhumation and burial, crustal thinning or thickening, mantle dynamics that cause
variations in subcrustal heat input, and crustal addition because of magma intrusion. Modem
basin-analysis techniques that model thermal histories of sedimentary basins (for summary see
Allen and Allen, 1990; Hermanrud, 1993) provide the vital connection between tectonics and
thermal fields.
Basin analysis is most effective in coupling tectonic and thermal regimes when the
following information is available: (1) the present-day thermal state of the region; (2} depth
and spatial dependence of thermo-physical parameters such as sediment porosity, thermal
conductivity, and radiogenic heat production; (3) detailed burial and exhumation histories and
how they differ within the region; and (4) thermal history indicators such as vitrinite
reflectance and fission track data that verify the appropriateness of a particular thermal history.
The results of basin analysis are most dramatic for basins that record active tectonic regimes
(preferably recent tectonism, so that thermal effects have not dissipated) and basins that
provide for internal calibration of models.
The Taranaki Basin in New Zealand (Fig. 1) is an ideal candidate for such basin analysis:
(1) it is an active plate boundary basin with demonstrated vertical crustal movement that
contributes large transient thermal signals; (2) it has been explored extensively for
hydrocarbons thus providing extensive data sets that allow thermal analysis; and (3) half of the
basin has not been significantly affected by plate-boundary deformation thus providing a
calibration region. In this paper we synthesize several papers related to the various aspects
of the Taranaki Basin tectono-thermal evolution: (Allis, Armstrong, and Funnell, 1995;
Armstrong and others, 1996, 1997; Funnell and others, 1996; and Armstrong and Chapman,
1998).

GEOLOGIC BACKGROUND

Taranaki Basin spans onshore and offshore territory of the western margin of the North
Island of New Zealand (Fig. 1). It is an active-margin basin located about 400 km west of the
Hikurangi Trench and 200 km above the subducted Pacific plate. The structural, tectonic, and
stratigraphic history of the region since the Late Cretaceous has been described extensively
(e.g., Pilaar and Wakefield, 1978; Knox, 1982; King and Thrasher, 1992, 1996; Holt and Stem,
1994) and we briefly review it here. An important aspect of the Taranaki region is its division
into two regions having distinctly different Neogene tectonic histories. The Western Platform
region (Fig. 1) has been relatively quiescent since the late Paleocene whereas the Eastern
COMBINING TECTONICS AND THERMAL FIELDS 153

Figure 1. Location map for Taranaki Basin showing distribution of wells (dots), present-day surface heat flow,
and major faults (teeth on thrust faults). Small numbers next to black dots are present-day surface heat flow (in
mW m-~ (Funnell and others, 1996). Contours are surface heat flow in 5 mW m-2 increments. Stippled region
is Western Platform (WP). Triangles show location ofQuatemary volcanoes. Inset shows location of Taranaki
Basin in New Zealand. Thick dashed lines outline region of plate-boundary deformation. Note that half of basin
is in plate boundary zone and half (Westem Platform) is outside this zone. EMB = Eastern Mobile Belt.

Mobile Belt has been variably affected by Miocene to recent plate boundary defonnation
resulting in significant and variable exhumation in the area.
The pre-Late Cretaceous geologic history of the Taranaki region is important in
establishing initial thennal conditions for the region prior to Taranaki Basin sedimentation.
In the Mesozoic, prior to initial rifting and sedimentation, the region (part of eastern
Gondwana) experienced terrane collision and regional crustal thickening (e.g., Kamp and
Hegarty, 1989; Tulloch and Kimbrough, 1989; Tulloch and Palmer, 1990; Richard and others,
1994). At the time of maximum crustal thickening, the region was intruded by crustal melts
154 ARMSTRONG AND CHAPMAN

and began a period of extensional collapse to fonn metamorphic core complexes that resulted
in extensive, and at times rapid, exhumation of the intrusions during possible footwall isostatic
uplift. Exhumation continued until the end of the Cretaceous when New Zealand rifted from
Australia to fonn the Tasman Sea.
First deposition of sediments in the Taranaki Basin proper began in the Late Cretaceous
(-75 Ma). Subsidence waned in the Paleocene and a relatively quiescent thennal subsidence
phase began that ended abruptly about 25-35 Ma in response to the development of the
convergent plate boundary to the east. In the earliest Miocene, contraction caused westward
overthrusting in the eastern part of the basin. About 10 Ma, the locus of contraction shifted
southward to the Southern Inversion Zone of King (1994), resulting in extensive exhumation.
During Plio-Pleistocene time, with the exception of the areas to the far south, the Taranaki
region generally was undergoing extension that probably was related to back-arc rifting. The
Western Platfonn region has been relatively quiescent and has experienced only subsidence
since the Late Cretaceous.
Sedimentation patterns in the basin have varied both spatially and temporally because of
the numerous tectonic events as described. Late Cretaceous terrestrial sandstones and coals
(Fig. 2) were deposited throughout much of the basin during the initial rifting episode. During
the Paleogene 'drift' phase, dominantly marine sedimentation occurred as the region evolved
as a passive margin. Contraction in the early Miocene brought about a change from carbonate
to terrigenous-dominated sedimentation. Thick sequences of silty mudstone were deposited
during the Miocene (Fig. 2). In Plio-Pleistocene time, uplift and erosion of the Southern Alps
to the south provided abundant terrigenous material deposited as sandstones and mudstones
in the east and dominantly mudstone and siltstone of the Giant Foresets Fonnation in the
Western Platfonn region. Plio-Pleistocene deposits are relatively thick across much of the
basin, locally reaching thicknesses of up to 3000 m (Thrasher and Cahill, 1990).

PRESENT-DAY THERMAL REGIME

Temperatures

The temperature field for the Taranaki Basin is based on bottom-hole temperatures
(BHTs) obtained from geophysical well-log headers and corrected for thennal effects of
drilling. The temperature relaxation model ofLachenbruch and Brewer (1959) was used to
correct BHTs for drilling disturbances. The mathematical representation of the method
(Bullard, 1947) is given by

Teq = T meas. + -Q- [ Ei - -(_r2) - Ei (4 (


_r2) )] ( 1)
41tA. 4at s a ts + tc
where Teq is fonnation equilibrium temperature, Tmeas. is measured BHT at sbut-in time 1:,., tc is
circulation time, Q is line source strength, ). and IX are combined borehole mud and rock
thennal conductivity and diffusivity, respectively, r is the borehole radius, and E j is the
exponential integral. A value of 5.0 x 10-7 m 2 S-1 was used for the system diffusivity as
suggested by Luheshi (1983). Because circulation times are seldom recorded on well log
headers, a value of 5 hours was used to calculate all equilibrium temperatures; errors arising
from this assumption are small because equilibrium temperatures are relatively insensitive to
circulation time (Deming, 1989; Luheshi, 1983). We use Equation (1) rather than its
simplified 'Homer plot' version (Homer, 1951) because it accounts for relatively short shut-in
COMBINING TECTONICS AND THERMAL FIELDS 155

N-NW S-SE
~
.---.---.-----------------~~~~~--~_r_O
1<- Taranaki Peninsula ~ ;l

Ma FORMATIONS / Egmont Volc. ~

• Matemateaonfa

III
10
z V / V v ~
III
U "5- Mohakatino
:;;:
0
~
-~ ;?::
_Manganui
20
'"'-'
Taimana 0
P::
-<
III
Z = Tikorangi f-
-<
III Tikorangi 0
U z
0 30
0
:J Tariki Mbr -<
0
---~
0
;:;;:
• Tangaro; : McKee .'
40

III
Z Mangahewa
III
U
0 50 Z
III
::J
Turi "-
Kaimiro -<
~

°
~z
III
-<Ill
,,-U
60
Tane Mbr
--
Farewell
Turi

N. Cape ::J

-
70 '-<

-
~;?::
I==[j
-<P::
....lU ~~
--. ~~
Rakopi ~-<

BASEMENT

Figure 2. Generalized chronostratigraphic section showing variation from north-northwest to south-southeast


for Taranaki Basin. Thin dashes are mudstone, thick dashes are coal, dots are sandstone, wiggles are marl, brick
patterns are limestone, and v's are volcaniclastics.

times in larger diameter well bores. Using the Homer plot method systematically
underestimates the corrected BHTs where wellbore diameter is large (Funnell and others,
1996). Seventy-nine of 115 Taranaki wells used in this study have at least one common-depth,
mUltiple temperature measurement that can be evaluated with Equation (1). The remaining
wells have only one temperature at any depth and therefore the equilibrium temperature cannot
be computed with Equation (1). A series of diameter-dependent thermal recovery slope versus
depth curves were developed for the Taranaki Basin that allow the single BHTs to be corrected
to equilibrium values (Funnell and others, 1996). The corrections are similar to the depth and
shut-in time dependent corrections of single BHTs given by Willett and Chapman (1987) and
Deming and Chapman (1988). The estimated equilibrium temperature uncertainties are 5-
lOoC for multiple BHTs at a common depth and 10-15°C for single BHTs.
156 ARMSTRONG AND CHAPMAN

Corrected present-day temperatures estimated from our BHT analysis range from 35-
50°C at 1 km to about 150°C at 5 km depth (Fig. 3). At any given depth below about 2 km
the range of temperatures is about 35°C (Fig. 3), considerably greater than the corrected
measurement error of -lOoC. We test the hypothesis that significant lateral variations exist
in the Taranaki Basin thermal field by plotting corrected BHTs by subregion (Fig. 3). The
range (10-15°C) of corrected BHT data in anyone of the six regions at depths between 2 and
5 km is much less than the spread of all the data. Some systematic bias in subsurface
temperatures may be introduced by differences in drilling practice at different times and
locations in the basin, but the overall scatter at least partly represents geographic variations in
temperature gradients. Even though the present-day surface temperature is 13-14°C, the
surface temperature has increased about 4°C in the last 14,000 years (Hornibrook, 1992).
Thus, a more representative surface temperature for calculating conductive geotherms is lOoe
(Funnell and others, 1996). The average present-day thermal gradient, assuming a surface
temperature of lOoC and a constant gradient with depth, is 29°C km· l ; local gradients range
from a high value of 33°C km· 1 in the Maui Field to about 22°C km· 1 in the southern onshore
and Kupe regions.

Temperature (Oe)
20 40 60 80 100 120 140 160
0
I:;.

1000 'b~

.D~
2000
IE • . . . . 0
~

8 ...•• 0
°
..•
'--'
..c: 3000
..... ~
0..
0
Q
4000 I:;. Southern Region ••• ~o o.
0
IE
Northern Taranaki
Western Platform ··.. 1 •
5000 •
0
Maui Field
Northern Onshore
••:. °
• Southern Onshore

and Kupe region
6000
Figure 3. Temperature versus depth for six different regions within Taranaki Basin based on corrected bottom-
hole temperatures.

Thermal Conductivity

We made 265 thermal conductivity measurements on washed drill cuttings from eight
wells distributed throughout the basin using the cell method described by Sass, Lachenbruch,
and Munroe (1971). At 20°C, matrix thermal conductivities range from 0.8 Wm·IK- I for a
COMBINING TECTONICS AND THERMAL FIELDS 157

coaly shale to 5.1 Wm-1K 1for clean, quartz sandstone. By estimating the lithologic mix of the
measured samples to the nearest 5% from composite well logs and mud logs for the interval
sampled, we used a geometric mixing formula and performed a least-squares inversion
(Funnell and others, 1996) to compute the average matrix thermal conductivity of six end-
member rock types: sandstone, siltstone, mudstone, limestone or marl, andesitic volcanic
rocks, and coal. The basinwide conductivity values (Table 1) range between 4.0 Wm-1K 1 for
sandstone and 0.2 Wm-1K 1 for coal and integrate the effects of local mineralogical variations
and anisotropy in the fine-grained rocks. In single-well heat-flow analyses, therefore, we
combined the basinwide end-member matrix conductivity values with lithology derived from
composite well logs and mud logs to construct a matrix thermal conductivity profile for each
well (Funnell and others, 1996).

Table 1. End-member matrix thermal conductivities.


Thermal Conductivity (Wm-'K')
Well Sandst. Siltst. Mudst. Limest. Vole. Coal

Ariki-I 4.4 3.2 2.8 2.3


Fresne-I 3.6 2.5 2.9 3.0
Kupe-I 3.6 2.4 2.7 3.6 0.8
Moki-I 4.4 2.4 2.5 2.8
Tane-I 3.7 3.3 2.5 2.7 0.2
Te Kiri-I 4.9 2.8 2.8 2.5 1.8 1.2
Te Ranga-I 3.9 2.2 3.0 1.8 0.4
Toko-I 4.7 2.5 2.7 1.5 0.4

Basinwide ave. 4.0 2.7 2.8 2.4 1.8 0.2


St. Error 0.1 0.1 0.1 0.2

Note: We use independent estimates of conductivity for coal


(Kappel meyer and Haenel, 1974; Herrin and Deming, 1996) because
it is poorly sampled in Taranaki. Measurements made a 20°C.

Porosity-Depth Relations and Exhumation

Porosity estimates are necessary to adjust matrix thermal conductivity values for in situ
conditions and to perform thermal history analyses. Generally Taranaki Basin porosity
relations were derived from density logs in the stable Western Platform region (Fig. 1). Where
density-log porosities are absent, porosities were computed from sonic logs by converting
travel times to porosity via the density log porosity-travel time transformation (Allis and
others, 1997a; Armstrong and others, in press). This method has been applied to many basins
and regions (e.g., Wells, 1990; Issler, 1992; Heasler and Kharitonova, 1996). A key aspect of
this analysis for the Taranaki Basin is that the Western Platform region has undergone only
minor uplift and exhumation, if any, and thus the porosity-depth relation derived for wells in
that area can be used as a standard, undisturbed porosity function. The best-fitting porosity
function for mudstone derived from all sonic-derived porosity data for 12 wells on the Western
Platform is ofthe form where porosity (<I» is in percent and burial depth (z) is in meters. For
158 ARMSTRONG AND CHAPMAN

(2)

mudstone, <1>0 is 50% and D is 2265 m. A second, but less well-constrained porosity-depth
curve, was generated for sandy sections, where <1>0 is 45% and D is 3000 m. Figure 4 shows
the mudstone porosity data for each Western Platform well relative to the standard porosity-
depth curve derived from the average of all the data. The good fit of data from each well to
the average porosity-depth curve indicates little variability in porosity-depth trends across the
approximately 200 kIn wide Western Platform region.

Porosity (%)
o o,-- 10 20 30 40 50 SO 50 so so SO SO SO 50 SO SO SO

,.......1000 :-~
E .
'-'
-5 2000
0.
a)
Q

4000 ~---

Figure 4. Composite plot of sonic-derived mudstone porosity versus depth for 12 Western Platform wells.
Porosity scale is offset in each plot to show data for each welL Solid curve for each well is the standard porosity-
depth curve for Western Platform.

Porosity-depth trends for the tectonically active Eastern Mobile Belt, in contrast, depart
markedly from the tectonically quiescent Western Platform region porosity curve (Fig. 5).
Because lithology is similar across the entire basin, the lower porosity values in Eastern
Mobile Belt wells at equivalent depths cannot be attributed to differential compaction. A
preferred explanation for low porosities is that they were buried to greater depths in the past
and subsequent erosion has displaced compacted rock volumes to shallower depths.
Comparison of standard porosity-depth profiles for areas representing undisturbed sections of
rock to porosity-depth profiles from eroded sections can be used to make quantitative estimates
of exhumation amounts (Magara, 1976).
An analysis of porosity-depth trends for the Eastern Mobile Belt indicates that
exhumation amounts differ markedly across the region (Allis and others, 1997a). Three wells
from different regions of the Eastern Mobile Belt illustrate the regional variability (Fig. 5).
The depth offset of the best-fitting curve for an individual well, assuming the same porosity
exponential decay constant of2265 m, relative to the standard curve gives a minimum estimate
of exhumation magnitude. The net exhumation is the sum ofthe porosity-depth offset and the
depth to the unconformity on which the erosion occurred. Well Fresne-l in the southern part
of the Eastern Mobile Belt (Fig. 1), has the largest porosity-depth offset (2800 m) in the basin.
Well Kaimiro-l located onshore on the northern Taranaki Peninsula is offset 850 m and well
Maui-2 located in the Maui Field adjacent to the Eastern Mobile Belt-Western Platform
boundary (Fig. 1) is offset only 270 m. The uncertainty in offset amount for all 43 wells
analyzed from the Eastern Mobile Belt is about ±20%.
COMBINING TECTONICS AND THERMAL FIELDS 159

Porosity (%) Figure 5. Sonic-derived mudstone porosity


versus depth for three representative wells from
o 10 20 30 40 50 Eastern Mobile Belt. Dashed curve is Western
o Platform standard porosity-depth curve.
Numbers in parentheses are average depth
offsets of porosity data from standard curve for
1000 each well. Curves through data are not best-fit
curves, but are depth offset standard curves.
r--.
S Note parallelism of offset data trends to standard
curve.
:;- 2000
.......
0.. " Kaimiro-I (850 m)
Q) I
03000 I Fresne-I (2800 m)
I
I
I

4000 I

Two regions of the Eastern Mobile Belt are distinct in tenns of variations in erosion
amounts and geologic histories. Exhumation amounts for ten well sites in the southern region
range from about 200 to 3000 m (Fig. 6). Erosion occurred generally on the top of anticlinal
structures associated with thrust faults and therefore should be considered as local in nature.
Ages of rocks above and below the unconfonnity indicate that the erosion occurred between
about 6 and 10 Ma and probably was associated with basin inversion and an increased rate of
convergence across the Alpine Fault system to the south (Walcott, 1987; King and Thrasher,
1996).
Exhumation in the eastern region increases from southwest to northeast (Fig. 6). Ir\ a
swath that extends from the Maui Field eastward to the Kupe and southern Taranaki Peninsula
region (Fig. 1), exhumation amounts are between 100 and 400 m (Fig. 6). To the northeast and
on the Taranaki Peninsula, exhumation steadily increases to a maximum of 1800 m (Fig. 6).
The age of erosion is Pliocene-Pleistocene in the eastern region and is younger than the erosion
in the southern region. The steady southwest to northeast increase in exhumation is consistent
with the regional 2_4 southwestern dip of the onshore strata and contrasts with the more
0

localized and variable nature of exhumation patterns in the southern region.


Variability in the style and magnitude of exhumation in the Eastern Mobile Belt attests
to the tectonic complexity of the Taranaki Basin. This complexity directly affects both the
present-day surface heat-flow pattern and past thenna1 regimes.

Present-Day Surface Heat Flow

Individual present-day surface heat-flow values (Fig. 1) were detennined by combining


temperature, porosity, temperature- and porosity-corrected conductivity, and sediment heat-
production estimates (Funnell and others, 1996). Temperature correction for conductivity is
similar to that of Chapman and Furlong (1992). For each well we detennined the surface heat-
flow value that produced a best fit between predicted temperature and the corrected bottom-
hole temperatures. Temperature at any depth is predicted from an equation for steady-state
heat conduction in a layered half space:

T(z) = To + ±[qi-I' ~zi _ Ai~Z?] (3)


i=1 Ai 2Ai
160 ARMSTRONG AND CHAPMAN

172E 173E

Km ,
50
38 S


0

i
{2

0
~

~
:i2
0 1840
(;j •

0
0
39 S
(£ 39 S

,t 0
J!i ~
f(]
~


0

40 S

Figure 6. Net exhumation patterns in Taranaki Basin. Values are in meters and contours are in 500 m
increments. Heavy dashed lines are boundaries between Western Platform and southern and eastern regions of
Eastern Mobile Belt.

where
(4)

and To is surface temperature, Ai' ilZi' and Ai are thermal conductivity, thickness, and
volumetric heat production, respectively, for the ith depth interval. Heat flow qi-l is the heat
flow at the top of the ith interval and qo is the surface heat flow. Depth intervals and material
properties are determined from well logs and laboratory measurements. Surface heat flow that
results in the best weighted least-squares fit to the BHTs is determined iteratively using
Equations (3) and (4).
Fifty-three heat-flow determinations were made for the Taranaki Basin (Fig. I). The
average surface heat flow of all the values is 57 m W m·2 • Surface heat flow ranges from less
than 50 m W m·2 in the southern Taranaki Peninsula and the Kupe-I region to about 70 m Wm-2
around Fresne-I in the southernmost part of the basin and near New Plymouth on the northern
Taranaki Peninsula.
COMBINING TECTONICS AND THERMAL FIELDS 161

Uncertainties in surface heat-flow values are difficult to assess. The main factors that
contribute to uncertainty are the estimate of thermal conductivity, and errors in BHT estimates
and corrections. Thermal conductivity uncertainties in turn involve laboratory errors in
measuring matrix conductivity, porosity estimates to convert lab measurements into in situ
conditions, temperature corrections (Lee and Deming, 1998), the appropriateness of thermal
conductivity mixing laws, and the effect of thermal conductivity anisotropy (Blackwell and
Steele, 1989a, 1989b; Deming, 1994; Van Wagoner, Armstrong, and Chapman, 1997). An
absolute heat flow uncertainty greater than ±1O mW m,2 therefore is possible. However,
applying different correction schemes for BHTs, different mixing laws, other conductivity
temperature corrections, or conductivity anisotropy would affect all the wells, thus increasing
or decreasing systematically all the surface heat-flow values. The relative uncertainty between
heat-flow sites is believed to be ±5 m W m,2 and the broad patterns of regional heat-flow
variation are considered real.

BASIN ANALYSIS AND THERMAL EVOLUTION OF TARANAKI BASIN

We quantify the thermal evolution of Taranaki Basin using standard methods of basin
analysis (e.g., Issler and Beaumont, 1989; Allen and Allen, 1990; Hermanrud, 1993). The
basin evolution is simulated using a 1-D finite element model, "BASSIM" (modified from
Willett, 1988), that accounts for the conductive thermal effects of variable burial and
exhumation rates, lithologic and depth-dependent thermal properties, sediment compaction,
and instantaneous sill intrusion. Sediments are deposited at decompacted rates (Sclater and
Christie, 1980) that are based on present sediment thicknesses, sediment ages, and a porosity-
depth trend suitable for the rock type. At each time step (typically 0.1 to 0.5 m.y.), rock
thermal properties are adjusted to in situ conditions, and temperatures are recomputed. Prior
to running our full-scale thermal evolution simulations, we evaluated the relative importance
of 1-D versus 2-D thermal regimes. Because most of the major structures and depocenters
trend N-S, we used east-west basin cross sections and a 2-D finite-element code adapted from
Smith and Chapman (1983) to compute potentia12-D steady-state heat-flow effects resulting
from conductivity variations. At maximum reasonable thermal conductivity contrasts between
basement and sediment matrix of 2.5 and 3.4 Wm,iK'i respectively (corrected for depth-
dependent porosity variations), temperatures computed in the 1-D and 2~D examples differ by
less than 10% at the locations of the wells relative to the major faults (wells are typically >5
km from major faults). Although there may be isolated regions where, for the purpose of oil
exploration, 2-D effects are significant, we believe that most of the temperature variation at
the well sites occurs in a 1-D sense and that the increased complexity of 2-D thermal modeling
is not justified in a general, basinwide analysis,
Boundary conditions in our basin thermal models include temperature for the top of the
sedimentary pile and heat flow at a specified depth, both of which change as a function oftime.
Present surface (top of sedimentary pile) temperatures are about 12-14 DC, and have ranged
from 4 DC during deep-water periods to 20 c C for most of the Tertiary. These estimates are
based on oxygen isotope and paleontologic evidence for surface-water temperatures
(Hornibrook, 1992) and bottom-water temperature versus depth estimates for offshore New
Zealand (Ridgway, 1969). This surface temperature variation affects subsurface temperatures
and leads to large transient thermal signals.
We apply a basal heat-flow boundary condition at a depth of 40 km below the base of the
sediments. This depth allows a thermal coupling between the sediments and lithosphere below
162 ARMSTRONG AND CHAPMAN

them; Nielsen and Balling (1990) showed that excluding transient thermal coupling between
basin strata and the underlying lithosphere can lead to inaccurate estimates of heat flow and
temperature structure of the basin.
Basement heat production measurements from well and outcrop samples show that the
heat production at the top of basement is 1.6 ±l.l ",W m-3 (55 samples; Armstrong, 1996). To
simulate decreasing crustal heat production with depth in the crust (Lachenbruch, 1970), we
assign an exponentially decreasing heat production with depth A(z) = A.,exp(-zlD), with
assumed values of A.,=1.6 '"W m-3 and D=11 Ian to a depth of 29 Ian; below 29 Ian the initial
heat production is zero. A 29 Ian thick subbasin crust therefore has an average heat production
of 0.62 ",W m-3, contributing an integrated 18 mW m- 2 to the surface heat flow.
Satisfactory thermal history models must predict the present-day temperatures (corrected
BHTs) in the wells, as well as thermal history indicators such as vitrinite reflectance and
apatite fission track ages (Table 2). During the burial and exhumational history, thermal
effects of lithospheric thickening or thinning, magma intrusion, or changes in basal heat flow
can be superimposed on the basin heat budget.

Table 2. Thermal history model parameters.


IBasal Heat Flow (mWm") no. RMS' Intrusion
Well 120-35 Ma i 35-0 Ma BHTs (0C) Parameters'
WESTERN PLATFORM !
Ariki-I 38 38 9 8.9
Wainui-I 40 40 2 1.4
Taimana-I 39 39 2 8.0
Tane-I 40 40 4 2.6
Witiora-I 40 40 2 7.7
Maui Field 40 40 7 5.1
Kiwa-I 40 40 3 6.4

EASTERN MOBILE BELT


Maui-4 39 36 3 1.7
N. Tasman-I 39 42 3 1.0
Fresne-I 39 24 3 3.0
Motueka-I 39 29 3 5.1
Surville-I 39 26 3 3.1
Kapuni Deep-I 39 3 20 5.1 7500.500.0.03
Kupe-I 39 5 II 3.2
Mokoia-I 39 -2 5 2.3
Waihapa-I 39 18 12 4.6 8000. 500. 0.03
TeKiri-1 39 22 9 5.3 7500.750.0.10
Kaimiro-I 39 20 7 5.8 7500. 750. 0.40
Mckee-I 39 22 10 4.7 7500. 300. 0.40
Turi-I 39 33 8 7.2

, RMS is root-mean-square misfit of model simulation temperatures to bottom-


hole temperatures (BHTs)
, Intrusion parameters are depth to top of intrusion in meters. thickness in meter
and timing of intrusion in m.y .• respectively.

iRMS is root-mean-square misfit of model simulation temperatures to bottom-hole


temperatures (BHTs)
2Intrusion parameters are depth to top of intrusion in mters, thickness in meters, and timing of
intrusion in m.y., respectively.

Site simulations were run for twenty wells; basin and thermal histories for four wells are
used to illustrate regional variations (Fig. 7). These illustration wells are from the Western
COMBINING TECTONICS AND THERMAL FIELDS 163

Platfonn (Tane-l), the southern region of the Eastern Mobile Belt (Fresne-l), and the low and
high-surface heat-flow parts of the eastern region of the Eastern Mobile Belt (Kapuni Deep-l
and Kaimiro-l, respectively). The southern and eastern regions of the Eastern Mobile Belt are
the same geographic regions as those discussed earlier in the porosity section of this paper.

Age (Ma) Temperature (OC) ViI. Ref. (%) Frage. (Ma)


70 60 50 40 30 20 10 0 0 40 80 120 160 0.3 0.6 0.9 0 20 40 60 80 100
0.0 ,,
1.0 , .
"~
.. .
E2.0
,,
:sc. 3.0
-'"

Cl4.0 ~ ~~~I"'''' ... ~....

A
5.0
6.0 ""

Age (Ma) Temperature (0C) ViI.Ref.(%) Fr Age. (M.)

.,
70 60 50 40 30 20 10 0 0 40 80 120 160 0.3 0.6 0.9 0 20 40 60 80 100
0.0
1.0
~ ~ tt"'~
" ~
]20 I :.
:sc. 3.0 ,,
,,
Cl4.0
,,
5.0 ,.
86.0

,
Age (Ma) Temperature (0C) ViI. Ref. (%) Fr Age. (Ma)
70 60 50 40 30 20 10 0 0 50 100 150 0.3 0.6 0.9 0 20 40 60 80 100
0.0
1.0
E2.0
:sc. 3.0
-'"

04.0
.......
5.0
C6.0
Age (M.) ViI. Ref.(%) Fr Age. (Ma)
40 30 20 0.3 0.6 0.9 0 20 40 60 80 100
0.0

E 2.0
1.0

:s 3.0
-'" ~
C.
Cl4.0 'I~'~ ~ ......
........... \
5.0
06.0
Figure 7. Results of burial and thermal history simulations for four regions of Taranaki Basin: A, Tane-l on
Western Platform; B, Fresne-l in southern part of Eastern Mobile Belt; C, Kapuni Deep-l in low heat-flow region
on Taranaki Peninsula; and D, Kaimiro-l in high heat-flow region on northern peninsula (Fig. 1). Simulations
were run from 120 Ma, but only last 75 m.y. are shown. Left-hand plot for each well shows burial history (bold
lines) and thermal history (thin lines). Isotherms are in 20 0 C intervals. Right three plots show measured
(squares) and modeled (heavy dashed and solid curves) temperature, vitrinite reflectance, and apatite fission track
(FT) ages versus depth. Bottom-hole temperature errors are ±10° and vitrinite reflectance and fission track age
error bars are ± one standard deviation. First-order Arrhenius reaction kinetic approach of Burnham and
Sweeney (1989) was used to model vitrinite reflectance. Model fission track ages were computed using
algorithm of Willett (1992). In temperature, vitrinite reflectance, and fission track plots for C and D, dashed
model curves are for recent intrusion heating and solid curves are for no intrusion heating. Note for well
Kaimiro-l that without recent intrusion heating, modeled vitrinite reflectance and fission track ages are too high
and too young, respectively. Thin dashed curves in fission track plots are stratigraphic ages for fission track
samples. All depths are relative to top of rock column.
164 ARMSTRONG AND CHAPMAN

On the Western Platform, initial rapid burial occurred between 75 and 44 Ma during the
rift phase of basin development. Rapid burial was followed by slow sedimentation until about
15 Ma when deposition rate increased. At 4 Ma, rapid deposition recommenced as the Giant
Foresets Formation (Fig. 2) was deposited. Isotherms are elevated in the early part of the
burial history (Fig. 7A) because of two factors. First, rapid deposition of high porosity, low
conductivity sediments on top of higher conductivity basement rocks causes temperatures to
rise in the basement and offsets a transient surface heat-flow decrease that accompanies rapid
influx of cold sediments. Second, the effect of syndepositional rifting imparts a long-lived
thermal signal that particularly affects the deeper horizons. Because we assume the same
initial conditions for the pre-burial lithosphere across the entire basin, the initial rise in the
isotherms is typical of all wells. The variations in early temperature histories depend on
sedimentation rates at each well site.
For all simulations (Fig. 7) there is a sharp depression in the isotherms at 20-28 Ma
because of increased water depth accompanying rapid subsidence and a corresponding
decrease in bottom water temperature. This isotherm depression probably is associated with
platform subsidence during the first stages of convergent margin tectonics between 35 - 25 Ma
(Holt and Stem, 1994).
The southern region experienced the same Late Cretaceous-Paleogene rift-related rapid
sedimentation followed by slow sedimentation similar to the Western Platform. More than 5
km ofterrestrial and coastal plain deposits of ages 75 to 65 Ma are followed by small burial
amounts until about 25 Ma at the Fresne-l site (Fig. 7B). At end of the Miocene and into the
Pliocene as much as 3 km of exhumation occurred on discrete thrust-related structures (Figs.
6 and 7B). One consequence of this exhumation was to raise isotherms closer to the surface
(Fig. 7B) and increase surface heat flow about 25 mW m-2 above a background of 60 m W m-2
in Pliocene time at well Fresne-l. More recent burial since has decreased surface heat flow
back to 70 mW m- 2 (Fig. 1). Present-day surface heat flow at most well sites in the southern
region (Fig. 1) is relatively high, reflecting this late Miocene to early Pliocene erosional event.
Well Kapuni Deep-l (Kapuni Field; Fig. 1), located in the eastern region on the southern
Taranaki Peninsula where present-day surface heat flow is relatively low (Fig. 1), displays the
same early rift and drift phases as discussed earlier (Fig. 7C). The well site experienced late
Miocene-early Pliocene uplift and erosion (King and Thrasher, 1996) followed by rapid
Pliocene burial of up to 2 km and a second exhumation event in Pliocene-Pleistocene time.
These relatively rapid and closely spaced (in time) exhumation and burial events have imparted
a complex thermal history that ends with a present-day surface heat flow of 53 m W m-2•
Well Kaimiro-l is located in the high heat-flow area on the northern Taranaki Peninsula.
In contrast to the other wells previously discussed, this site experienced relatively minor rapid
subsidence in the Late Cretaceous and more rapid burial in the Paleogene (Fig. 7:0). This
region did not experience the late Miocene-early Pliocene erosional event evident in Kapuni
Deep-l to the south, but did experience about 1 km of erosion in Pliocene-Pleistocene (Figs.
6 and 7D). Present-day surface heat flow in the Kaimiro region is >70 mW m-2, which is 15
mW m-2 greater than the average southern peninSUla value.
In summary, combined basic basin analysis and thermal modeling techniques that account
for the thermal effects of burial and exhumation show that significant variations have occurred
in the thermal histories of different regions of Taranaki Basin. Because most of the basin
experienced the same pre-Miocene tectonothermal history, most thermal variations have taken
place in the Neogene coincident with subduction tectonics to the east. In the following
sections we introduce factors other than simple burial and exhumation that alter the Neogene
thermal regime.
COMBINING TECTONICS AND THERMAL FIELDS 165

SUPERPOSITION OF INTRUSION THERMAL EFFECTS

The most significant lateral variation in present-day surface heat flow in the basin is the
40% increase in a south-to-north direction across the Taranaki Peninsula (Fig. 1). This lateral
surface heat-flow gradient in part is spatially coincident with the andesitic Mt. Taranaki
volcano (also referred to as Mt. Egmont), which reaches an elevation 2518 m above sea level,
and the eroded remnants of older Quaternary andesitic volcanoes (Fig. 1). The volcanoes
decrease in age to the southeast from 1.74 Ma near New Plymouth, followed to the south by
remnant volcanoes with ages of 0.57 and 0.25 Ma, and Mt. Taranaki with a maximum age of
0.12 Ma (Neall, 1979).
Three-dimensional gravity modeling results (Locke, Cassidy, and MacDonald, 1993)
show that the feeder dike system below intermediate age volcanic centers extends at least to
basement depths (>5 km). Therefore, any larger magma chambers are deeper than 5 km. Allis,
Armstrong, and Funnell (1995) used simple two-dimensional intrusion models to bound the
range of admissible intrusion depths and thicknesses that could account for the amplitude (~15
mW m-2) and width (25-50 km, but poorly determined) of the high heat-flow anomaly on the
northern peninsula. Their study showed that the anomalously high heat flow could be caused
by either crustal underplating of about 5 km of basaltic material between 2 and 4 Ma, or upper
crustal (10 km depth) sill intrusion of 0.5 km thickness during the past 0.2 to 0.5 m.y., or a
combination of both. Therefore, thermal modeling alone cannot significantly constrain
intrusion timing or depth.
In order to limit the timing of the high heat-flow anomaly, we compared apatite fission
track (AFT) ages and vitrinite reflectance (Ro) values with those measured from samples
collected in selected Taranaki wells (see Armstrong and others [1997] for complete data set,
data evaluation, and analysis). Armstrong and others (1997) show that AFT ages and Ro values
from the high heat-flow region are consistent with relatively low temperatures and heat flow
at the time of maximum burial. Relative to the southern part of the peninsula, the temperatures
must have increased by about 20°C at depths of3-4 km within the last <1 m.y., and perhaps
in the last 0.1 m.y. In addition, this relatively recent heating implies that the intrusion must
have been shallow, probably in the upper 7-8 km. This depth roughly corresponds to the depth
of the basement, where a strength or rheological boundary may have influenced localization
of magma ponding (Roman-Berdiel, Galais, and Brun, 1995; Armstrong and others, 1997).

INTRUSION MODELING

Thermal effects of solidification and conductive cooling within and around horizontal sill
intrusion contemporaneous with burial and exhumation are incorporated into our basin
simulator. We used the formulation of Peacock (1989) with an assumed emplacement
temperature of 1000°C and specific heat of 1000 J kg·toe t in solidified portions of the
intrusion. Specific heat is adjusted to higher values in the liquid part of intrusion to account
for latent heat of crystallization of the magma according to

c*=c+LI~T (5)

where c* is specific heat of magma, c is specific heat of solid intrusion, L is latent heat of
crystallization (400 kJ kil), and ~T is temperature range between liquidus and solidus
(l00°C) (Peacock, 1989; Furlong, Hanson, and Bowers, 1991).
166 ARMSTRONG AND CHAPMAN

We tested the sensitivity of differing intrusion depths and reasonable thicknesses (up to
1 km) assuming an emplacement age of 0.40 Ma for well Kaimiro-l in the high heat-flow
region. This emplacement age is consistent with the ages of nearby volcanic edifices and is
within the heating time of <1 m.y. suggested by the AFT and Ro analysis. Depths of7-8 km
and a thickness of 750 m gave the best fit to temperatures in northern peninsula wells. Deeper
intrusions require a greater thickness and earlier emplacement times to satisfy the temperature
constraints, but result in high temperature transients that are longer lived than 1 m.y. In other
wells that surround Mt. Taranaki on the peninsula, an intrusion depth of 7.5-8.0 km was
assumed and the intrusion thickness and timing was differed to give the best temperature fits.
The depth, thickness, and timing parameters are given in Table 2 for modeled well sites near
known Quaternary volcanoes. The depth range predicted by our modeling is roughly
coincident with the depth of basement in the eastern part of the basin.
The model fits of temperature, Ro, and partial annealing zone AFT ages for well Kaimiro-
1 are good for the situation oflate heating resulting from intrusion (Fig. 7D). Ifno intrusion
is assumed and past temperatures are based on the relatively high heat flow that is present
today, then computed vitrinite reflectance values and fission track ages are too high and too
low, respectively (Fig. 7D).
Well sites in the Kapuni and Waihapa fields (Fig. 1) are directly in line with the volcanic
trend (Kapuni) or adjacent to the youngest volcano (Waihapa) and therefore may have been
affected by intrusion-related heating. The present-day temperature gradient in Kapuni Deep-l
increases below about 4.5 km (Fig. 7C). Well Waihapa-l (Waihapa Field; Fig. 1) does not
penetrate below 4.5 km, but the temperature gradient does increase slightly near the maximum
well depth. This increase in temperature gradient may be the result of transient effect of the
recent intrusions. However, anomalously low in situ thermal conductivity may be contributing
to the increased gradient below 4.5 km. In the Kapuni Deep-l temperature plot (Fig. 7C), the
dashed and solid model curves are for best-fit simulations that include intrusion and no
intrusion, respectively. If shallow intrusion is responsible for the increased gradient, then
intrusion modeling suggests instantaneous intrusion at about 0.03 Ma (Table 2). This intrusion
age is reasonable if the volcanic center is migrating southeastward as indicated by the
orientation of the volcanic chain on the Taranaki Peninsula (Fig. 1) and the age (0.12 Ma) of
Mt Taranaki. Because ofthe relatively recent age of possible intrusion, its thermal effects may
have not yet been sensed at depths less than 4.5 km at Kapuni Deep-l and the present-day heat
flow at the surface yet is relatively low.
Although our solutions are nonunique, as is typical of multiparameter models, the heat
flow, AFT age, and Ro data in the high heat-flow region are consistent with shallow (7-8 km
deep), thin (500-800 m thick), recent «1 Ma) intrusion. Even though surface heat flow is
presently low south of the youngest volcano, the increased temperature gradient in the deepest
well in this region suggests that intrusion heating may be occurring there at present.

HIDDEN THERMAL ANOMALIES - BEYOND SURFACE HEAT FLOW

In our thermal models discussed so far we have accounted for burial, exhumation, and
local magmatic thermal effects. These, however, are only a few of the possible processes that
may contribute to the overall past and present thermal budget. Other nonclimate-related
sources of surface heat-flow variability, especially in active tectonic regions such as Taranaki
Basin, include: (1) variations in mantle heat flow, (2) integrated heat production variations in
the crust, (3) thermal effects offluid flow including depressed heat flow in hydrologic recharge
COMBINING TECTONICS AND THERMAL FIELDS 167

areas and elevated heat flow in discharge areas, and (4) large-scale tectonic modification of the
crust such as crustal thickening and thinning, which respectively depress and elevate heat flow.
In this section we ask the questions: are there hidden thermal anomalies that are not apparent
in the present-day surface heat-flow distribution of Figure 1 and if so, how might they be
reconciled in terms of the possibilities listed?
Surface heat flow ('lsurf) in steady state can be considered as the sum of six components:

qsurf =qbase + qrad + qexhlbur + qfluid + qrnagrna + qtect (6)

The terms on the right are the heat-flow contributions/perturbations: basal heat flow (qbase)'
integrated sediment and basement radiogenic heat production (qra~' exhumation and burial
(qexhlbur)' fluid flow (qflui~' magmatism ('lmagma)' and heat-flow variations caused by crustal
thinning or thickening (qtecJ. Under normal circumstances qbase' qrad' qexh' and qmagma are
positive, qbur is negative, but qfluid and qtect may be either positive or negative. Some of the
terms can be highly transient with different time lags so that at any time the full effect of a
particular term may not be sensed at the surface. As many geodynamics problems focus
primarily on tectonic heat, the challenge is to estimate or bracket all the other terms
sufficiently well to relate with confidence the desired quantity qtect.
Basal heat flow (qbasJ is simply the quantity of heat flowing out of Earth across a plane
at some arbitrary level or depth; if this level were selected as the crust-mantle interface, then
it would be synonymous with mantle heat flow. In this study we applied the qbase condition
at a depth of 40 km below the base of the sediments. A different basal depth changes the
sensitivity to changes in qbase' but does not significantly alter our conclusions.
If the relative magnitudes of the tectonic, fluid flow, variable heat production, and mantle
heat-flow variation effects are unknown, then the apparent qbase can be treated as if it
potentially contains a component of each. We therefore can calculate an apparent qbase subject
to the conditions applied in the model, but it should not be confused with reduced heat flow
that is related to the linear surface heat flow versus heat-production relations (Birch, Roy, and
Decker, 1968; Lachenbruch, 1970) and need not represent the actual mantle heat flow. qbase
as used in this study refers to the apparent heat flow applied at the base of the model needed
to satisfy observational constraints. The qbase is determined through iterative model
simulations, keeping track of known transient effects such as burial and exhumation. Spatial
and temporal variations in qbase then are interpreted with respect to the tectonic, magmatic,
fluid flow, and variable heat production terms.

Application to Taranaki Basin

We have applied the basal heat-flow analysis method to thermal histories of twenty wells
(Fig. 8 and Table 2), but illustrate the variations with one well each from the Western Platform
and the Eastern Mobile Belt.
Well Tane-l has a burial history (Figs. 7A and 9A) that is representative of most Western
Platform wells. A constant qbase of 40 m W m·2 for the entire buriaVexhumation history in
Tane-l (Table 2) satisfied all the thermal constraints (Fig. 7A). The qbase for six other wells
on the Western Platform are remarkably consistent at 38 to 40 m W m·2 (Table 2 and Fig. 8),
which is close to the generally used basal heat flow for undisturbed continental lithosphere
(Roy, Blackwell, and Birch, 1968; Lachenbruch and Sass, 1978). Because of the qbase
consistency for the Western Platform wells, and because they are far from possible
topographically driven fluid-flow effects, recent volcanic activity, and Miocene-Recent plate
168 ARMSTRONG AND CHAPMAN

-----60

5.0

Figure 8. Map showing pattern of modeled basal heat flow. Large numbers next to wells are modeled basal heat
flow as given in Table 2. Solid contours are surface heat flow in 5 mW m·2 increments from Figure 1. Thin
dashed lines are gravity contours in 20 mGal increments (after Rose, 1991). Hatured region is R magnetic
anomaly of Davy (1992). Stippled region is Western Platform. Triangles show locations of volcanoes on
peninsula.

deformation, these qbase values are used as control or background values to which %ase values
from wells within the Eastern Mobile Belt can be compared.
Well Kapuni Deep-l has a burial/exhumation history (Figs. 7C and 9A) similar to many
other Eastern Mobile Belt wells. If a qbase of39 mW m-2 as at Tane-l is used, then the present-
day surface heat flow at Kapuni Deep-l would be 70 mW m-2 (Fig. 9B) and fission track ages
and vitrinite reflectance values would be too low and too high, respectively. The burial and
thermal constraints in Kapuni Deep-l are satisfied with a qbase of 22 m W m-2 for the entire
history (KD model 2; Fig. 9B). However, plate boundary zone deformation associated with
subduction to the east began 25-35 Ma (Walcott, 1987) and prior to this time, there are no
apparent reasons for the qbase values between the Western Platform and Eastern Mobile Belt
to be different because both regions were part of the same passive margin system. Thus the
COMBINING TECTONICS AND THERMAL FIELDS 169

qbase for Kapuni Deep-l was held constant at 39 mW m-2 from the start of the simulations until
35 Ma, then changed and held constant until 0 Ma at a new value that best satisfies the thermal
constraints (KD Model 3)_ Given these conditions, the required qbase for the last 35 My at
Kapuni Deep-l is 3 mW m-2• This Neogene qbase value is considerably lower than the Western
Platform value of 39 mW m-2, even though the present-day surface heat flow is comparable.

A. O+-~~~~~~~~~~~~~~+

2 Tane-l

~
..s0..
4

Q 6 Kapuni Deep-l
8
80 0 60
B 80
--Tane-I
.•••• K_D_ model I
- - -K.D_ model 2
- - K.D. model 3

40~~,,~~~on~~,,~on,,~~h+
80 70 60 50 40 30 20 10 o
Age (Ma)
Figure 9. Burial histories A, and corresponding surface heat flow (1J.un) histories B for Tane-l and Kapuni Deep-
1 simulations. In A, burial history is for top of basement horizon; in B, K.D. (Kapuni Deep-I) model 1 basal heat
flow (<Ib.sJ is held constant at 39 mW m·2 (same as Tane-l). InK.D. model 2, 'lbase is held constant at 22 mW m·2 •
In K.D. model 3, Cb...c is decreased at 35 Ma to 3 mW m-2• Note that K.D. model 1 overestimates Cisurr by 17 mW
m-2 and therefore lower model 'lbase is implied. Variations in Cisurr in early part of history reflect initial conditions
and differences in early burial rates (Armstrong, 1996).

Large variations in qbase when there is a small variation in Qsurr between wells is partially
related to differences in the transient thermal effects of burial and exhumation. The heat-flow
differences between Tane-l and Kapuni Deep-l result from the two exhumation events in
Kapuni Deep-l in the last 10 My, while Tane-l experienced rapid burial (Fig. 9). Burial
decreases heat flow in both wells but exhumation increases it by 11 mW m-2 in Kapuni Deep-I.
The overall effect for the last 10 m.y. is to decrease surface heat flow in Tane-l by 6 mW m-2
and increase it by 4 mW m-2 in Kapuni Deep-l (Fig. 9). The exhumationlburial differences in
the last 10 m.y. therefore accentuate the modeled qbase variations even though present-day
surface heat-flow values are similar between the areas.
The same early qbase history applied to the other eastern Eastern Mobile Belt wells as in
Tane-l and Kapuni Deep-l Model 3 (39 mW m-2 prior to 35 Ma) yields qbase values for the
170 ARMSTRONG AND CHAPMAN

last 35 m.y. between -2 and 5 mW m·2 on the southern peninsula and just offshore, to 33 mW
m o2 in the northernmost Eastern Mobile Belt, with intermediate values in between (Fig. 8). In
the southern Eastern Mobile Belt, qbase values range from 24 to 42 mW m o2 • Whereas the
modeled qbase values on the Western Platform are consistent, those in the Eastern Mobile Belt
are overall lower and more variable, even where present-day surface heat flow is relatively
invariant (Fig. 10).

Distance from trench (Ian)


500 450 400 350 300
80 ~~__~~~~__~__~~-r~__~~80
, B~O
6 ~ 60 o ~~OdfO cPo
t;::N
., 'e
"iii
• 0 o 4:01 El
60

..c:
., ~ 40 0 0 <OcPo~ 0 Cb
1 0 0
u
~
§. 1
1
o 0
.... 20 1
000 0 0
"
en o Surface 1
1
0 o Basal
1
1
1
o o
WESTERN PLATFORM ~ EASTERN MOBILE BELT
1
1
Tane-l ~ Kapuni Deep-l
1
1
1

~100
--+---l----i----+---+---- 1

!..c: 38 to 40 mW m- 2 1 <0 to 40 mW m- 2
.,
0. \ )'
a 200 Apparent basal
heat flow
Subducting plate
Figure 10. Surface and modeled (apparent) basal heat-flow values versus distance from Hikurangi trench in
Figure 1 inset. Bottom is sketch cross section showing Taranaki Basin location relative to subducting Pacific
plate. Filled squares in top plot are present-day surface heat-flow values for Tane-l and Kapuni Deep-l as shown
in sketch (lower). Surface heat-flow values for two wells are similar (56 and 53 mW m o2 , respectively), yet
required subcrustal basal heat values vary greatly.

Interpretations of Basal Heat-Flow Variations

Causes of low apparent basal heat-flow values for the Neogene in the Eastern Mobile
Belt compared to the Western Platform can be divided into two categories: (1) variations in
heat flow at depth below the model boundary that might cause as much as a 10-fold variation
in modeled deep heat flow between the region for the last 35 m.y.; and (2) thermal changes
between the model lower boundary and the top of the sediment pile that lower temperatures
in the well and thus suggest a lower basal heat flow to satisfy the heat budget.
Heat flow from the mantle can vary with space, but tectonothermal mechanisms
responsible for the mantle heat-flow variations are poorly understood. Taranaki Basin is
located above the subducting Pacific Plate (Fig. 10) where the relatively cool subducting plate
should cool the area under the eastern region more than the western region. However,
COMBINING TECTONICS AND THERMAL FIELDS 171

Taranaki Basin is 150-200 km above the subducting plate (Fig. 10). If subduction began 35
Ma, conductive cooling (slab refrigeration) would not be sensed yet at the model base; thermal
length arguments suggest that about 40 My would be required for significant heat-flow
changes.
Holt and Stern (1994) suggested that the broad subsidence patterns in Taranaki Basin
were caused by mantle flow induced by the subducting slab. The mantle flow could
potentially set up a hydrodynamic circulation such that the upper mantle to the west is warmer
than that to the east. However, the required basal heat-flow change of>20 m W m·2 from the
Western Platform to the Eastern Mobile Belt occurs in about 50 km, which is a shorter
distance than suggested by the several hundred kilometer wavelength mantle flow scenario.
Large-scale mantle flow therefore is not considered a viable mechanism for heat-flow changes
near the base of the crust.
If the model qbase anomalies cannot be explained by variations in mantle heat flow, then
they must be explained by changes between the model boundaries that, when not accounted
for, require variations in the qbase' Next we describe thermal property and intra- and
subbasinal processes that could thermally perturb the basin such that qbase could be the same
(39 mW m- 2) in the Western Platform and Eastern Mobile Belt.
The models described previously do not take into account possible fluid-flow effects
within the basin sediments. Mt. Taranaki is a potential source oftopographically driven water
flow onshore. Modeling of the topographically driven fluid flow utilizing Taranaki-derived
relations between lithology-dependent porosity and permeability shows that flow velocities
are low at depths of>0.5 km and distances of>20 km from Mt. Taranaki (Allis and others,
1997b), suggesting small thermal disturbances at typical BHT depths. Also, temperature
profiles from north and south ofMt. Taranaki show that there are no apparent perturbations
that might be associated with groundwater flow in the upper 5 km (Allis, Armstrong, and
Funnell, 1995b). Deep regional fluid flow below 5 km depth may be affecting temperature
gradients farther up-section, but inter-well variations in potential regional deep flow patterns
that might cause the necessary lowering of qbase at one location and not others probably are
small.
A possible explanation for the low apparent qbase is that the subbasin crustal heat
production in the Eastern Mobile Belt may be lower than in the Western Platform. If the total
integrated heat production were lower in the Eastern Mobile Belt, then the required qbase there
would be higher in our simulations. The low qbase region corresponds to an ~25-50 km wide
large-amplitude (200-400 nT) magnetic anomaly that is interpreted to be caused by upper
crustal mafic material (Davy, 1992). This region coincides with a major tectonic boundary
that separates continental-type crust to the west from an assemblage of mafic accretionary
complexes to the east (Bradshaw, 1993). Compilations of heat production versus density show
that heat production decreases with increasing density (e.g., Cermak and others, 1990) so that
if the basement that underlies the Eastern Mobile Belt wells consists of mafic bodies, then its
heat production may be lower than in the Western Platform.
To match the Western Platform average qbase of39 mW m-2 an average basement heat
production of <0.15 ~ W m-3 is required for the Eastern Mobile Belt, which contributes <5 m W
m-2 to the total surface heat flow. This low average heat production would require an
ultramafic composition for the entire subbasin crust in the low qbase region. The low %ase
anomalies, therefore, cannot easily be accounted for solely with low heat production in the
Eastern Mobile Belt relative to the Western Platform.
Finally, another mechanism possibly responsible for much of the qbase variation is crustal
thickening. Stern and Davey (1990) noted a deepening of the reflection Moho under the
172 ARMSTRONG AND CHAPMAN

southern Eastern Mobile Belt and suggested that the crust there is at least 10 Ian thicker than
the Western Platfonn crust. A gravity anomaly difference of about 100 mGal between the
Western Platfonn and Eastern Mobile Belt (Fig. 8) suggests a mass deficit or relatively thick
region oflower density (relative to mantle) material in the Eastern Mobile Belt region. A
relatively dense Eastern Mobile Belt crust (2900 versus 2750 kg m-3), and therefore with
relatively low heat production, has been interpreted based on gravity modeling (Holt and
Stem, 1994). Heat production versus density compilations imply an average heat production
of 0.45 IlW m-3 for rocks with a density of 2900 kg m-3 • With this heat production
contribution, the heat sink effects of crustal thickening must decrease surface heat flow 15-20
mW m-2 • Bulk crustal thickening models (Armstrong, 1996) show that 10-15 Ian thickening
in the last 20 m.y. can decrease surface heat flow enough to offset the exhumation and burial
effects and therefore account for the basal heat flow anomalies. Our crustal thickening
interpretation is consistent with gravity, magnetic, and seismic data. But the anomalies would
have gone unnoticed in the thennal data without the aid of detailed basin analysis and thennal
modeling.

SUMMARY

The Taranaki Basin (New Zealand) is a Cenozoic active-margin basin located adjacent
to the Pacific/Australian plate boundary that owes much of its origin to the evolving plate
interactions in the area. Combining tectonics and thennal fields in the Taranaki Basin is
important because (1) the late Cenozoic tectonic history has produced thennal transients that
are large and can be identified, analyzed, and their possible causes interpreted; and (2) half of
the basin (Western Platfonn) is outside the region disturbed by Miocene to Recent plate
boundary defonnation and half (Eastern Mobile Belt) is within it. The Western Platfonn
thennal regime can be used as a standard against which the Eastern Mobile Belt thennal
regime can be compared for analysis of transient thennal signals.
Analysis of bottom-hole temperatures in 115 wells reveals considerable geographic
variation in the present-day temperature structure. Average temperature gradients range from
<25 °CIkm to 33°CIkm. Temperatures were combined with extensive thennal conductivity,
heat production, and porosity data to detennine present-day surface heat-flow values for the
basin. The average heat flow is about 60 mW m-2, but ranges from 45 to 75 mW m-2, mostly
because of transient thennal effects related to tectonic and igneous influences.
Late Neogene exhumation amounts, based on well-defined porosity-depth trends, differ
significantly across the different regions of the Eastern Mobile Belt. Much of the variation
in surface heat flow is accounted for by the transient thennal effects of this exhumation and
periods of rapid burial. In addition, Quaternary volcanism and possible associated intrusions
also have affected the thennal regime. Analysis of heat flow, apatite fission track age, and
vitrinite reflectance data from a region of anomalously high heat flow on the northern Taranaki
Peninsula suggests that the high heat flow there may be the result of upper crustal igneous
intrusion in the last 1 m.y.
After accounting for the transient thennal effects of burial, exhumation, and possible
intrusion, 1-0 burial and thennal history modeling results reveal anomalies that are not seen
in the present-day heat-flow distribution. Heat flow applied at the 40 Ian deep model base
(apparent basal heat flow) that is needed to satisfy the temperature and burial history
constraints is unifonn across the Western Platfonn. Apparent basal heat flow for the Eastern
Mobile Belt is at least 20 mW m-2 less than for the Western Platfonn, even where surface heat
COMBINING TECTONICS AND THERMAL FIELDS 173

flow is comparable between the regions. Even though we cannot constrain our modeling
uniquely, our favored interpretation, corroborated by gravity, seismic, and magnetic data, is
that a combination oflow heat-producing crust and the heat sink effects of crustal thickening
in the Eastern Mobile Belt relative to the Western Platform region have caused the thermal
anomalies. More work is needed to constrain further the transient nature of the thermal
history, but the anomalies would have gone unnoticed without the aid of detailed basin
analysis and thermal modeling.

ACKNOWLEDGMENTS

This project was funded by the donors of the Petroleum Research Fund administered by
the American Chemical Society. We thatik Rick Allis, Rob Funnell, Peter King, and Glen
Thrasher for helping us understand the geology and geophysics of Taranaki. Peter Kamp
provided the aportite fission track data. We gratefully acknowledge the staff at the Institute
of Geological and Nuclear Sciences in Lower Hutt, New Zealand for much needed scientific
and logistical support. Two anonymous reviewers provided detailed and helpful reviews.

REFERENCES

Allen, P. A., and Allen, J. R., 1990, Basin analysis - principles and applications: Blackwell, London, 451 p.
Allis, R. G., Annstrong, P. A., and Funnell, R. H., 1995, Implications of high heat flow anomaly around New
Plymouth: New Zealand Jour. Geol. Geophys., v. 38, no. 2, p. 121-130.
Allis, R. G., Annstrong, P. A., Funnell, R. H., and Chapman, D. S., 1997a, Exhumation of continental crust
deduced from offset porosity-depth trends (abst.): EOS, Trans. AGU, v. 78, p. F640.
Allis, R. G., Zhan. X, Evans, C., and Kroopnick, P., 1997b, Groundwater flow beneath Mt. Taranaki, New
Zealand, and implications for oil and gas migration: New Zealand Jour. Geol. Geophys., v. 40, no. 2, p.
137-149.
Armstrong, P. A., 1996, Thermal history of the Taranaki Basin, New Zealand: unpubl. doctoral dissertation,
Univ. Utah, 245 p.
Annstrong, P. A., and Chapman, D. S., 1998, Beyond surface heat flow: an example from a tectonically active
sedimentary basin: Geology, v. 26, no. 2, p. 183-186.
Annstrong, P. A., Allis, R. G., Funnell, R. H., and Chapman, D. S., in press, Late Neogene erosion patterns in
Taranaki Basin (New Zealand): evidence from offset porosity-depth trends: submitted to Jour. Geophys.
Res.
Annstrong, P. A., Chapman, D. S., Funnell, R. H., Allis, R. G., and Kamp, P. J. J., 1996, Thermal modelling and
hydrocarbon generation in an active-margin basin: Taranaki Basin, New Zealand: Am. Assoc. Petroleum
Geologists Bull., v. 80, no. 8, p. 1216-1241.
Armstrong, P. A., Kamp, P. J. J., Allis, R. G., and Chapman, D. S., 1997, Thermal effects of intrusion below the
Taranaki Basin (New Zealand): evidence from combined apatite fIssion track age and vitrinite reflectance
data: Basin Research, v. 9, no. 2, p. 151-169.
Birch, F., Roy, R. F., and Decker, E. R., 1968, Heat flow and thermal history in New England and New York,
in Zen, E., White, W. S., Hadley, J. B., and Thompson, J. B., eds., Studies of Appalachian Geology--
Northem and Maritime: Interscience, New York, p. 437-451.
Blackwell, D. D., and Steele, J. L., 1989a, Thermal conductivity of sedimentary rocks: measurement and
significance, in Naeser, N. D., and McCulloh, T. H., eds., Thermal History of Sedimentary Basins:
Methods and Case Histories: Springer-Verlag, New York, p. 13-36.
Blackwell, D. D., and Steele, J. L., 1989b, Heat flow and geothermal potential of Kansas: Kansas Geol. Survey
Bull. 226, p. 267-295.
Bradshaw, J. D., 1993, A review of the Median Tectonic Zone: terrane boundaries and terrane amalgamation
near the Median Tectonic Line: New Zealand Jour. Geol. Geophys., v. 36, no. I, p. 117-125.
174 ARMSTRONG AND CHAPMAN

Bullard, E. C., 1947, The time necessary for a borehole to attain temperature equilibrium: Mon. Notes Roy. Astr.
Soc., v. 5, no. 5, p. 127-130.
Burnham, A. K., and Sweeney, J. J., 1989, A chemical kinetic model of vitrinite maturation and reflectance:
Geochim. Cosmochim. Acta, v. 53, no. 10, p. 2649-2657.
Cermak, V., Bodri, L., Rybach, L., and Buntebarth, G., 1990, Relationship between seismic velocity and heat
production: comparison of two sets of data and test of validity: Earth Planet. Sci. Let., v. 99, no. 1-2, p.
48-57.
Chapman, D. S., and Furlong, K. P., 1992, Thermal state of the continental lower crust, in Fountain, D. M.,
Arculus, R., and Kay, R. M., eds., Continental lower crust: Developments in Geotectonics, v. 23, Elsevier,
Amsterdam, p. 177-199.
Davy, B., 1992, Magnetic anomalies of the New Zealand basement: New Zealand Oil Explor. Conf. Proc. 1991,
p. 134-144.
Deming, D., 1989, Application of bottom-hole temperature corrections in geothermal studies: Geothermics, v.
18, no. 5-6, p. 775-786.
Deming, D., 1994, Estimation of thermal conductivity anisotropy ofrock with application to the determination
of terrestrial heat flow: Jour. Geophys. Res., v. 99, no. Bll, p. 22,087-22,091.
Deming, D., and Chapman, D. S., 1988, Heat flow in the Utah-Wyoming thrust belt from analysis of bottom-
hole temperature data measured in oil and gas wells: Jour. Geophys. Res., v. 93, no. B11, p. 13,657-
13,672.
Funnell, R. H., Allis, R. G., Chapman, D. S., and Armstrong, P. A., 1996, Thermal regime of the Taranaki basin,
New Zealand: Jour. Geophys. Res., v. 101, no. B11, p. 25,197-25,215.
Furlong, K. P., Hanson, R. B., and Bowers, J. R., 1991, Modelling thermal regimes, in Kerrick, D. M., ed.,
Contact Metamorphism: Mineral. Soc. America, Reviews in Mineralogy, v. 26, p. 437-505.
Heasler, H. P., and Kharitonova, N. A., 1996, Analysis of sonic well logs applied to erosion estimates in the
Bighorn Basin, Wyoming: Am. Assoc. Petroleum Geologists Bull., v. 80, no. 8, p. 630-646.
Hermanrud, C., 1993, Basin modelling techniques - an overview, in Dore, A. G., and others, eds., Basin
Modelling: Advances and Applications: Elsevier, Amsterdam, p. 1-34 ..
Herrin, J. M., and Deming, D., 1996, Thermal conductivity of U.S. coals: Jour. Geophys. Res., v. 101, no. Bl1,
p.25,381-25,386.
Holt, W. E., and Stem, T. A., 1994, Subduction, platform subsidence, and foreland thrust loading: the late
Tertiary development of Taranaki basin, New Zealand: Tectonics, v. 13, no. 5, p. 1068-1092.
Homer, D. R., 1951, Pressure buildup in wells: 3,d Proc. World Petrol. Cong., v. 2, p. 503-521.
Homibrook, N., 1992, New Zealand Cenozoic marine paleoclimates: a review based on the distribution of some
shallow water and terrestrial biota, in Tsuchi, R., and Ingle, J. C., eds., Pacific Neogene, the Environment,
Evolution, and Events: Univ. Tokyo Press, Tokyo, p. 83-106.
Issler, D. R., 1992, A new approach to shale compaction and stratigraphic restoration, Beaufort-Mackenzie basin
and Mackenzie Corridor, northern Canada: Am. Assoc. Petroleum Geologists Bull., v. 76, no. 8, p. 1170-
1189.
Issler, D. R., and Beaumont, c., 1989, A fmite element model of the subsidence and thermal evolution of
extensional basins: application to the labrador continental margin, in Naeser, N. D., and McCulloh, T. H.,
eds., Thermal History of Sedimentary Basins: Methods and Case Histories: Springer-Verlag, New York,
p.239-268.
Kamp, P. J., and Hegarty, K. A., 1989, Multigenetic gravity couple across a modem convergent margin:
inheritance from Cretaceous extension: Geophys. Jour., v. 96, no. 1, p. 33-41.
Kappelmeyer, 0., and Haenel, R., 1974, Geothermics with special reference to application, in Rosenbach, 0.,
and Morelli, C., eds., Geoexploration Monograph Seriel I: Geopublication Association, Berlin.
King,P. R., 1994, The habitat of oil and gas in the Taranaki basin: New ,Zealand Explor. Conf. Proc. 1994, p.
180-203.
King, P. R., and Thrasher, G. P., 1992, Post-Eocene development of the Taranaki basin, New Zealand,
convergent overprint of a passive margin, in Watkins, J. S., Zhiqiang, F., and McMillen, K. J., eds.,
Geology and Geophysics of Continental Margins: Am. Assoc. Petroleum Geologists Mem. 53, p. 93-118.
King, P. R., and Thrasher, G. P., 1996, Cretaceous-Cenozoic geology and petroleum systems of the Taranaki
Basin, New Zealand: Inst. Geol. And Nuclear Sciences Ltd., Monograph 13, Lower Hutt, New Zealand;
243 p., 6 enclosures.
Knox, G. J., 1982, Taranaki Basin, structural style and tectonic setting: New Zealand Jour. Geol. Geophys., v.
25, no. 2, p. 125-140.
COMBINING TECTONICS AND THERMAL FIELDS 175

Lachenbruch, A. H., 1970, Crustal temperatures and heat production: implications of the linear heat-flow
relationship: Jour. Geophys. Res., v. 75, no. 17, p. 3291-3300.
Lachenbruch, A. H., and Brewer, M. C., 1959, Dissipation of the temperature effect of drilling a well in arctic
Alaska: U.S. Geol. Survey Bull. 1083C, p. 73-109.
Lachenbruch, A. H., and Sass, J. H., 1978, Models of an extending lithosphere and heat flow in the Basin and
Range province: Geol. Soc. America Mem. 152, p. 209-250.
Lee, Y., and Deming, D., 1998, Evaluation of thermal conductivity temperature corrections applied in terrestrial
heat flow studies: Jour. Geophys. Res., v. 103, no. B2, p. 2447-2454.
Locke, C. A., Cassidy, J., and MacDonald, A., 1993, Three-dimensional structure of relict stratovolocanoes in
Taranaki, New Zealand: evidence from gravity data: Jour. Volcanology and Geothermal Res., v. 59, no.
1-2, p. 121-130.
Luheshi, M. N., 1983, Estimation offormation temperatures from borehole measurements: Geophys. Jour. Roy.
Astro. Soc., v. 74, no. 3, p. 747-776.
Magara, K., 1976, Thickness of removed sedimentary rocks, paleopore pressure, and paleotemperature,
southwestern part of Western Canada basin: Am. Assoc. Petroleum Geol. Bull., v. 60, no. 4, p. 554-566.
Neall, V. E., 1979, Geological map of New Zealand - New Plymouth, Egmont, and Manaia: Department of
Scientific and Industrial Research, Wellington, Scale: 1:50,000 ..
Nielsen, S. B., and Balling, N., 1990, Modelling subsidence, heat flow, and hydrocarbon generation in
extensional basins: First Break, v. 8, no. 1, p. 23-31.
Peacock, S. M., 1989, Thermal modeling of metamorphic pressure-temperature-time paths: forward approach,
in Crawford, M. L., and Padovani, E., eds., Metamorphic Pressure-Temperature-Time Paths: Am.
Geophys. Union, Washington, D.C., p. 57-102.
Pilaar, W. F. H., and Wakefield, L. I., 1978, Structural and stratigraphic evolution of the Taranaki Basin,
offshore North Island, New Zealand: Aust. Petrol. Expl. Assoc. Jour., v. 18, pt. 1, p. 93-101.
Richard, S. M., Smith, C. H., Kimbrough, D. L., Fitzgerald, P. G., Luyendyk, B. P., and M.McWilliams, M. 0.,
1994, Cooling history of the northern Ford Ranges, Marie Byrd Land, West Antarctica: Tectonics, v. 13,
no. 4, p. 837-857.
Ridgway, N. M., 1969, Temperature and salinity of sea water at the ocean floor in New Zealand region: New
Zealand Jour. Marine and Freshwater Res., v. 3, p. 57-72.
Roman-Berdiel, T., Galais, D., and Brun, J. P., 1995, Analogue models of laccolith formation: Jour. Struct.
Geology, v. 17, no. 9, p. 1337-1346.
Rose, K. H., 1991, Cook gravity anomaly map, oceanic series: New Zealand Dept. Sci. and Ind. Res.,
Wellington, scale 1:1,000,000 ..
Roy, R. F., Blackwell, D. D., and Birch, F., 1968, Heat generation of plutonic rocks and continental heat flow
provinces: Earth Planet. Sci. Lett., v. 5, no. 1, p. 1-12.
Sass, J. H., Lachenbruch, A. H., and Munroe, R. J., 1971, Thermal conductivities of rocks from measurements
on fragments and its application to heat flow determinations: Jour. Geophys. Res., v. 76, no. 14, p. 3391-
3401.
Selater, J. G., and Christie, P. A. F., 1980, Continental stretching: an explanation of the post-mid-Cretaceous
subsidence of the central North Sea Basin: Jour. Geophys. Res., v. 85, no. B7, p. 3711-3739.
Smith, L., and Chapman, D. S., 1983, On the thermal effects of groundwater flow I. Regional scale systems:
Jour. Geophys. Res., v. 88, no. B1, p. 593-608.
Stern, T. A., and Davey, F. J., 1990, Deep seismic expression of a foreland basin: Taranaki basin, New Zealand:
Geology, v. 18, no. 10, p. 979-982.
Thrasher, G. P., and Cahill, J. P., 1990, Subsurface maps of the Taranaki basin, New Zealand: New Zealand
Geol. Survey Rept. G 142,45 p., 14 sheets ..
Tulloch, A. J., and Kimbrough, D. L., 1989, The Paparoa metamorphic core complex, New Zealand: Cretaceous
extension associated with fragmentation of the Pacific margin of Gondwana: Tectonics, v. 8, no. 6, p.
1217-1234.
Tulloch, A. J., and Pahner, K., 1990, Tectonic implications of granite cobbles from the mid-Cretaceous Pororari
Group, southwest Nelson, New Zealand: New Zealand Jour. Geol. Geophys., v. 33, no. 2, p. 205-217.
Van Wagoner, T., Armstrong, P. A., and Chapman, D. S., 1997, Thermal conductivity anisotropy oflaminated
sedimentary rocks (abst.): EOS, Trans. AGU, v. 78, p. F677.
Walcott, R. 1.,1987, Geodectic strain and the deformational history of the North Island of New Zealand during
the late Cainozoic: Phil. Trans. Roy. Soc. London, v. A 321, no. 1557, p. 163-181.
176 ARMSTRONG AND CHAPMAN

Wells, P. E., 1990, Porosities and seismic velocities of mudstones from Wairarapa and oil wells of North Island,
New Zealand, and their use in determining burial history: New Zealand Jour. Geol. Geophys., v. 33, no.
1, p. 29-39.
Willett, S. D., 1988, Spatial variation of temperature and thermal history of the Uinta Basin: unpubl. doctoral
dissertation, Univ. Utah, 121 p.
Willett, S. D., 1992, Modelling thermal annealing of fission tracks in apatite, in Zentilli, M. ,and Reynolds, P.
H., eds., Short Course on Low Temperature Thennochronology: Mineral. Assoc. Canada Short Course
Handbook, p. 42-74.
Willett, S. D., and Chapman, D. S., 1987, Analysis of temperatures and thennal processes in the Uinta Basin,
in Beaumont, C., and Tankard, A. J., cds., Sedimentary Basins and Basin-Forming Mechanisms: Can. Soc.
Petrol. Geology Mem. 12, p. 447-461.
THERMAL HISTORY OF A DEEP WELL
IN THE MICHIGAN BASIN: IMPLICATIONS
FOR A COMPLEX BURIAL HISTORY

William D. Everham and Jacqueline E. Huntoon

Department of Geological Engineering and Sciences


Michigan Technological University, Houghton, Michigan

ABSTRACT

Most of the Paleozoic strata in the Michigan Basin exhibit levels of thermal maturity
that are higher than expected given present-day geothermal gradients, burial depths, and heat
flow. Because ofthe complexity of the basin's thermal history, no previous study has resulted
in a single model that predicts the observed thermal maturity for the entire stratigraphic section
in the basin. Two hypotheses, however, have resulted from previous studies. They are: (1) that
elevated basal heat flow affected the basin during its early history, and (2) that a significant
amount of Late Carboniferous and Permian strata must at one time have been present in the
basin.
This paper describes the results of several I-dimensional model simulations of the
time-dependent conductive thermal history of the Mobil-Jelinek well, located in Shiawassee
County, Michigan. The Mobil-Jelinek well has been the focus of several previous studies
because of abundance of thermal maturity data for the well. These data were used in this study
to constrain some model input parameters and to evaluate the modeling results. A substantial
subsurface database also was used to construct isopachous maps and structural cross sections
that aided reconstruction of a burial history for the Mobil-Jelinek well.
Results of the best-fit model predict thermal maturities that are consistent with
observed data for the entire stratigraphic section of the well. The best-fit model requires
elevated heat flow (greater than present-day values) through the basin until the end of the
Silurian. A significant amount of erosion (2.3 km) also is incorporated in the model at the
SilurianlDevonian boundary. This boundary represents the transition from the Tippecanoe to
Kaskaskia Sequences. Finally, about 2 km of excess Late Carboniferous and Permian strata
(which was eroded during the Triassic) is included in the best-fit model.

177
178 EVERHAM AND HUNTOON

INTRODUCTION

The Michigan Basin is an intracratonic basin located in the upper Midwestern United
States. The basin is radially symmetric, covers an area of 207,000 km2, and is centered on
Michigan's Lower Peninsula. It includes all of the Lower Peninsula, part of the Upper
Peninsula, portions of eastern Wisconsin, western Ontario, and small sections of northern
Illinois, Indiana, and Ohio. The basin's strata reflect deposition during the Sauk through Zuni
Sequences; the Tejas sequence is present only as unconsolidated glacial debris (Sloss, 1963;
Catacosinos, Daniels, and Harrison, 1991). The youngest rocks (disregarding the glacial
deposits) are Jurassic in age and are located only in the center of the basin. Permian and
Triassic age rocks are missing in the basin (Moyer, 1982). Figure 1 shows the basin's extent,
several of its bounding features, and the location and ages of the glacial subcrop. Figure 2
shows a generalized NE-SW cross section of the basin's Lower Ordovician to Jurassic strata
and an E-W representative diagrammatic cross section of the Precambrian basement including
several prominent features.

Canadian Shield

J
11'
Jurassic
Pennsylvauian
S Silurian
0 Ordovician
* Mobil-Jelinek Well

M Mississippian £ Cambrian
D Devonian P€ Precambrian

Figure 1. Location and extent of Michigan Basin in upper midwestern United States. Several structural trends
and basin bounding features are shown, along with glacial subcrop ages. Star (*) indicates location of Mobil-
Jelinek well. Also shown are locations of cross sections in Figure 2. Modified from Catacosinos, Daniels, and
Harrison (1991).
THERMAL mSTORY OF THE MICmGAN BASIN 179

o 100 200 300 400 500 600


Distance (km)

E
HOWELL ST. CLAIR

Figure 2. Cross section A (cross-section line shown in Fig. 1): generalized depiction of Michigan Basin's
Middle Ordovician to Jurassic strata (modified from Allen and Allen, 1990, p. 232). Horizon ages are from
AAPG COSUNA Project (1985). Cross section B (cross-section line shown in Fig. 1): diagrammatic
representation of Precambrian basement (modified from Fisher and others, 1988, p. 367).

No previous modeling study has produced results that satisfy the observed thermal
maturity data for the entire stratigraphic section in the basin. Results of Cercone's (1984)
study suggested that an elevated geothermal gradient existed in the basin from the Cambrian
through the Carboniferous, and that the basin was once buried by approximately 1 km of post-
Pennsylvanian sediment that was removed prior to the Late Jurassic. One of Cercone's models
(using a geothermal gradient of 45°CIkm from the Cambrian to the beginning of the Permian)
successfully predicted the thermal maturity of upper Paleozoic rocks, but overestimated the
maturities of older rocks. Conversely, Cercone's model using a geothermal gradient of
35°CIkm (from the Cambrian up to the beginning of the Permian) reasonably predicted the
maturities oflower Paleozoic rocks, but underestimated the maturity of younger strata.
A study by Nunn, Sleep, and Moore (1984) suggested that petroleum would have been
produced only from Ordovician and older rocks in the basin. Because in reality, petroleum is
sourced from younger strata (Gardner and Bray, 1984; Illich and Grizzle, 1983, 1985; Pruitt,
1983), Nunn, Sleep, and Moore suggested that the temperature gradient and depth of burial
was greater in the past, or that the model used to calculate the maturation of organic matter was
not appropriate. The gradient (45-50°CIkm) necessary to produce oil from Silurian and
180 EVERHAM AND HUNTOON

Devonian rocks is significantly higher than predicted by their tectonic model, so they
concluded that the depth of burial of units in the central part of the basin was underestimated
by the modeling.
Cercone and Pollack (1991) suggested that the addition of 2 km of Permo-
Carboniferous strata, that was removed subsequently by the early Mesozoic, could produce
observed thermal maturities, even without elevated Paleozoic heat flow. The results of their
models, however, did not accurately predict the thermal maturities for the entire stratigraphic
section. Cercone and Pollack (1991) further tested the hypothesis that deposition of excess
Permo-Carboniferous strata combined with elevated geothermal gradients could produce the
observed maturities. For example, one model that incorporated 1 km of Permo-Carboniferous
overburden and a geothermal gradient of 40 °CIkm accurately predicted the thermal maturity
of Devonian and younger strata but underestimated the thermal maturity of pre-Devonian
strata. On the other hand, a different model that included 1 km of Permo-Carboniferous
overburden and a geothermal gradient of 60 °CIkm roughly predicted the thermal maturity of
pre-Devonian strata but overestimated the thermal maturity of Devonian and younger strata.
In their study of several wells in the Michigan Basin, Daly and Lilly (1985) used a
subsidence and temperature history similar to that ofNunn, Sleep, and Moore (1984), but
employed the Lopatin (1971), rather than the Tissot and Espitalie (1975), method for
determining thermal maturity. Their results indicate that the Upper Ordovician across much
of the basin is within the oil window and that at the basin center it is at or near peak levels of
oil generation. Their results also show that at the basin center, the Upper Silurian shows
marginal maturity. These results differ significantly from those ofNunn, Sleep, and Moore
and the difference entirely reflects the selection of organic maturation model.
Wang, Crowley, and Nadon (1994) investigated a different method of calculating
organic maturity. They used Sweeny and Burnham's (1990) method for Devonian and
younger strata and a single activation energy (48 kcaVmole) for pre-Devonian organic material.
The pre-Devonian activation energy was based on results of Hunt, Lewan, and Hennet (1991)
for marine organic material. They could match vitrinite reflectance data for Devonian and
younger strata as well as a single point in pre-Devonian strata (the base of the Utica Shale)
(Fig. 3) by incorporating deposition of 1.8 km of Carboniferous material that was eroded
subsequently during the Permian and Triassic. Wang, Crowley, and Nadon used a geothermal
gradient of 26.2 ° CIkm (equivalent to the present-day gradient) in their model. Although their
model produced results that matched thermal maturity data for post-Devonian strata and the
base of the Utica Shale, their model severely overestimated the maturity of strata older than
the Upper Silurian and underestimated the maturity of pre-Devonian strata above the Utica
Shale.
There is debate about the suitability of using "vitrinite" reflectance data for pre-
Devonian rocks, in that there is no true pre-Devonian vitrinite. Cercone and Pollack (1991)
indicate that the reflectance measurements for the pre-Devonian strata in the Mobil-Jelinek
well were made on chitinous and macrinitic material. Alpern (1980), Bertrand and Heroux
(1987), and Bertrand and others (1985) stated that the reflectances of vitrinite and chitinous
material are almost the same. In his study comparing maturation indicators, Bertrand (1990)
concluded that chitinozoans and other zooclasts can be utilized effectively as thermal maturity
indicators in vitrinite-free pre-Devonian strata and that vitrinite reflectance values are only
slightly higher than that of chitinozoans.
Although previous studies have been unsuccessful in accurately predicting thermal
maturities for the basin's entire stratigraphic section, they have resulted in two main
hypotheses about the basin's thermal history: (1) that elevated basal heat flow affected the
THERMAL HISTORY OF THE MICHIGAN BASIN 181

SYSTEM lM~ FORM~


~"'"
IQUAR'tr.KNA~Y 1.6 TEJAS ~
~TI~RY
144
"REDBEDS"
JURASSIC ZUNI
200
290 GRANDRNER
==
rJl
P ABSAROKA
0 U

~
~G~
r- 330
Z
0 ;w
L
~
U
of A

BEREA
365 ~
ELSWORTH --=s
~
,,~~
U KASKASKIA ~VE~FORMATION"-

r- fRA VERSE lIMESTONE

~
0 M DUNDEE
:>
r.< LUCAS -
~
r- DETROIT RNER GP. '~~~S~i:RG

L GARDEN ISLANJOIS BLANC


405 BASS ISl 'V'illi
Z U
~ r- ~
§
rJl
L
.. BURN r BI dtl':
flE-'"
TIPPECANOE M LIN
425
CiliCINNATIAN NOIF.ROCKS
~U U
I'lC.Ac

!> BL ;K~

0
§
0
~ ~~
PRAlREDl lIEN JP
500 EA
~ U
(AN(

~ SAUK fALl>
~
r- 510
~
u
L
I P'H'C'AMRJ> O\N 570
PRE-! SIMON

Figure 3. Michigan Basin stratigraphy. Modified from Catacosinos, Daniels, and Harrison (1991) and AAPG
(1985).

basin during its early history (e.g., Nunn, Sleep, and Moore, 1984; Cercone, 1984); and (2) that
significant amounts of Pennsylvanian and Pennian strata must have at one time been present
in the basin (e.g., Cercone and Pollack, 1991; Wang, Crowley, and Nadon, 1994).
This modeling study centers on the Mobil-Jelinek well located in Shiwassee County,
Michigan (Fig. 1). The Mobil-Jelinek well has been the subject of several previous modeling
studies (e.g., Cercone and Pollack, 1991; Wang, Crowley, and Nadon, 1994) because vitrinite
reflectance data are available for all of the penetrated stratigraphic section. In this study, the
hypothesis is examined that purely vertical conductive heat transfer coupled with a complex
burial and erosion history can explain the observed levels of vitrinite reflectance in the Mobil-
Jelinek well. All of the vitrinite reflectance measurements from the Mobil-Jelinek well are
assumed to be of equal accuracy. The numerical model used here produces results similar to
those of previous studies when the same initial conditions, boundary conditions, and burial
history models are used.
182 EVERHAM AND HUNTOON

This paper first will discuss the modeling method used as well as the numerical model
itself. Secondly, vitrinite reflectance data from the Mobil-Jelinek well are presented and
discussed. Thirdly, several sensitivity models and their results are described. These models
incorporate a variety of heat-flow and burial histories. Finally, the development of input
parameters and results of the best-fit model are discussed.

MODELING METHOD

The first step in the thermal modeling study was to assemble available data that could
be used to constrain the thermal history of the basin. These data include tectonic modeling
studies that interpret the basin's thermal history based on observed subsidence patterns (e.g.,
Haxby, Turcotte, and Bird, 1976; Nunn, Sleep, and Moore, 1984; Nunn, 1994). Organic
thermal maturity indicators such as thermal alteration indices, vitrinite reflectance, and
conodont color alteration indices (e.g., Cercone, 1984; Hogarth and Sibley, 1985; Bowers,
1989; Cercone and Pollack, 1991; Wang, Crowley, and Nadon, 1994;) are highly desirable
because they can be measured directly and are also correlated readily to stages of hydrocarbon
generation. Fission-track data are also available, and they provide good estimates of the time
of maximum burial and subsequent unroofing for the Michigan Basin and surrounding area
(e.g., Crowley, 1991; Wang, Crowley, and Nadon, 1994). In this study the forward simulation
begins at a time corresponding to deposition of the top ofthe Mount Simon Sandstone. This
horizon was selected for the start of the modeling because the overlying units are the oldest to
reflect nearly radially symmetric subsidence within the Michigan Basin. Figure 3 shows the
basin's stratigraphy.
The second step was to construct an initial burial history for a specific site within the
basin. The specific site (Mobil-Jelinek well) was selected because high-quality organic
thermal maturity data are abundant for this well. The initial burial history was modified during
the course of the modeling study when it was apparent that the initial history could not
reproduce observed levels of maturity. The Correlation of Stratigraphic Units of North
America (COSUNA) Midwestern Basin and Arches Region chart (AAPG, 1985) provided age
data for the formations included in each burial unit. The age data were used to estimate the
time of each burial unit's deposition, and were used to estimate the timing of erosion events.
The group and formation thickness data were collected from the Aangstrom Precision
Corporation Michigan Oil and Gas Well Database (1990), and descriptions in Lilienthal (1978)
and Fisher and others (1988). The Aangstrom Precision Corporation database provided Top
Measured Depth (TMD) values for all but one of the burial units. The depth to the Mt. Simon
Sandstone (the lowest horizon monitored in this model) was determined from a structure
contour map and cross section produced from Aangstrom Precision Corporation data using the
GEOGRAPHIX Exploration System © software. Lithology information for each burial unit in
the burial history was based on Lithologic Sample Descriptions (Aangstrom Precision
Corporation, 1990) and published formation descriptions (Lilienthal, 1978; Fisher and others,
1988).
The thermal conductivity assigned to each burial unit was based on the unit's lithology
and the thermal conductivity values compiled by Cermak and Rybach (1984). Burial units
consisting of multiple lithologies were assigned a weighted average thermal conductivity.
Uncertainties in thermal conductivity values constitute the largest source of error in this
modeling study because measured thermal conductivity values for basin strata are lacking.
THERMAL mSTORY OF THE MICmGAN BASIN 183

Initially, the heat-flow history was based on results of previous tectonic models of the
basin's evolution (e.g., Haxby, Turcotte, and Bird, 1976; Nunn, Sleep, and Moore, 1984). This
heat-flow history was modified after it was determined to be inconsistent with observed levels
of thermal maturity in the basin. The surface temperature in the models was set to 20°C
because during much of the basin's depositional history it was near the paleoequator (Cercone,
1984).
The third step in the modeling was to input the burial history, heat-flow history,
thermal conductivities, and surface temperature to a numerical model that calculates the
temperature of specified stratigraphic levels through time. Calculated temperatures then were
used to calculate vitrinite reflectance and Time Temperature Index (TTl) values for specified
stratigraphic levels as a function oftime.
The final step in the modeling was to compare the model output to organic thermal
maturity data to determine whether the model was effective at predicting the thermal history
of the basin.

THE NUMERICAL MODEL

Thermal histories are calculated for each burial unit using a forward model that solves
the one-dimensional heat conduction equation.
iJ 2 T 1 or A
(1)
&2 -a8t=-T
In this equation, T = temperature COC), Z = depth (km), IX = thermal diffusivity (m2/s), t = time
(m.y.), A = heat production (mW/m 3), and k = thermal conductivity (W/mK). The numerical
model calculates the temperature distribution in a 30-km thick section of the lithosphere as a
function of time, and is described elsewhere (Furlong and Edman, 1989; Huntoon, 1990;
Huntoon and Furlong, 1992; Price, Huntoon, and McDowell, 1996).
The model uses surface temperature and basal heat flow as boundary conditions. At
the start of each model run the user explicitly specifies surface temperature (20°C for all
models in this study). The user specifies a function describing a desired change in surface heat
flow through time. Surface heat flow is specified because that value (as opposed to basal heat
flow) usually is measured in sedimentary basins. Because heat generated by decay of
radioactive elements in crustal rocks contributes to surface heat flow, the basal heat-flow
boundary condition (imposed at a depth of 30 km) is calculated from the user specified surface
heat flow in the numerical model to account for cumulative heat production in the upper 30 km
of the crust.
(2)

In Equation (2), A = heat production as a function of depth (J.1W/m3 ), Ao = heat production at


the surface (2.5 J.1W/m 3 ), Z = depth (km), and Zo = depth at which heat production is equal to
lie of the surface heat production (10 km), where e is the natural logarithm base. Heat flow
is varied through time in the modeling because all tectonic studies and most thermal studies
of the Michigan Basin suggest that heat flow was higher in the past than it is at the present
time.
The numerical model requires that the user specify the timing, duration, and magnitude
of deposition and erosion events, as well as the thermal conductivities of deposited sediments
and basement materials. Thicknesses at the time of deposition of stratigraphic units which are
184 EVERHAM AND HUNTOON

present now in the basin were estimated from the measured present-day thicknesses using the
lithology-dependent decompaction algorithms of Baldwin and Butler (1985) for sandstone,
shale, and carbonate lithologies. Units composed dominantly of salt were not decompacted.
For decompaction purposes each stratigraphic unit was classified as a single lithology;
mixtures oflithologies were classified as the dominant lithology. In each model run the user
inputs the present-day thicknesses of stratigraphic units and the model calculates the maximum
depth of burial attained by the unit during evolution of the basin. Units are assumed to reach
their present-day thickness at the time of maximum burial so that the initial thicknesses are
calculated based on only the present-day thickness, maximum burial depth, and lithology.
During each model run, units compact from their initial thicknesses as they are buried. It is
assumed that units do not decompact during erosional events.
Thermal conductivity values for the burial units do not change with temperature or
depth of burial. Thermal conductivity is a function of density, specific heat, and thermal
diffusivity (Drury, Allen, and Jessop, 1984). Specific heat, diffusivity, and, to a lesser extent,
density are temperature dependant. Mongelli, Loddo, and Tramacere (1982) measured changes
in specific heat, diffusivity, and thermal conductivity with respect to increases in temperature
for several rock types. From their data it can be interpreted that if density is constant, an
increase in temperature causes thermal diffusivity and conductivity to decrease and specific
heat to increase. Robertson (1988) showed thermal conductivity values decreasing with
increasing temperature. Compaction of sediments (increasing solidity), however, also affects
thermal conductivity (Robertson, 1988). Compaction should result in an increase in thermal
conductivity because of the expulsion of water and enhanced grain to grain contacts (Cermak
and Rybach, 1985). Assuming that compaction is primarily a function of depth of burial,
calculations were made on several sediment types using the Baldwin and Butler (1985)
compaction curves and the solidity vs. conductivity and temperature vs. conductivity plots
from Robertson (1988). These calculations indicate that, whereas the decrease in thermal
conductivity resulting from increased temperature does not completely offset the increase in
thermal conductivity because of compaction, the differences in the resulting conductivities is
well within the uncertainties of the values themselves.
After the surface temperature, burial history, heat flow through time, and thermal
conductivity information are input, the numerical model calculates the time-dependent
temperature history of a one-dimensional (vertical) profile. Finite-difference techniques are
used to solve the heat equation at each time step. The results for each simulation include the
time-temperature history of every stratigraphic interval that is present now in the basin. From
this history, the thermal maturity of each horizon is estimated based on calculated TTl
(Waples, 1980) and vitrinite reflectance (%Ro) values. %Ro is calculated using Sweeny and
Burnham's (1990) Arrhenius first-order parallel-reaction method. The kerogen type used in
their model contains vitrinite and, therefore, is appropriate for use in modeling because the
results can be compared to observed %Ro data. TTl and %Ro are calculated because they are
used and readily correlated with other organic thermal maturity indicators.
The numerical model is a forward modeling simulation; time in the model progresses
forward from the past toward the present day. The initial conditions for the model therefore
are estimates of conditions at a particular time in the past. Because forward models are
nonunique, the best-fit model resulting from this study represents a single potential solution.
The best-fit model agrees with observed organic thermal maturity data and incorporates burial
and heat-flow histories that are geologically reasonable.
THERMAL mSTORY OF THE MICmGAN BASIN 185

HEAT-FLOW TYPE MODELS VS. GRADIENT-TYPE MODELS

The modeling method used here differs significantly from that used in previous studies
of the Michigan Basin's thermal history (e.g., Cercone, 1984; Cercone and Pollack, 1991;
Wang, Crowley, and Nadon, 1994). The previous studies utilized a gradient method, rather
than numerical solution techniques, to determine the temperature at specified depths as a
function of time. In gradient-type models subsurface temperatures depend on depth. Gradient
methods are accurate if the basin's heat-flow history is simple (e.g., constant heat flow through
time) (McCulloh and Naesar, 1989), if the deposition and erosion history is straightforward,
if neither deposition nor erosion proceed at a rapid rate (Deming and others, 1990), and if the
modeled rocks have constant thermal conductivities (Blackwell and Steele, 1989). None of
these criteria are satisfied in the Michigan Basin (e.g., Howell, 1993; Cercone, 1984; Wang,
Crowley, and Nadon, 1994). For example, all previous studies have suggested that the basin
experienced elevated heat flow during its early evolution (e.g. Cercone, 1984; Cercone and
Pollack, 1991; Wang, Crowley, and Nadon, 1994; Nunn, Sleep, and Moore, 1984). The large
variation in lithology of basin deposits (e.g., shale, sandstone, limestone, dolomite, and salts)
results in large variations in thermal conductivities (Robertson, 1988). In addition, several
previous studies have suggested that the basin has experienced episodes of significant and
rapid burial and erosion (e.g. Howell, 1993; Fisher and others, 1978).
Gradient-type models cannot adequately handle the effect of changes in heat flow
through time or the transient thermal effects of rapid burial and erosion. For example, rapid
burial has the thermal effect of moving the surface boundary condition downward through
time, because sediment is added faster than it can reach equilibrium temperatures. A gradient-
type model cannot accurately account for this process. Gradient models also are unable to
account adequately for changes in basal heat flow through time because a change in basal heat
flow propagates upward through the modeled interval during a finite period of time.

MOBIL-JELINEK WELL VITRINITE REFLECTANCE DATA

Vitrinite reflectance data collected from the Mobil-Jelinek well are used to constrain
the results of the modeling. Data from this well seem to be consistent in that wide variations
in reflectance at individual stratigraphic levels are not reported. This probably is because all
of the samples were analyzed at a single time by a single lab (Wang, Crowley, and Nadon,
1994).
Figure 4 shows a semilog plot of%Ro vs. depth for the Mobil-Jelinek well. A series
ofleast-squares best-fit lines are drawn through the data. In the post-Silurian section, a single
best-fit line is shown. In the pre-Devonian section, two different lines are shown. One line
fits data near the bottom of the well, whereas the second fits data near the top of the pre-
Devonian section. In the modeling we attempted to obtain calculated %Ro values that lie along
the upper (post-Silurian) best-fit line, and lie somewhere between the two lines shown for the
pre-Devonian section. Observed scatter in the data essentially is smoothed by using best-fit
lines. Inversions in the vitrinite reflectance vs. depth pattern, for example %Ro ~ 1.45 at 1737
m and %Ro ~ 1.19 at 1920 m may reflect measurement error or superposition of the effect of
groundwater movement on the conductive regime (Mansure and Reiter, 1979). Inversions
observed in the data may reflect the presence of reverse or thrust fault in the stratigraphic
section penetrated by the well (Dow, 1977), although no evidence for compressional features
at this site have been reported.
186 EVERHAM AND HUNTOON

0.1 1 "loRa 10
o
100 I I
II I
200

300 ,

400 I ,
500 • I

600 •
700 I

BOO • !

BOO I

!
g 1000 . Devonian I

i 1100 Silurian \ ,'


I
I
CD
C
1200 \
1300
\
1400

~.
1500 i

1600

1700

,\'I"
1800

1900
\
2000
It \
2100

2200
\ \
Figure 4. Semilog plot of measured vitrinite reflectance values (circles) vs. depth for Mobil-Jelinek well. Graph
shows offset data between pre-Devonian and Devonian and younger strata, suggesting presence of erosional
unconformity. Solid lines represent best-fit lines through data. In pre-Devonian section, two different lines are
shown. One line fits data near bottom of well, whereas second fits data near top of pre-Devonian section. In
modeling we attempted to obtain calculated %Ro values that lie along upper (post-Silurian) best-fit line, and lie
somewhere between two lines shown for pre-Devonian section. Dashed line roughly bisects angle between pre-
Devonian best-fit lines and is included to emphasi2e offset in pre-Devonian and post-Silurian %Ro values.

Two features of the data need to be emphasized because they are important in the
thermal modeling study. First, the post-Silurian data lie along a least-squares best-fit line that
is significantly offset from the best-fit lines through the pre-Devonian data. This offset
suggests that an erosional unconformity (Waples, 1985) is present at the SilurianlDevonian
boundary. Second, the slopes of the pre-Devonian best-fit lines are steeper than the slope of
the post-Silurian best-fit line. Pre-Devonian strata are made up predominantly of carbonates
and anhydrites, but Devonian and younger strata are mostly sands and shales with some
carbonates (Lilienthal, 1978; Fisher and others, 1988; Aangstrom Precision Corporation,
1990). Calculated average conductivities for pre-Devonian and Devonian and younger strata
THERMAL HISTORY OF THE MICHIGAN BASIN 187

are 2.8 W/m-K and 2.1 W/m-K, respectively. Because thermal conductivities of the pre-
Devonian section generally are higher than those of the post-Devonian section, the steep slopes
of the pre-Devonian least-squares best-fit lines are interpreted to indicate that higher heat flow
affected the basin during pre-Devonian time than during the Devonian to Recent. If the heat
flow in the basin was constant throughout time, the higher thermal conductivity of the pre-
Devonian strata would result in a smaller change in temperature with depth in the pre-
Devonian strata compared to the Devonian and younger strata. These two aspects of the data
were used in part to guide the modeling effort.

MOBIL-JELINEK WELL THERMAL STUDY

In this section the results of the sensitivity models are described first. These models
provide information that was used to develop more complex models of the basin's thermal
history. The burial history and heat-flow history that produced the best fit to the observed data
when modeled are described at the end of this section.

Sensitivity Models

The sensitivity models described here are grouped into three suites. The first suite of
models assumes that the burial history is simple and that heat flow is constant through time.
The second suite of models again assumes that heat flow is constant through time, but
examines the effect of burial by a thick section ofCarboniferouslPermian strata. The third
suite of models allows the heat flow to vary through time, and examines the effect of
deposition of a thick succession of excess material at the SilurianlDevonian boundary.
The burial history for the first suite of models includes excess strata assumed to have
been present once in the basin and subsequently eroded. Excess material is included at every
horizon represented by an erosional unconformity in the basin (Figs. 3, 5). The amount of
excess material added at each unconformity in the model was calculated by dividing in half
the time period represented by each lacuna, and then assuming that during the first half of the
lacuna deposition occurred at the same rate as calculated for the unit (that is present today)
immediately below. Erosion was assumed to occur at the identical rate during the second half
of the lacuna. This burial history results in addition and subsequent erosion of 180 m of
CarboniferousIPermian (CIP) strata. The amounts of additional strata included at the other five
the horizons now represented by unconformities in the basin are listed in Table 1.
The lithology of each eroded unit is assumed to be the same as that of the burial unit
immediately below the unconformity (Fig. 5). Because deposition in the eroded units is
followed immediately by erosion, the deposited material is not compacted fully and, therefore,
should have a lower thermal conductivity (Cermak and Rybach, 1985; Robertson, 1988). As
shown previously, both compaction and temperature affect the thermal conductivity. All or
a large portion of the eroded units undergo minimal amounts of burial and compaction.
Because of this, the decrease in thermal conductivity resulting from an increase in temperature
is small compared to the effect of reduced compaction and, therefore, the effect of temperature
change is not included. Thermal conductivity values for the eroded units were calculated,
based on the thermal conductivity values determined for the underlying units, using
compaction vs. depth curves (Baldwin and Butler, 1985) to determine solidity values and the
following equation (Robertson, 1988):
188 EVERHAM AND HUNTOON

(3)

where Kci is the corrected conductivity, Ci is the thennal impedance correction factor, y is the
solidity (or fractional grain volume), and Kc is the calculated, or measured conductivity.
Solidity (Robertson, 1988; Baldwin and Butler, 1985) constitutes a suitable frame of reference
for sediment compaction calculations because it is detennined by the proportion of solid grains
(solidity) as opposed to the proportion of pore space (porosity). Solidity, similar to porosity,
can be treated as a percentage of sediment thickness. When compaction is the only pore space

LITHOLOGY COMMENTS

SOm REMOVED

2300 m REMOVED

SOm REMOVED

200 m REMOVED

ANHYDRITE
- SHALE

Figure 5. Lithologic succession for Mobil-Jelinek well. Lithologies (Aangstrom Precision Corporation, 1990;
Lilienthal, 1978; and Fisher and others, 1988) are shown in relation to their depth, with zero representing surface.
Also shown are Sequence names, geochronologic time units, and ages. Amounts of material added and
subsequently eroded at each unconformity in best-fit model are shown on right side of figure (under Comments).
THERMAL HISTORY OF THE MICHIGAN BASIN 189

Table 1. Amounts of excess material included at lacunas in suite 1,2, and 3 sensitivity models.

Lacuna Amount of excess material added

Suite 1 Suite 2 Suite 3


AQe (Mal Nunn et al. Cercone
Carboniferous- 300-200 180m 1-2 km 2.4km 1 km
Triassic

Lower -Middle 347-325 200m 200m 200m 200m


Carboniferous

Middle Devonian ·389-387 50m 50m 50m 50m

Upper Silurian- 405-395 100m 100m 6.5km 4km


Lower Devonian

Upper Ordovician- 428-424 50m 50m 50m 50m


Lower Silurian

Lower -Upper 489-459 200m 200m 200m 200m


Ordovician

reducing process occurring, however, the thickness of solid grains can be considered a
constant. This leads to a linear relationship between solidity and length reduciton, whereas a
nonlinear relationship exists between porosity and length reduction (Baldwin and Butler,
1985). The use of solidity simplifies arithmetic calculations, because the volume of solid
grains, unlike pore space, does not change during compaction (Baldwin, 1971). Using the
solidity values from Baldwin and Butler's (1985) compaction curves along with thermal
impedance values from Robertson (1988) and Equation (3) permits calculation of thermal
conductivity values for the eroded units. Use of this method suggested that thermal
conductivities for the sediment in the eroded units (Table 2) should be reduced by between 30
and 50% from that of the underlying burial unit.
The first suite of models uses constant heat flows of 56 mW/m2, 100 mW/m2, and 200
mW/m2; 56 mW/m2 is the present-day heat flow near the Mobil-Jelinek well (Combs and
Simmons,1973). Models using heat flows of 100 and 200 mW/m2 were run to demonstrate
the effect of variations in heat flow for sensitivity analysis. Results of three of this first suite
of models can be seen in Figure 6. The models using constant heat flows of 56 mW/m2 and
100 m W1m2 underestimated the thermal maturities for the entire stratigraphic section. The
model using a heat flow of 200 mW/m2 underestimated the maturities for the Lower and
Middle Carboniferous strata and overestimated the maturities for the Middle Ordovician and
older strata. These results indicate that a simple burial history and constant heat flow through
time cannot be used to match the observed thermal maturities in the Mobil-Jelinek well.
The second suite of models again assumed constant heat flow through time, but
included various amounts of excess CIP strata. The additional CIP strata were eroded during
the Triassic in the entire second suite of models. The results of these models (Fig. 7) indicate
that a heat flow of 56 mW/m2 coupled with 2 km of excess CIP strata, or a heat flow 100
mW/m2 coupled with 1 km of additional CIP deposition, both produce an acceptable match to
observed maturities in Devonian and younger strata. Both of these models, however,
underestimate pre-Devonian maturities. The model using a heat flow 56 mW/m2 coupled with
190 EVERHAM AND HUNTOON

Table 2. Best-fit burial history of Mobil-Jelinek well; Shiawassee County, Michigan.

Formations Age Thick- Thermal Conductivity


Era/Period/Epoch Absolute (Ma) ness (m) (W/mK) & lithology

Trempealeau Formation Upper Cambrian- 510-474 640 2.49 (Is & dol)
Oneota Dolomite Middle Ordovician
New Richmond Sandstone
Shakopee Dolomite
Prairie Du Chein Group
EROSION Middle-Upper 474-459 200 1.25
Ordovician
St Peter Sandstone Upper Ordovician 459-456 120 2.29 (Is & sh)
Black River Group
Trenton Group Upper Ordovician 456-450 140 2.29 (Is)
Utica Shale Upper Ordovician 450-426 250 2.10 (sh & Is)
Cincinnatian
Undifferentiated Rocks
EROSION Upper Ordovician 426-424 50 1.05
Manitoulin Dolomite Lower-Upper 424-415 90 2.57 (dol, Is, & sh)
Cabot Head Shale Silurian
Niagara Group
Salina Group Upper Silurian 415-406 520 3.66 (dol, st, sh,
Is, & an)
Bass Island Group Upper Silurian- 406-400 2320 3.62 (dol)
Lower Devonian
EROSION Lower Devonian 400-395 2300 2.00
Bois Blanc Formation Lower-Middle 395-388 380 2.98 (dol, Is, sh, &
Sylvania Sandstone Devonian ss)
Lucas Formation
Detroit River Group
Dundee limestone
EROSION Middle Devonian 388-387 50 1.49
Bell Shale Middle-Upper 387-385 120 2.27 (dol,sh, & Is)
Traverse limestone Devonian
Traverse Formation
Traverse Group
Antrim Shale Upper Devonian- 385-363 110 2.23 (ss & sh)
Ellsworth Shale Lower Carboniferous
Bedford Shale
Berea Sandstone
Sunbury Shale Lower Carboniferous 363-359 300 2.38 (sh, dol, ss, &
Weir Sandstone Is)
Coldwater Shale
Marshall Sandstone Lower Carboniferous 359-336 310 2.22 (ss & sh)
Stray Sandstone
Michigan Formation
EROSION Lower-Middle 336-325 200 1.11
Carboniferous
Parma Sandstone Pennsylvanian- 325-250 2230 2.27 (ss & sh)
Saginaw Formation Permian
EROSION Triassic 250-200 2200 1.75
Glacial Drift Holocene 1-0 30 2.00 (sh & ss)
THERMAL HISTORY OF THE MICHIGAN BASIN 191

%Ro
0.1 10
0
I\ •
\ \ • • Measured values

\\ \ •• (c.....:ooe & Pollack, 1881)


()
II)
C-
O \ - - Constant q .. H mWlm2
\, \

::J
~ \
a I \

---~--~. Constant q • 100 mW/m2

\\
\
C \
'" \
\ • - - - - - Constant q ~ ZOO mWlm2

\.•
\
\
500- f-
\.
c
~
\
CD
< \ \.

\•
\
0
::J
iir • \
\
::J
• \
\

• \
\

\
\
\


\
~ 1000_ \
f- \
.s. \

.
\
J:: \
Q.
\

\ .
.!l \

!o. \\
§. \
\
II)
\

\\
::J \
\
\
1500 .\\
\
\


\

.
-f- \

\
\
.\
\

0
a. \ \

\
\
\
\

\
\
~.
Q. • \
\


\

.
II)
\
2000 ::J \

\ ••
\
\
\
\ \
\
. \
\
-~

Figure 6. Sernilog plot of%Ro vs. depth for first suite of sensitivity models. Models incorporate constant heat
flow through time and relatively simple burial history. Solid diamonds indicate measured %Ro values. Solid,
dashed, and dash-dot lines represent model results using heat flows of 56 mW/m2, 100mW1m2, and 200 mW1m2
respectively.

1 km of excess CIP strata severely underestimates the maturities throughout the stratigraphic
section in the well. On the other hand, the model using a heat flow 100 m W/m2 coupled with
2 km of excess CIP strata overestimates the maturities for the entire stratigraphic section in the
well. No combination of constant heat flow and any amount of CIP deposition could produce
acceptable results for both the pre-Devonian and Devonian and younger strata.
The third suite of sensitivity models investigated thermal histories proposed by Nunn,
Sleep, and Moore (1984) and Cercone (1984). Nunn, Sleep, and Moore suggested that heat
flow in the basin reached a maximum of 87 mW/m2 at about 460 Ma (Middle Ordovician) and
then gradually decreased to a constant of 53 mW/m2 at about 300 Ma (approximately the
CarboniferouslPermian boundary). One of Cercone's (1984) models assumed a geothermal
gradient of 45 °CIkm from the Cambrian through the Carboniferous and about 22°CIkm from
192 EVERHAM AND HUNTOON
O/ORo
0.1 10
a

\

f
()
• Measured values

a-0"'
(Cercone & Pollack, 1991)

::J - - Constant q .. 56 mW/m 2


;; w/1 km CIP of erosion

ac
tJI
.\~ Constant q =56 mW/m2

~;1
wi 2 km of C/P erosion

,\0 ------ Constantq .. 100 mW/m2


500
- ~
w/1 km of CIP erosion

..'.,
~
\t
--_.- Constantq .. 100 mW/m2
\ wi 2 km of C/P erosion
i I
i
0
(J)
<
\~ i
i
0 \\ i
::J
iii' \ i
::J i
+\ i

.\
.\ \ i
i
1000 \ I

'.§.
- - \
i
i
..
.
.<:
Q. i
c i
\ i
i

\...
!!l i
E" i
ffio i
::J
i
i
1500 i
i
, i
i
I-- • i
,, i
,, • • ii
,, i
,,

..
0
i3. , i
0
<
n: \. i
i

...
\ i
2000
"'
::J
\ i
i
\ i
I \ i
r--
c'>

Figure 7. Semilog plot of%Ro vs. depth for second suite of models. This graph shows results of models that
assume constant heat flow through time, but include addition of 1 and 2 Ian ofCarboniferouslPermian (CIP)
strata that is eroded during Triassic. Filled diamonds represent measured %Ro values. Lines represent heat flow
of 56 mW/m2 , with I Ian (solid) and 2 Ian (dotted) ofCIP deposition, respectively. Other lines represent heat
flow of 100 mW/m2 , with 1 Ian (dash-dot) and 2 Ian (long-dash-dot-dot) ofCIP deposition, respectively.

the Pennian to the Recent. Assuming a thennal conductivity of 2.1 WImK, the average
thennal conductivity ofthe rocks in the basin (Nunn, Sleep, and Moore, 1984) and steady-state
conditions, these gradients can be converted to heat flows of 95 mW/m2 and 46 mW/m2,
respectively. The models in this third suite that are based on Cercone's thennal history use a
heat flow of 95 mW/m2 from the beginning of the model (510 Ma) until 260 Ma (the
CarboniferouslPennian boundary), when it is decreased to an equilibrium value of 46 mW/m2.
The results of models using the Nunn, Sleep, and Moore and Cercone heat-flow histories (Fig.
8) and a relatively simple burial history, that is where the amount of excess material once
present at the unconfonnities is the same as for the suite 1 models, underestimated thennal
maturities for the entire stratigraphic section.
THERMAL HISTORY OF THE MICHIGAN BASIN 193

Modifications of the burial histories improved the results (Fig. 8). The model based
on Cercone's thermal history and a modified burial history, requiring 1 kIn of excess CIP
strata, and 4 kIn of excess strata at the SilurianlDevonian boundary, adequately matches the
observed thermal maturities in the Devonian and Lower Ordovician sections. The maturity of
the Upper Ordovician, Silurian, and Carboniferous sections is underestimated. The model
based on Nunn's thermal history and a modified burial history produced an acceptable match
to observed thermal maturities for the entire stratigraphic section. The modified burial history,
however, required 2.4 kIn of excess CIP strata, and 6.5 kIn of excess strata at the
SilurianlDevonian boundary.

%Ro
10

• Measured values
(Cercone & Pollack, 1991)

-- Cercone (1984) estimated


heat flow & simple burial history

---- Cen:one (1984) esUmated


heat ftow & best.ftt burial hl.tory

Nunn.t al. (1984) heat !low


& simple burial history

Nunn etal. (1884) heat now


& besMlt burial hiStory

o
~
0:
2000 !!l

Figure 8. Semilog plot of %Ro vs. depth for third suite of sensitivity models. Filled diamonds are measured
%Ro values. This graph shows model results using estimated heat flow based on thermal history of Cercone
(1984) with relatively simple burial history (solid line) and modified burial history to produce best fit to observed
thermal maturity data (dotted line). Also shown on this graph are results of models using heat flow ofNunn,
Sleep, and Moore (1984) with relatively simple burial history (dashed-dot line) and modified burial history to
produce best fit to observed thermal maturity data (long dash-dot-dot line).
194 EVERHAM AND HUNTOON

The conclusions drawn from these sensitivity models are that:


(1) A constant heat flow cannot be used successfully to model the thennal history of
the basin.
(2) Addition of excess CIP strata is necessary to produce observed levels of maturity
in Devonian to Carboniferous strata without requiring heat-flow values that are
unrealistically high.
(3) An additional amount of deposition is needed at the SilurianlDevonian boundary
to produce observed maturities in the pre-Devonian strata.

Best-Fit Model for the Burial History

A burial history for the Mobil-Jelinek well was produced using infonnation from a
variety of sources. When the simple burial history (as described for the first suite of sensitivity
models) failed to produce reasonable results, the burial history was modified (Table 2). For
example, the amount of material eroded during the Triassic (250-200 Ma) erosion event
previously was estimated at 1.2-2.0 km (Cercone and Pollack, 1991) and 2.0-5.0 km (Crowley,
1991). A thickness of2.2 km was used in our final best-fit model and this value is consistent
with the ranges suggested by previous studies. In addition, our modeling demonstrated that
it is necessary to add 2.3 km at the SilurianlDevonian unconfonnity. Additional burial at this
boundary was necessary to account for the offset between pre-Devonian and Devonian and
younger %Ro values observed in data from the Mobil-Jelinek well. Without this extra burial,
models that matched observed maturities for Devonian and younger strata could not match
maturities in the pre-Devonian section of the well. Although 2.3 km of excess strata is greater
than previously proposed in the literature, the SilurianlDevonian boundary does represent a
major unconfonnity corresponding to the transition from the Tippecanoe to Kaskaskia
Sequences (Sloss, 1963), and the unconfonnity at that level is used for interregional correlation
of strata.
For modeling purposes the eroded material at this boundary was assumed to be
composed primarily of carbonates. Deposition of2.3 km of carbonate material during the Late
Silurian would require a deposition rate of 0.46 rnmIyear in the burial history. Measured rates
of carbonate sedimentation differ between depositional environments. Ancient shallow marine
environments had sedimentation rates as low as 0.03 mm/year (Wilson, 1975). Modem rates
of carbonate sedimentation range from a maximum for sabkha-intertidal settings of 0.5
mm/year to about 3rnm1year for reefs and forereef talus (Wilson, 1975). Tippecanoe
subsidence rates for the Michigan Basin were estimated by Sloss (1988) to range up to 30m/my
(0.03 rnmIy). These subsidence rates, however, were generated from isopachous maps based
on present-day thicknesses and therefore represent a minimum subsidence rate. The
sedimentation rate (0.46 mm/year) required by the best-fit model for the deposition of the
excess strata at the SilurianlDevonian boundary therefore is reasonable.
The best-fit burial history requires two periods of major basin subsidence and
deposition followed by uplift and erosion of most of the deposited material. The most recent
event involved continued subsidence of the basin during the Carboniferous-Pennian, with
uplift and erosion occurring during the Triassic. The other event involves a period of rapid
subsidence in the Late Silurian with uplift and erosion starting at the end ofthe Silurian and
continuing into the Early Devonian. Although others have noted the CarboniferouslPennian
erosion, the SilurianlDevonian deposition and erosion event proposed here has not been
considered yet by tectonic models. Rapid subsidence of the basin during the Late Silurian may
THERMAL HISTORY OF THE MICHIGAN BASIN 195

be partially the result of Appalachian tectonics. Howell and van der Pluijm (1990) concluded
that variations in exponential basin subsidence might be related to Appalachian tectonics, in
that there seems to be a temporal correlation between Appalachian orogenic events and
changes in basin subsidence. They suggest that increased basin subsidence during the Late
Silurian may be related to the Acadian orogeny. Sea-level curves (Vail, Mitchum, and
Thompson, 1977) and studies of Silurian reef growth by Shaver and others (1983) indicate a
continual sea-level rise until approximately the Late Silurian. Because, under favorable
conditions, carbonate production can keep up with nearly any amount of sea-level change and
tectonic subsidence (Wilson, 1975), a combination of subsidence and continual sea-level rise
may have been enough to provide the accommodation space necessary for the required rate of
deposition.
Appalachian tectonics also may be responsible for the SilurianlDevonian uplift in the
basin; forebulges caused by overthrust loads along the Appalachian Orogen possibly affected
the geometry of the Michigan Basin (Quinlan and Beaumont, 1984). When the thrust-related
forebulge coincided with the location ofthe Michigan Basin, destructive interference between
the forebulge and the basin possibly resulted in uplift. This uplift, even when combined with
the regression that occurred during the Early Devonian (Vail, Mitchum, and Thompson, 1977;
Sloss, 1988), might not be of sufficient magnitude to account for erosion of all of the excess
SilurianJDevonian material however. Multiple basin heating events have been hypothesized
as a cause of discontinuous basin subsidence (Nunn and Sleep, 1984). Recent work by Sleep
(1996) indicates that mantle plume material can be transported hundreds of kilometers along
the base of the lithosphere. Sleep (1997) indicates that plume material moving along the base
of the lithosphere should pond beneath thinned lithosphere. The lithosphere beneath the
Michigan Basin may be thinned, although interpretations about lithospheric thickness are
dependant upon the interpreted mechanism of basin formation (Beaumont, 1978). Ponding of
a hot mantle plume, if it occurred, would result in uplift and erosion of previously deposited
sediment. It, however, would also require a heat-flow history more complex than the one used
in this study. Because forward models are nonunique, the best-fit burial and heat-flow
histories proposed in this study are just one of a number of possible models that could be
generated to fit the observed data. All models that reproduce the observed data, however,
require deposition and erosion of a significant amount of material at the SilurianJDevonian
boundary. Incorporation of a more complex heat-flow history (multiple heating events) could
reduce the total amount of excess strata required to reproduce the observed data.
Table 3 is a comparison of model results in which the thermal conductivity is varied
for the eroded material at the SilurianJDevonian boundary and the excess material deposited
during the Late CarboniferouslPermian. These model results indicate that changes in the
thermal conductivity of the eroded sections affect the amount of excess deposition required.
The amount of excess CarboniferouslPermian strata required by the best-fit model can be
reduced by 700 m if the thermal conductivity of the excess strata is reduced from 1.75 W/mK
to 1.15 W/mK. A conductivity of 1.15 W/mK is roughly the shale conductivity determined
by Blackwell and Steele (1989) and also that of the dark gray shale of Cercone, Deming, and
Pollack, (1996). Cercone and Pollack (1991) assumed the missing CarboniferouslPermian
strata to be fluvio-deltaic in nature, consisting of dark shales, carbonaceous mudstones, and
coaly material. These lithologies may have low thermal conductivities. Preserved
Pennsylvanian strata in the basin, however, consist mainly of interbedded sandstone, shale,
limestone, and coal (Lilienthal, 1978; Catacosinos and others, 1990). The thermal conductivity
of the eroded material would be higher than 1.15 WImK if it included sandstone and limestone.
196 EVERHAM AND HUNTOON

The amount of excess strata at the SilurianlDevonian boundary can be reduced by 300
m if thermal conductivity of the eroded material is reduced from 2.00 W/mK. to 1.70 W/mK..
The Silurian system in the basin is composed mainly of dolomite, limestone, and evaporites
(Lilienthal, 1978; Catacosinos and others, 1990). The uppermost Silurian rock unit, the Bass

Table 3. Comparison of best-fit thickness of eroded strata at


SilurianlDevonian boundary and excess CarboniferouslPermian material with
changes in thermal conductivity of eroded sections. These simulations used
previously determined heat-flow history from best-fit model.

Silurian/Devonian Carboniferous/Permian

Best-fit eroded Best-fit eroded


Conductivity section thickness Conductivity section thickness
(W/mK) (m) (W/mK) (m)

2.20 2500 1.95 2400


2.10 2400 1.85 2300
2.00 2300 1.75 2200
1.90 2200 1.65 2100
1.80 2100 1.55 2000
1.70 2000 1.45 1900
1.60 1900 1.35 1800
1.50 1800 1.25 1700
1.40 1700 1.15 1500
1.20 1500 1.00 1300

Islands Group, is composed of dolomite with some limestone intervals (Fisher and others,
1988; Lilienthal, 1978). The eroded material at the SilurianlDevonian boundary was assumed
to be a continuation of the deposition of the Bass Islands Group. A thermal conductivity less
than 1.70 W/mK. for the eroded strata would require a solidity less than that estimated for
carbonates at the time of their deposition (Baldwin and Butler, 1985; Robertson, 1988). Even
with the reduction in thermal conductivities of the eroded CarboniferouslPermian and
SilurianlDevonian material, significant amounts of excess strata are required to predict
adequately the observed thermal maturities in the basin.

Boundary Conditions. Measured values of the present-day terrestrial heat flow in the
basin range from 33 to 58 mW/m2 (Combs and Simmons, 1973; Nunn, 1994). Numerous
studies (e.g. Nunn, Sleep, and Moore, 1984; Crowley, 1991) have shown that the present-day
heat-flow values cannot explain the thermal history of the basin. For example, the in situ
generation of hydrocarbons in Silurian and Devonian formations would require higher
paleotemperatures (Crowley, 1991) than possible if present-day heat-flow values were
assumed to have been effective in the past. Even accounting for thermal effects of significant
amounts of burial and subsequent erosion, heat-flow values must have been higher in the past
(Cercone, 1984).
THERMAL HISTORY OF THE MICHIGAN BASIN 197

The initial heat flow used in the best-fit model was detennined from the plot of
measured vitrinite reflectance (%Ro) values vs. depth (Fig. 4) for the Mobil-Jelinek well.
Modeling could reproduce the slopes of the best-fit lines through the two groups of pre-
Devonian data when heat flows of 106 and 134 mW/m 2 were used. The average heat flow of
120 mW/m2 was selected as the initial heat flow for the model. It was necessary to use this
heat flow from the start of the model (510 Ma) to the SilurianlDevonian boundary (407 Ma)
to predict accurately observed vitrinite reflectance values in the pre-Devonian section. This
heat flow is consistent with tectonic models that predict elevated heat flow during the basin's
early evolution, although previous tectonic models incorporate a lower heat flow that
decreased to present-day values prior to the Silurian.
In the best-fit model the initial heat flow of 120 mW/m2 was decreased, using an
equation of the form y = A (lIx) + c, until it reached an equilibrium value of 56 mW/m2 at 403
Ma. This equilibrium heat-flow value is within the range of present-day measured values.

Results. The results of the modeling using our heat-flow history can be seen in Figure
9. This graph shows the results of models using: (1) a relatively simple burial history, (2) a
burial history with 2.1 km of excess CarboniferousIPennian deposition only, and (3) our best-
fit burial history with 2.2 km of excess CarboniferousIPermian deposition and 2.3 km of excess
strata at the SilurianlDevonian boundary. This figure in particular demonstrates that the heat-
flow history described here coupled with deposition of excess CarboniferousIPennian and
SilurianlDevonian strata produces an excellent fit to the observed data in both the pre-
Devonian and Devonian and younger sections. Figure 10 shows calculated depth vs. time and
temperature vs. time, respectively, for the best-fit model.

DISCUSSION

Models incorporating constant heat flow through time (Fig. 6), or constant heat flow
and deposition of excess CarboniferousIPermian strata (Fig. 7), could not match the observed
levels of thermal maturity for both the pre-Devonian and Devonian and younger strata in the
Mobil-Jelinek well. Models using heat-flow histories suggested by previous studies (Fig. 8)
could not be made to match observed thermal maturities throughout the well without the
deposition of an extreme amount of material at the SilurianlDevonian boundary. Cercone's
(1984) heat-flow history required the addition of 4 km of material at the SilurianlDevonian
boundary and 1 km of excess CarboniferouslPermian strata to match observed thermal
maturities in the Lower Ordovician and Devonian sections of the well. This model, however,
underestimated maturities in the Upper Ordovician, Silurian, and Carboniferous sections. The
thermal maturities of the Upper Ordovician, Silurian, and Carboniferous sections could be
matched adequately with the addition of 5.5 km of excess deposition at the SilurianlDevonian
boundary and 1.5 km of excess CarboniferouslPermian deposition. This led, however, to an
overestimation of the maturities in the Devonian and Lower Ordovician sections of the well.
Nunn's (1984) heat-flow history required 6.5 km of deposition and erosion at the
SilurianlDevonian boundary and 2.4 km of excess CarboniferousfPermian strata to match
adequately observed thermal maturities throughout the well.
The results of the modeling suggest that elevated heat flow affected the basin until the
end of the Silurian. Our best-fit model also suggests that as much as 2.3 km of material may
198 EVERHAM AND HUNTOON

%Ro
0.1 10
0r------------,-------T--------------------1

(')

'"
~ ~
f • Measured values
(cen:one& PoII_, 1991)

- - Simple burial hIstory

;' to ...- - Burtal history wi 2.1 km of Permo--


ac: ti: Cltl'bonlferouseroslon
<II
~ ------ Best.ftt model results
\\
500 __ ~.
I\.
~

.'
.1:
il
l:
~
.1,
1000
-- ~\ ,,
,
,,
,,

\
,,•
,
1500
\.
\.
\:
-

\.
o
~ \ .\

.
2000 Q. .,:
!!l eo:
.:
\

Figure 9. Semilog plot of %Ro vs. depth using elevated heat flow through Silurian and three different burial
histories. Filled diamonds are measured %Ro values. Solid line represents relatively simple burial history,
dashed line denotes addition of 2.1 km of excess CarboniferousIPermian deposition, and dash-dot line represents
results of best-fit model. Best-fit model required addition of 2.3 km of excess strata at SilurianlDevonian
boundary and 2.2 km of excess CarboniferousIPermian deposition.

have been present at the SilurianlDevonian boundary. Other models could be generated that
incorporate less material at the SilurianlDevonian boundary but would require a higher heat
flow. For example, using 1 km of eroded material at the SilurianlDevonian boundary requires
a heat flow of about 200 m W1m2 to produce acceptable results. The best-fit model presented
here represents the effect of the addition of a moderate amount of excess material at the
SilurianlDevonian boundary while assuming a moderate pre-Devonian heat-flow value. The
SilurianlDevonian boundary represents a major unconfonnity and the transition from the
Tippecanoe to Kaskaskia Sequences (Sloss, 1963), and it is likely that a major erosion event
is represented at that stratigraphic level. Our results support the hypothesis that a significant
THERMAL HISTORY OF THE MICHIGAN BASIN 199

1----=--~-~I\___+:s.L~":__-_J_.ffr---_;s;:;••::::n.:rG;;:ro::;;u.:---t -1.0

~-~~~fI-IIi=~~~-2.0
I
rI -------lIf_-\-...Y.----.l;"--VYHfII-r====;;;=;:~T.:::;;:;::==t
Mt. Simon Sandstone
-2.5 ..
I
I-------If_-~--..l~~\f-lll+---------__t -3.0

1-------I'--------\~HfH---------__t -3.'
....0

.....
550

5SO
500

500
-
45.
<00

.00
350

3SO
300

...
250
TIm_{M_'
250
200

,..
'50

150 ,..
100 50

50
~ ".0

20

Mt. Simon. Sand.tolM


1------~f_---~~~~r---------~'OOu
j
~-----_lf_----~~~---------~120~

1--------1f-----------------~ 14.
1-_____-1_________________1'8.

1-------I----------------r==~1~
L-_ _ _ _ _ ___________________ [!J

Figure 10_ Graphs of depth vs_ time (A), and temperature vs. time (B) for best-fit model. Burial history is
summarized in Table 2. Horizons monitored are from deepest to shallowest and coolest: Mt. Simon Sandstone,
Prairie Du Chein Group, Black River Group, Trenton Group, CincDmatilm Undiff. Rocks, Niagara Group, Salina
Group, Bass Island Group, Dundee Limestone, Tranverse Group, Berea sandstone, Coldwater Shale, Michigan
Fonnation, and Saginaw Fonnation.

amount of Pennsylvanian and Permian strata were at one time present in the basin. Our best-fit
model requires 2.2 Ian of Upper Carboniferous and Permian sediment to have been eroded
during the Triassic to match the observed thermal maturities in the Devonian and later strata.

CONCLUSIONS

The results of this study support the following conclusions:


(1) Thermal maturities within the Mobil-Jelinek well are inconsistent with a thermal
history characterized by constant heat flow through time.
200 EVERHAM AND HUNTOON

(2) A period of elevated heat flow affected the basin during its early history and
probably continued through the Silurian.
(3) A relatively complex burial history is necessary to account for observed levels of
thermal maturity in the basin.
(4) Approximately 2 km of Late Carboniferous and Permian strata were once present
in the basin and were eroded later, probably during the Triassic.
(5) In the absence of extremely high-heat flow, about 2 km of excess strata at the
SilurianlDevonian boundary is required to produce observed levels of maturity in
the pre-Devonian section.

ACKNOWLEDGMENTS

This project was funded by the U.S. Department of Energy, contract # DE-FC22-
94BC14983. Special thanks to GEOGRAPmx, Inc. and their University Program for providing
the GEOGRAPHIX Exploration Systems © software used in part of this study. Thanks also to
the Subsurface Visualization Lab, Department of Geological Engineering and Sciences,
Michigan Technological University and its director James R. Wood for providing computer
facilities.

REFERENCES

Aangstrom Precision Corporation, 1990, Michigan Oil and Gas Well Database.
AAPG, 1985, Midwestern basin and arches region, Correlation of stratigraphic units of North America
(COSUNA) project. Lindberg, F.A. editor.
Allen, P.A., and Allen, J.R.,1990, Basin analysis: principles and applications: Blackwell Scientific Pub!.:
London. 451 p.
Alpern, B., 1980, Petrographie du kerogene, in Durand, B., ed., Kerogen, Insoluble Organic Matter from
Sedimentary Rocks: Editions Technip, Paris, p. 339-384.
Baldwin, B., 1971, Ways of deciphering compacted sediments: Jour. Sed. Pet., v. 41, no. I, p. 293-310.
Baldwin, B., and Butler, C.O., 1985, Compaction curves: Am. Assoc. Petroleum Geologists Bull., v.
69, no. 4, p. 622-626.
Beaumont, C., 1978, The evolution of sedimentary basins on a viscoelastic lithosphere: theory and examples:
Geophys. Jour. Roy. Astr. Soc., v. 55, no. 2, p. 471-497.
Bertrand, R, 1990, Correlations among the reflectances of vitrinite, chitinozoans, graptolites and scolecodonts:
Organic Geochem., v. 15, no. 6, p. 565-574.
Bertrand, R., and Heroux, Y., 1987, Chitinozoan, grapolite, and scolecodont reflectance as an alternative to
vitrinite and pyrobitumen reflectance in Ordovician and Silurian strata, Anticosti Island, Quebec,
Canada: Am. Assoc. Petroleum Geologists Bull., v. 71, no. 8, p. 951-957.
Bertrand, R., Berube, J.-C, Heroux, Y., and Achab, A., 1985, Petrographie du kerogene dans Ie Paleozoique
inferieur: methode de preparation et exemple d'application: Revue de l'Institut Francais du Petrole, v.
40, no. 2, p. 155-167.
Blackwell, D.D., and Steele, J.L., 1989, Thermal conductivity of sedimentary rocks: measurement and
significance, in Naesar, N.D., and McCulloch, T.H., eds., Thermal History of Sedimentary Basins:
Methods and Case Histories: Springer-Verlag, New York, p. l3-36.
Bowers, T.L., 1989, Upper Niagara-Lower Salina (Mid-Silurian) sedimentology, and conodont-based
biostratigraphy and thermal maturation of the southeast Michigan Basin: unpub!' masters thesis, Wayne
State Univ., 119 p.
Catacosinos, P.A., Daniels, P.A., Jr., and Harrison, W.B., III, 1991, Structure, stratigraphy and petroleum geology
of the Michigan Basin, in Leighton, M.W., Kolata, D.R. Oltz, D.F., and Eidel, J.J., eds., Interior
Cratonic Basins: Am. Assoc. Petroleum Geologists Mem. 51, p. 561-602.
THERMAL mSTORY OF THE MICmGAN BASIN 201

Cercone, K.R., 1984, Thennal history of Michigan Basin: Am. Assoc. Petroleum Geologists Bull., v. 68, no. 2,
p.130-136.
Cercone, K.R., and Pollack, H.N., 1991, Thermal maturity of the Michigan Basin: Geo!. Soc. America Spec.
Paper 256, p. 1-11.
Cercone, K.R., Deming, D., an4 Pollack, H.N., 1996, Insulating effect of coals and black shales in the
Appalachian Basin, westem Pennsylvania: Organic Geochem., v. 24, no. 2, p. 243-249.
Cermak, V., and Rybach, L., 1982, Thermal properties, in Hellwege, K.-H., and Angenheister, G., eds., Landolt-
Bomstein numerical data and functional relationships in science and technology, new series, group V:
Geophys. Space Res., v. 4, p. 305-371.
Combs, J.C., and Simmons, G., 1973, Terrestrial heat flow determinations in the north-central United States:
Jour. Geophys. Res., v. 78, no. 2, p. 441-461.
Crowley, K.D., 1991, Thermal history of Michigan Basin and southem Canadian Shield from apatitie fIssion
track analysis: Jour. Geophys. Res., v. 96, no. Bl, p. 697-711.
Daly, A.R., and Lilly, D.H., 1985, Thermal subsidence and generation of hydrocarbons in Michigan Basin:
Discussion: Am. Assoc. Petroleum Geologists Bull.,v. 69, no. 7, p. 1181-1184.
Deming, D., Nunn, J.A., Jones, S., and Chapman, D.S., 1990, Some problems in thermal history studies, in
Nuccio, V.F., Barker, C.E., Dyson, S.J., eds., Applications of Thermal Maturity Studies to Energy
Exploration: Eastwood Print. and Pub!., Denver (SEPM), p. 61-80.
Dow, W.G., 1977, Kerogen studies and geological interpretations: Jour. Geochem. Exploration, v. 7, no. 2, p.
79-99.
Drury, M.J., Allen, V.S., and Jessop, AM., 1984, The measurement of thermal diffusivity of rock cores:
Tectonophysics, v. 103, no. 1-4, p. 321-333.
Fisher, J.H., Barratt, M.W., Droste, J.B., and Shaver, R.H., 1988, Michigan Basin, in Sloss, L.L., ed.,
Sedimentary Cover North American Craton: U.S.: Geo!. Soc. America, Geo!. of North America, v. D-2,
p.361-382.
Furlong, K.P., and Edman, J.D., 1989, Hydrocarbon maturation in thrust belts: thermal consideration, in Price,
RA, ed., Origin and Evolution of Sedimentary Basins and Their Energy and Mineral Resources: Am.
Geophys. Union Geophys. Mon. 48, p. 137-144.
Gardner, W.C., and Bray, E.E., 1984, Oils in source rocks of the Niagaran reefs (Silurian) in the Michigan Basin,
in Palacas, J.G., ed., Petroleum Geochemistry and Source Rocks: Am. Assoc. Petroleum Geologists
Studies in Geology no. 18, p. 33-44.
Haxby, W.F., Turcotte, D.L., and Bird, J.M., 1976, Thermal and mechanical evolution of the Michigan Basin:
Tectonophysics, v. 36, no. 1-3, p. 57-75.
Hogarth, C.G., and Sibley, D.F., 1985, Thermal history of the Michigan Basin: evidence from conodont color
alteration indices, in Cercone, K.R. and Budai, J.M., eds., Ordovician and Silurian rocks of the
Michigan Basin and its margins: Michigan Basin Geo!. Soc. Spec. Paper 4, p. 45-57.
Howell, P.D., 1993, Styles of subsidence in the Michigan Basin: unpub!' doctoral dissertation, Univ. of
Michigan, 276 p.
Howell, P.D., and van der Pluijrn, B.A., 1990, Early history of the Michigan Basin: subsidence and Appalachian
tectonics: Geology, v. 18, no. 12, p. 1195-1198.
Hunt, J.M., Lewan, M.D., and Hennet, R. J-C., 1991, Modeling oil generation with time-temperature index
graphs based on Arrhenius equation: Am. Assoc. Petroleum Geologists Bull., v. 75, no. 4, p. 795-807.
Huntoon, J.E., 1990, An integrated model of tectonics and sedimentation for the Newark basin: unpub!' doctoral
dissertation, Pennsylvania State Univ., 346 p.
Huntoon, J.E. and Furlong, K.P., 1992, Thermal evolution of the Newark Basin: Jour. Geology, v. 100, no. 5,
p.579-591.
Illich, HA, and Grizzle, P.L., 1983, Comment on "Comparison of Michigan Basin crude oils" by Vogler et a!.
Geochirnica et Cosmochirnica Acta, v. 47, no. 6, p. 1151-1155.
Illich, H.A, and Grizzle, P.L., 1985, Thermal subsidence and generation of hydrocarbons in Michigan Basin:
Discussion: Am. Assoc. Petroleum Geologists Bull., v. 69, no. 9, p. 1401-1403.
Lilienthal, R.T., 1978, Stratigraphic cross-sections of the Michigan Basin: 'Michigan Geo!. Survey Rept. of
Invest. 19,27 p.
Lopatin, N.V., 1971, Temperature and geologic time as factors in coalifIcation (in Russian): Akaderniya Nauk
SSSR Izvestiya, Seriya Geologicheskaya, no. 3, p. 95-106.
Mansure, AJ., and Reiter, M., 1979, A vertical groundwater movement correction for heat flow: Jour. Geophy.
Res., v. 84, no. B7, p. 3490-3496.
202 EVERHAM AND HUNTOON

McCulloh, Th.H. and Naesar, N.D., 1989, Thermal history of sedimentary basins: introduction and overview,
in Naesar, N.D., and McCulloch, T.H., eds., Thermal History of Sedimentary Basins: Methods and Case
Histories, Springer-Verlag, New York, 319 p.
Mongelli, F., Loddo, L., and Tramacere, A., 1982, Thermal conductivity, diffusivity and specific heat variation
of some Travale Field (Tuscany) rocks versus temperature: Tectonophysics, v. 83, no. 1, p. 33-43.
Moyer, R.B., 1982, Thermal maturity and organic content of selected Paleozoic formations-Michigan Basin:
unpubl. masters thesis: Michigan State Univ., 62 p.
Nunn, J.A., 1994, Free thermal convection beneath intracratonic basins: thermal and subsidence effects: Basin
Research, v. 6, no. 2/3, p. 115-130.
Nunn, J.A., and Sleep, N.H., 1984, Thermal contraction and flexure of intracratonal basins: a 3-d study of
Michigan Basin: Geophysical Jour. Roy. Astr. Soc., v. 76, no. 3, p. 587.
Nunn, J.A., Sleep, N.H., and Moore, W.E., 1984, Thermal subsidence and generation of hydrocarbons in
Michigan Basin: Am. Assoc. Petroleum Geologists Bull., v. 68, no.3, p. 296-315
Price, K.L., Huntoon, IE., and McDowell, S.D., 1996, Thermal history of the 1.I-Ga Nonesuch Formation, North
American mid-continent rift, White Pine, Michigan: Am. Assoc. Petroleum Geologists Bull., v. 80, no.
1, p. 1-15.
Pruitt, J.D., 1983, Comment on "Comparison of Michigan Basin crude oils· by Vogler et al.: Geochirnica et
Cosmochirnica Acta, v. 47, no. 6, p. 1151-1155.
Quinlan, G.M., and Beaumont, C., 1984, Appalachian thrusting, lithospheric flexure, and the Paleozoic
stratigraphy of the Eastern interior of North America: Can. Jour. Earth Science, v. 21, no. 9, p. 973-996.
Robertson, E.C., 1988, Thermal properties of rocks: U.S. Geol.Survey, Open-File Rept. 88-441, 106 p.
Shaver, R.H., Sunderman, J.A., Mikulic, D.G., Kluessendorf, J., McGovney, J.E.E., and Pray, L.C., 1983,
Silurian reef and inter-reef strata as responses to a cyclical succession of environments, southern Great
Lakes area (Field Trip 12), in Field trips in midwestern geology: Geol. Soc. America, Indiana Geol. Soc.
and Indiana Univ., Dept. of Geology, Bloomington, Indiana, v. 1, p. 141-196.
Sleep, N.H., 1996, Lateral flow of hot plume material ponded at sublithospheric depths: Jour. Geophys. Res., v.
101, no. B12, p. 28,065-28,083.
Sleep. N.H., 1997, Lateral flow and ponding of starting plume material: Jour. Geophys. Res. v. 102, no. B5, p.
10,001-10,012.
Sloss, L.L., 1963, Sequences in the cratonic interior of North America: Geol. Soc. America Bull., v. 74, no. 2,
p.93-114.
Sloss, L.L., 1988, Tectonic evolution of the craton in Phanerozoic time, in Sloss, L.L., ed., Sedimentary Cover
North American Craton: U.S.: Geol. Soc. America Geol. of North America, v. D-2, p. 25-51.
Sweeney, lJ., and Burnharn, A.K., 1990, Evaluation of a simple model of vitrinite reflectance based on chemical
kinetics: Am. Assoc. Petroleum Geologists Bull., v. 74, no. 10, p. 1559-1570.
Tissot, B., and Espitalie, J., 1975, L'evolution therrnique de la matiere organique des sediments: applications
d'une simulation mathematique: Revue del'lnstitut Francais du Petrole, v. 30, no. 5, p. 743-777.
Vail, P.R., Mitchum, R.M. Jr., and Thompson, S. m, 1977, Seismic stratigraphy and global change of sea level,
Part 4; Global cycles of relative changes of sea level, in Payton, C.E., ed., Seismic
Stratigraphy-Applications to Hydrocarbon Exploration: Am. Assoc. Petroleum Geologists Mem. 26,
p.83-97.
Wang, H.F., Crowley, K.D., and Nadon, G.C., 1994, Thermal history of the Michigan Basin from apatite fission-
track analysis and vitrinite reflectance: Am. Assoc. Petroleum Geologists Mem. 61, p. 167-177.
Waples, D.W., 1980, Time and temperature in petroleum formation: application of Lopatin's Method to
petroleum exploration: Am. Assoc. Petroleum Geologists Bull., v. 64, no. 6, p. 916-926.
Waples, D.W., 1985, Geochemistry in petroleum exploration: International Human Resources Development
Corporation, Boston, 232 p.
Wilson, J.L., 1975, Carbonate facies in geologic history: Springer-Verlag, New York, 471 p.
RISING MUD DIAPIRS AND THEIR THERMAL ANOMALIES

E. Bagirovl and I. Lerche2

IGeological Institute
Azerbaijan Academy of Sciences
Baku, Azerbaijan
2Department of Geological Sciences
University of South Carolina
Columbia, SC 29208

ABSTRACT

Mud diapirs in the South Caspian Basin are noted for being massively gas-charged, for
having extremely low temperatures compared to regional values, for having cold to ice-cold
frothy mud and water emissions from gryphon!! and salses despite originating from several
kilometers depth (as judged by mineral contents ofthe waters), and for having inferred heat-
flux values which can be an order or magnitude or more higher than those inferred regionally.
Using both a self-consistent geometric evolution code for diapirs and sediments,
together with a dynamical evolution code, GEOPETII, for sediment basins, this paper accounts
for the observations in terms of a low thermal conductivity of mud diapirs relative to regional
values. The consequence is to cause heat to defocus away from the diapiric base, thereby
causing a higher temperature than in sediments regionally at the same depth, and to cause heat
to focus towards the diapiric crest thereby providing a low temperature compared to regional
ones. For the Abikh diapir the crestal domain isotherms can be about 2-3 km deeper than
regionally. In addition, the higher diapiric basal temperature promotes more rapid conversion
of kerogen to oil, and of oil to gas, which not only aid in the gas-charging of the mud diapir
and which lowers the thermal conductivity even farther, but which also contributes to the
buoyant rise of the mud diapir in the last 5-10 million years, coincident with massive
sedimentation of about 10 km in the South Caspian Basin during the same time period.
The dynamical basin analysis model, GEOPETII, also includes exsolution of gas from
oil and water at low temperatures and pressures, which cools the diapir even more by direct
adiabatic gas expansion and by gas-charging of crestal pore space, which again lowers the
diapiric thermal conductivity.
Numerical estimates indicate that the thermal influence of the Abikh diapir extends
laterally about 2-4 diapir diameters into surrounding sediments near the crest. Near the
diapiric base the thermal influence on hydrocarbon generation is profound, whereas sediment
formation upturning distortions occur out to about six diapir diameters for all formations from

203
204 BAGIROV AND LERCHE

Cretaceous through to Holocene, indicating a major influence of the diapir on gas and oil
accumulations in and around the diapir.

INTRODUCTION

Mud diapirs occur in a number of modern accretionary wedges (Brown and Westbrook,
1988) and provide significant contributions to structural development, mass motion and heat
transportation, fluid pressure development, sedimentary bed distortions, and hydrocarbon
generation, migration, and accumulation. A complete understanding of the dynamical
development of mud diapirs and their associated driving forces is far from clear under different
geological conditions (Brown, 1990; Ivanov and Guliev, 1986). Here, we concentrate on
providing an understanding of just the thermal anomalies of evolving mud diapirs, with
particular emphasis on the Abikh mud diapir in the offshore South Caspian Basin as an
archetypical illustration.
Observations in the South Caspian Basin, a classical region of mud diapirs and mud
volcanoes with approximately 300 known, show systematic thermal anomalies around mud
volcanoes relative to regional conditions far from the structures. Heat fluxes inferred from
measurements are typically one to two orders of magnitude higher than in surrounding
sediments (Ivanov and Guliev, 1986), yet measured temperatures of mud flows and of waters
associated with salses and gryphons from mud volcanoes, are typically lower than surficial
sedimentary temperatures - the mud flows and waters literally may be ice-cold compared to
sediment temperatures of about to-20°C (direct measurements by the authors in July 1996;
Ivanov and Guliev, 1986). The question addressed here is to account for these observations
at the present day, and also to tie thermal anomalies in the past to the evolution of mud diapirs,
so that an understanding is achieved of the influence with time of such thermal anomalies on
hydrocarbon generation.
Elsewhere (Lerche and others, 1996), the time-development has been examined of both
thermal anomalies, deformation and stress around the growing Vezhirov mud diapir in the
South Caspian Basin using a self-consistent procedure described in more detail in Lerche and
Petersen (1995) and Lerche and others (1996). That procedure, however, is concerned with
only the local geometric distortion of evolving diapirs and surrounding sediments and does not
allow for regional effects including excess pressure and fluid-flow development. Here, in a
novel approach, the Abikh mud diapir is investigated using both the local geometric self-
consistent method of Lerche and Petersen (1995) and, in addition, using a 2-D basin analysis
code, GEOPETII, which does allow for regional development of fluid-flow factors. In this
way thermal and dynamical effects on both the local scale of a diapir and on a broader basinal
scale are exhibited to the sharpest extent possible.

PRESENT-DAY SHAPE AND DIAPIR EVOLUTION

Diapir Geometry

The Abikh mud diapir is located in the central deep-water part of the South Caspian
Basin (Bagirov and Lerche, 1997a; figs. 2 and 3 ofNadirov and others, 1996). A seismic
section of 12 seconds two-way travel time crosses the Abikh structure from east to west
(Gambarov and others, 1993; Lerche and others, 1996). The shape of the diapir on the seismic
MUD DIAPIRS AND THERMAL ANOMALIES 205

section is complex (Fig. 1) and penetrates all Meso-Cenozoic sedimentary fonnations present.
The diapiric roots extent to the full 12 seconds of the seismic section, corresponding to depths
of 22-26 kIn (Gambarov and others, 1993), whereas the crest of the diapir almost reaches sea
bottom; the diapiric width is approximately 10 kIn.
E
0

4
u-
! s
:c 6
>-
Cl.
W
D 7

10

11

12

Figure I. South Caspian Basin Abikh mud diapir on 12-second two-way travel-time west-east cross section;
location of seismic line as given in figures 2 and 3 ofNadirov and others (1996). Horizon symbols are: (i) Q
" Quaternary; (ii) N 1 2Upper Pliocene; (iii) N" _Middle Pliocene; (iv) N" _Lower Pliocene; (v) Mi "_Miocene;
(vi) ~ = Maikop (Oligocene-Lower Miocene); (vii) P2 "_ Eocene; (viii) PI "_ Paleocene; and (ix) Mz
"_Mesozoic.

The right (east) side has a wall-type diapir shape, with a rim syncline at the base ofthe
diapiric stem; sediment layers are upturned by the rising diapir. The left (west) side of the
diapir is more complex, with an overhang near the crest and a "bulge" on the stem.
Sedimentary fonnations, correspondingly, are more disturbed on this side. Depths of the same
stratigraphic units on the left and right sides of the diapir can be 2-6 kIn different. Without
additional geological and geophysical data, it is not clear if that difference is caused only by
the diapir, or if the difference signals some other major geological or tectonic processes.

Local Geometric Diapir Evolution

The self-consistent model of diapir and surrounding sediment evolution (Lerche and
Petersen, 1995) operates in two parts. First, the observed diapir shape is used to detennine
values for present-day parameters in a suite of equations describing different possible diapiric
structures, a pillow (mound) structure; a vertical mud diapir (wall); a diapir with a developed
overhang; and a "tear-drop" diapir with negligible physical connection to any underlying
mother fonnation. Second, once the present-day parameters are detennined so that a minimal
mismatch is obtained between predicted present-day diapir shape and observed shape, then the
paleovalues of the parameters are allowed to differ away from the present-day values in order
to describe the diapir evolution. Constraint infonnation used to control such an evolution are
all the present-day sedimentary bed configurations relative to assumed horizontal deposition
of all beds. Thus the paleovalues of parameters are adjusted until a minimum mismatch is
206 BAGIROV AND LERCHE

achieved between all present-day predicted sedimentary bed locations surrounding the diapir.
In this way one has a self-consistent evolution description of both the diapir and the
surrounding sedimentary beds. For both the parameters describing the present-day diapir
shape, and their paleocounterparts describing sedimentary bed distortions from horizontal as
a consequence of diapir evolution, a systematic inverse procedure is used. Details of this two-
stage method and the inverse procedure are given in Lerche and Petersen (1995).
Using the time to depth conversion of Gambarov and others (1993), the present-day
shape of the Abikh diapir in physical depth as well as the corresponding sedimentary beds are
shown in Figure 2. From the observed diapir shape alone, Figure 3 shows the detailed model
fit to the diapir shape with the "best" present-day parameter values (the first-stage of
development of the local self-consistent model).

10
--------------~~--­
Mi+N21
'--------------~~---
Mp

::c 15
b:
w
o Mz
20

25

30 25 20
Iii iii iii iii i I (. i , j iii,. Iii i ,
15 10 10 15 20 25

Horizontal distance (km)

Figure 2. Present-day Abikh diapir shape with depth converted from seismic section using time to depth
conversion of Gambarov and others (1993).

30
30

25
25

20 20

.[ 15
~ 15

J
~
:l 10
10
.. ~

A
B
o+r~~~~~~~~~~~~~
o to 15 20 25 30 35 to 15 20 25 30 35
Horilon1a1 dl$tance (km) Horizonlal distance (un)

Figure 3. Visual comparison of predicted numerical (dashed line) and observed (solid line) present-day Abikh
diapir shape. A, Right side of diapir; B, Left side of diapir.
MUD DIAPIRS AND THERMAL ANOMALIES 207

For the second-stage of development ofthe local self-consistent model, the present-day
sedimentary bed configurations, plus the assumed horizontality of the beds at deposition, both
are used to constrain the paleovalues of parameters. Ahead of detailed quantitative modeling
it is noted that present-day beds provide clues to the probable evolution of the Abikh mud
structure, particularly the diapiric phase, which occurs when there is primarily vertical flow
of mud associated with the breakthrough, distortion, and penetration of the then overlying
formations. The source of mud is mainly from the domain close to the diapir. During the
evolution, the mud thickness below the secondary rim syncline decreases, the sediment load
increases, and so the pressure gradient on the mobile mud is amplified. In turn this gradient
enhances the flow of mud into the diapiric stem and so creates a larger rim syncline, thereby
amplifying the process even further.
Continued self-amplification can deplete the mobile mud supply. This phase of
evolution may be characterized by a thickening of sediments towards the incipient diapir
structure. However, some care has to be attached to using this effect in a general sense,
because mobile mud supply can be not only from formations below the rim syncline but also
from adjoining formations where sediments are undercompacted. In the local geometric self-
consistent model here both effects are included in the evolution, as shown by the
nonconservation of area of the evolving mud diapir.
Thus, qualitative observations of Figure 2 already provide some guidance in evaluating
quantitative model needs. With the two controls of (i) horizontal beds at deposition; and (ii)
present-day sedimentary formation geometries, the "best" paleoparameters were determined
using the systematic inverse procedure for self-consistent evolution of the Abikh diapir and
surrounding sedimentary formations.
The results of this local dynamic evolution modeling are shown in Figure 4, indicating
that before about Early Pliocene time, the Abikh diapir was not present. Diapiric rise began
in Middle Pliocene and by roughly Quaternary time the diapir had reached its maximum
height, close to present-day values. This quantitative modeling of the Abikh diapir evolution
and surrounding sedimentary formations is consistent in both timing and formation uplift with
estimates made elsewhere (Guliev, 1996; Ivanov and Guliev, 1986; Nadirov and others, 1996)
of: (i) massive sediment supply (~1-2 kmlmillion years) during the last 5-10 million years;
(ii) timing, originating depths, and volumes of produced hydrocarbons, particularly gas
generation which is thought to be the dominant buoyancy driving force of South Caspian Basin
mud diapirs (Ivanov and Guliev, 1986; Guliev, 1996); and (iii) gas flaming from mud diapirs
and mud volcanoes using a physico-chemical model (Guliev, 1996). Taken together, these
consistent indicators suggest that the dominant aspects of evolution of the Abikh diapir are
captured correctly by the local geometric self-consistent model of diapir and sediment
evolution and distortion.
A diapiric rise initiation some 3-5 million years ago, and a present-day observed
diapiric height of some 22-26 km, indicate a sustained average vertical speed of diapiric rise
of some 5-8 kmlmillion years - between two to three times as rapid as the sedimentation rate
since Middle Pliocene. If Darcy's law is appropriate to describe this diapiric rise phase, then
the corresponding mud permeability is about 10.2 mD, making for a "tight" seal capable of
retaining generated and entrapped hydrocarbon gas - a point that had been noted previously
as necessary by Guliev (1986) in his physico-chemical model of gas flaming from mud
volcanoes.
Consideration of regional modeling of the seismic section using the basin analysis
code, GEOPETII, will be taken up after the local thermal influence of the Abikh diapir is
evaluated using the geometric self-consistent model.
208 BAGIROV AND LERCHE

a
20
M,.N;
25 M,
,5 P'+P2
...E M,
20
i '0
~

'5

M. 0
1-5.0 MYBP

I ':j "~_ A
'0
Horizontal
'5 20
distance (1cm)
25 30 3,

o [i I , • I i •• , I ' , , i i • ,
.[. " 'i'" I ['" .[
o 5 10 15 20 25 30 3S
Horizontal distance (km)
E
,5 ~
P,...P2

..-------------
________ N~

I
"-
'0
lA,

25' ~

t-16.0MY8P E
20
0
0 S '0 '5 20 25 30
'5
! Horizontal .1S1anco(l<m)
M,

J '0
,5

I P.+Pz
B
J
M,
t.l.65MYBP '0

F
Horizontal distance (km) t_35.5MYBP

25 '0 20 25 30
'5
Horizontal distance pan)

'5
M, E
"-
IA.
.~ '0
~

c
G
iii, i I' i ',1 I i i i' i • iii •• [
t =65 MYBP
5 10 15 20 25 30 3S
'0 15 20 2S 30 3'
Honzonlal distance (km)
Horizontal distance (Icrn)

Figure 4. Evolution of right side of Abikh diapir and sediment formation geometries: A, t=O (present-day); B,
t=1.6 MYBP; C, t=2.8 MYBP; D, t=5.0 MYBP; E, t=16.0 MYBP; F, t=35.4 MYBP; G, t=65.0 MYBP.
Abbreviations for stratigraphic units as for Figure 1.

THERMAL EFFECTS OF THE ABIKH DIAPIR: PRESENT AND PAST

The majority of mud volcanoes are known to be massively gas-charged through direct
observations of gas eruptions, flame ignition (up to 1.2 kIn flame height), and through gas
hydrate measurements (Bagirov and Lerche, 1996; Guliev, 1996), and indirectly through low
seismic stacking velocities recorded to considerable depth (-20 kIn) in mud diapirs (Gambarov
and others, 1963).
MUD DIAPIRS AND THERMAL ANOMALIES 209

So to evaluate the local thennal influence of a mud diapir, it is taken that the diapir is
gas-charged. Now gas-saturated mud has a thennal conductivity which is about one-half to
one-third that of water-saturated sediments (Ivanov and Guliev, 1986); typical values in near-
surface mud diapirs are about 0.6 W/mK and 2.0 W/mK, respectively.
Mud diapirs in the South Caspian Basin exhibit large vertical structural fonns
compared to their widths, so that their high thennal resistance (low thennal conductivity)
compared to neighboring water-charged sediments indicates that heat will tend to avoid an
uprising mud diapir to the extent possible as set by basinal conditions. Accordingly, sediments
close to the crest of a mud diapir will be cooler than sediments at the same depth but far from
the diapir, whereas sediments close to the base of a mud diapir, and in the rim syncline, will
be wanner than equal depth sediments far removed from the diapir - precisely the opposite
effect to that occurring for a salt diapir which has a thennal conductivity about three times
higher than sediments.
To illustrate the focusing and defocusing of heat and the corresponding spatial
variations of temperature, three groups of calculations are shown here (although many more
have been done, all yielding similar results). The first two groups use the local geometric
model described earlier with gas-charged mud thennal conductivity taken to be either constant
or to increase with increasing temperature (depth). And only present-day thennal patterns are
exhibited here. The second group of calculations uses the two-dimensional basin analysis
code, GEOPETII, to provide a more regionally relevant picture ofthennal evolution around
the developing Abikh diapir.
Consider each group in tum.

Local Thermal Influence

Under steady-state conditions, and with a basement heat flux of34mW/m 2 (=0.8 HFU),
typical of measured values in the South Caspian Basin (Bredehoeft, Djevanshr, and Belitz,
1988), the present-day temperature pattern is presented in Figure 5. The extreme end-member
situation of constant thennal conductivities for both the mud and surrounding sediments is
where the sediment surface is an isothenn at 9°C because of the temperature of the overlying
water. Note the isothenn pattern shows defocusing of heat near the diapiric base and focusing
towards the diapiric crest. The temperature difference laterally is large: at 20-25 kIn depth the
temperature near the center of the diapir is about 50°C higher than in sediments at the same
depth but 35 kIn laterally distant from the diapir. Near the crest of the diapir the opposite
behavior is noted. Isothenns are depressed to a depth of approximately 3 kIn in the diapir and
recover to their regional pattern at about a diapir diameter or so laterally removed, that is over
a lateral scale of about 10-20 kIn.
As a second extreme end-member, the thennal conductivity of the gas-charged mud
is taken to increase with depth, gradually reaching the thennal conductivity of the surrounding
sediments. The argument in favor of such a behavior is that as the gas density reaches 0.1-0.2
gcm·3, it behaves more similar to a quasiliquid than a compressible gas. Thus, the gas-charged
mud thennal conductivity increases. For illustration this increase has been taken to occur at
a temperature scale of 100°C, corresponding to about 5-10 kIn of burial which, under
conditions prevailing in the South Caspian Basin is where the pressure is between 500-1000
atm, enough to compress gas to the "quasi liquid" state. The same steady-state temperature
calculations as in the preceding example then were run, all other factors being equal, with the
present-day temperature pattern being exhibited in Figure 6. Again, the crestal regions around
the diapir show the depression of isothenns to about 3 kIn depth at the diapiric center, with
210 BAGIROV AND LERCHE

recovery to the regional pattern for a 10-20 km lateral scale away from the diapir. However,
because the mud diapir and sediment thennal conductivities are the same at great depth in this
example, there is no longer the enhanced temperature regime in the region around the base of
the diapir.

°ll-===~~~~~~=======--'
---------.• --------------------- - .... - .. ---.5O'"C'''- ---.... ---- .. ---------------------------

5 --- 7J-C--- -------------------------------------


I
___------------ _____ ______ 'j-C O-- ~-------------------------~--
E 10
------------------~~- --- ---------.. ,2 i-c--- -- --------------..:-
~
_____________ ----______ _ _______________ I'OC·-- - ______________ _
ii: 15

;- ;:~ :~: ~:~ ~ ~-~=:L:~:~!/!~: !~ ~ L:- -~:~-~:~:~ :~ ~ ;


UJ
o
20

25

30

Figure 5. Present-day isothenns for constant thermal conductivity of mud.

0
-_
.. ---_ ............ ---- ........ ---- ---_ .. --
5
--
I
-- 5O"C---
___________________________ _
--- - ---- 75OC--- -
_______ ,lco-- ----""--~=-----_I
E 10
'2~"C·--
~ ___________ ----------------------------

__ I
:c
b: 15
UJ
0 ~~--------:-~---------::--------------~ ~~- ~~~~~~~~~~~~J~:~-= -~~~------------------------
20 I

25
:~~:~:m:::=::::-:-:::=::::::::::::~F=::--:=::::::::;:5k;;;-----
_______________________________________ - - - - - - o2So"C - - -- ---________________________ _

30

Figure 6. Present-day isothenns for thermal conductivity of mud increasing with temperature.

Regional and Temporal Thermal Influence

Although the general pattern ofisothenn depression (and consequent heat focusing)
in the vicinity of the mud diapiric crest is portrayed simply by the two examples, there are
several simplifications in the calculations that make it desirable to present a more appropriate
description. Some of these concerns are tied to the geometric self-consistent model and others
to the thermal model used. Thus, the model of the evolving thennal regime was taken to be
steady-state with the sediment thennal conductivity held fixed and with heat transport only by
conduction. Further, the self-consistent diapir and sediments geometric model used as a basis
for determining the spatial regions where mud and sediment have separate thermal properties
did not allow for bed compaction and fluid flow (and hence increasing thennalconductivity
MUD DIAPIRS AND THERMAL ANOMALIES 211

with burial), nor for transient transport of heat by convection, or even of the influence of the
increasing sediment load on basinal distortion. Thus, the modeled behaviors of dynamic
evolution. and associated thermal evolution around a rising mud diapir are not completely
unassailable. However, it would certainly seem that the correct orders of magnitude for both
the dynamical mud diapir evolution and the diapir crestal thermal depression are captured by
the simple self-consistent local model. And yet, there is the concern that the models used may
be too restrictive and may not be producing the relevant behavior when allowance is made for
departures from the confining model assumptions.
To address such potential concerns a more sophisticated basin analysis code,
GEOPETII, has been used to evaluate the combined dynamical and thermal evolution of an
uprising mud diapir. The code, GEOPETII, allows lateral as well as vertical flow of three
separate phases (water, oil, gas) simultaneously. Also included are: variable deposition rates
laterally in a basin, unconformities, stratigraphic' variations laterally, and paleooverpressuring.
The code combines conductive and convective heat transfer together with the fluid flow. The
model includes thermal expansion pressure and capillary pressure. An N-channel hydrocarbon
generation/migration model has been added which allows separate oil and gas phase flows, as
well as a fracture model, a cementation/dissolution routine, and paleotectonic, and paleowater
depth modules. Solubility of gas in oil, oil in water, and gas in water are included. Also the
oil generation is broken down into light, medium, and heavy fractions. In addition, the model
will allow up to 10 vertical or normal slant faults, each of arbitrary throw (up or down), the
faults may be open or closed at user discretion for different time periods and each fault may
extend across an arbitrary number of horizons.
A detailed description of the operation of the GEOPETII code has been given
elsewhere (Lerche, 1990) and there is no need to repeat that description here. The basic input
data for evaluating the Abikh mud diapir evolution are obtained from well information,
outcrops, as extrapolated to the 12-second cross section in the manner ofNadirov and others
(1996) and Tagiyev and others (1996). Eight seismic-stratigraphic units (SSU), confined to
different age sediments of the South Caspian Basin were picked on the regional 12-second
seismic section; a part of this section is shown in Figure 1.
Lithological input data, formation ages, and total organic carbon (TOC) content were
evaluated (in Chapter 3 of Lerche and others, 1996) for use on a 80-km wide part of the section
centered on the Abikh diapir; a precis of the dominant inputs is given in Table 1, with more
detailed information available in Lerche and others (1996).
The present-day depth values for both basement and for six sediment layer surfaces
[Mesozoic, Eocene, Oligocene-Miocene (the Maikop of the South Caspian Basin), Middle
Pliocene (Productive Series), Upper Pliocene, and Quaternary] were determined in 25
pseudowells along 80 km of the 12-second seismic section, using the depth-time conversion
of Gambarov and others (1993). These SSU then were divided into finer thickness units in
accord with stratigraphic and lithologic information from both outcrops and borehole
information. The model lithologies were tied to (i) lithology of the same age layers from
outcrops and South Caspian Basin wells; (ii) breccia fragments from mud volcanoes; and (iii)
trends in lithologic variations inferred in the basin abyssal direction from wells. This
procedure led to 26 layers being identified in the section with prescribed lithologies as
exhibited in Table 1.
For the mud diapir, the lithology was defined as "shaly" with the same surface
depositional porosity, permeability, thermal conductivity, and density as shale sediments.
However, for modeling purposes, the porosity of the mud is taken to change but little with
depth of burial, allowing one to insert a "sloppy," unconsolidated mud, with all other features
212 BAGIROV AND LERCHE

Table 1. Some lithologic, heat flux, and kerogen inputs for GEOPETII calculations.

LAYER Kerogen Type %


Index Symbol Age, HFU TOC% I Lithology (%)
MYBP II
26 QUAT 1.3 0.75 1.5 0 100 Sh=70,Sd=30
25 APSH 1.6 0.75 1.5 0 100 Sh=80,
Sd=10,
Lm=10
24 AKCH 1.8 0.75 1.5 0 100 Sh=90,Sd=10
23 MPL5 2.5 0.76 1.5 0 100 Sh-90,Sd-10
22 MPL4 2.7 0.76 0.3 0 100 Sh=10,Sd=90
21 MPL3 3.6 0.77 1.5 0 100 Sh-90,Sd-10
20 MPL2 3.9 0.78 0.3 0 100 Sh=10, Sd=90
19 MPL1 5.0 0.79 1.5 0 100 Sh-90,Sd-10
18 PONT 6.2 0.82 1.5 10 90 Sh=90,Sd=10
17 MI03 10 0.85 1.5 50 50 Sh-80,Sd-20
16 MI02 16 0.86 1.5 30 70 Sh=40,Sd=50
15 MI01 23 0.90 3.0 20 80 Sh-90,Sd-10
14 OUG 35 0.95 4.5 20 80 Sh=100
13 EOC3 39 0.98 0.7 0 100 Sh-70,
Sd=20,
Lm=10
12 EOC2 50 1.03 0.7 0 100 Sh=80,
Sd=10,
Lm=10
11 EOC1 57 1.05 0.7 0 100 Sh-50,
Sd=20,
Lm=20,Do=10
10 PLC2 61 1.07 0.7 0 100 Sh-50,
Sd=20,
Lm=20,Do=10
9 PLC1 65 1.10 0.3 0 100 Sh-40,
Sd=20,
Lm=20,Do=20
8 CR2C 74 1.15 0.7 0 100 Sh=60,
Sd=10,
Lm=20,Do=10
7 CR2T 97 1.20 0.7 0 100 Sh-60,
Sd=10,
Lm=20,Do=10
6 CR1V 124 1.30 0.7 0 100 Sh-60,
Sd=10,
Lm=20,Do=10
5 CR1L 145 1.40 0.7 0 100 Sh-60,
Sd=10,
Lm=20,Do=10
4 MALM 157 1.50 0.7 0 100 Sh-60,
Sd=10,
Lm=20,Do=10
3 DOGR 173 1.60 3.0 0 100 Sh-60,
Sd=10,
Lm=20,Do=10
2 VLUP 178 1.63 0.0 0 100 Cr-100
1 BSMT 180 1.65 0.0 0 100 Cr=100
Sh = shale, Sd = sand, Lm = limestone, Do - dolomite, Cr - crystalline rock.
MUD DIAPIRS AND THERMAL ANOMALIES 213

of shaly sediments, at the diapir location of the section, and so to have a dynamically active
mud diapir during basinal evolution.
Measured heat fluxes for wells in the offshore South Caspian Basin (except for the
Apsheron-Balkhan zone - the eastward submarine extension of the Great Caucasus Mountains)
range between 20-42 mW/m 2 (0.5-1 HFU) with an average of about 28 mW/m2 (0.65 HFU)
for all 25 pseudowells along the section. For paleoheat flux variations, two factors were used
to assess a higher heat flux differing linearly with time to 70 mW1m2 in the Jurassic, the start
of the modeling calculations. First, regional tectonic evolution of the South Caspian Basin
(Nadirov and others, 1996) plus the presence of early volcanic uplifts (the Lesser Caucasus,
and the Godin Uplift towards the West Turkmenistan side of the Basin) suggest a higher heat
flux in early times; second, the massive recent sedimentation (about 10 km in the last 5-10
million years) suggests a "thermal blanketing" effect making for a higher paleoheat flux
(Tagiyev and others, 1996). The amount and types of kerogen for each depositional layer,
together with inferred paleoheat flux at deposition also are recorded in Table 1, with the
kerogen information coming from outcrops and well data (Tagiyev and others, 1996). Water
bottom temperature is held at the present-day value of 9 0 C throughout the calculations.
Because the GEOPETll code traces the evolution of fluid flow and overpressure with
time and space, it also is necessary to allow for fracture formation if the fluid overpressure
rises too high. Conventionally, this is allowed for by setting a value for the fracture coefficient
(ratio oftotal fluid pressure to lithologic load at a given depth) of 0.83 (Lerche, 1990) and is
so treated here.
For the generation of hydrocarbons from kerogen the simple kinetic model of Tissot
(as detailed in Tissot and Welte, 1978) was used, which uses six channels for kerogen
degradation to oil and for oil converting to gas. Other kinetic models that also have been used
provide similar results to those presented here.
The base of the Abikh mud diapir was taken to be rooted in Cretaceous sediments and
the timing to diapiric growth occurs as for the simple geometric model considered previously
(Fig. 4). The starting point of diapiric development was in the Late Miocene or Early
Pleistocene and by the beginning of Middle Pliocene (Productive Series time), a diapir of some
8-9 km height developed (Fig. 7A) with the crest of the diapir at a submudline depth of2-3
km. The additional growth of the Abikh diapir was accompanied by increased rates of
sedimentation; by the end of Late Pliocene the crest of the diapir remained at a submudline
depth of 2-3 km despite a high rate of diapiric rise (to the beginning of the Quaternary, the
diapiric column was 16-18 km tall). At the present, the crest of the Abikh diapir is almost at
the sediment surface (sea bottom). By the beginning of Middle Pliocene, the diapir had
penetrated all Cretaceous, Paleogene, and Lower Miocene (Maikopian) sediments; by the end
of Pliocene time, the diapiric crest was in the upper part of the thick Productive Series and,
today, penetrates all the sedimentary cover from Cretaceous to Quaternary. The sequence of
subfigures that constitute Figure 7 shows how the thermal conductivity changes in and around
the Abikh mud diapir resulting ·from both retention of porosity in the body of the diapir, and
from gas charging of the diapir as a consequence of hydrocarbon (gas) generation at depth.
The low thermal conductivity of the diapir relative to surrounding more compacted
sediments forces heat flux to avoid the diapiric structure. It should be remembered that the
GEOPETll code allows for both transient as well as steady-state transport of heat by
conduction and convection and, in addition, includes solubilities of gas in oil and water that
are temperature and pressure dependent, together with adiabatic expansion (and so cooling)
of gas as pressure and temperature decrease near the diapiric crest. Thus thermal cooling
effects not included in the prior simple geometric model now are considered and add to the
214 BAGIROV AND LERCHE

17 19 23

1l1000
,1000
31;000

GOOD )
5<000

6JOOO
21000
moo
,~ooo A
81000
11000 ~TlC1C. ~ llJtll.I:TtYITl (eallf.l~:sec)
9>X»
-5.00 HYBP
~ ~------------------------------------------------~

0.0 2.0 ~. o '.0 8.0 10.0 12.0 1., 0


....,,"' ......... ...,.... ..............
............................................ ~ ••••••• • 4

catlC/""u£

u Il 17 I'l 23

1l1000

27COO
31;000

.sooo I
S<OOO
6JOOO
moo
"000 B 81000
:2700'.) \£RJlr..ct.. TlDH't... CIlitU:11Vl1l (e.oIIUCRI.seel
9>X»
-LBO tI"I"W
~ ~------------------------------------------------~

14.0

Figure 7. Thermal conductivity variation in space and time around Abikh diapir across the 2-D section: A,
begiuning of Middle Pliocene; B, end of Middle Pliocene.

perception of a cool diapir relative to regional conditions. Figure 8A-8D shows the patterns
of temperature anomalies at different time-steps, corresponding to different stages of
development of both the basin and the diapir. The patterns of heat focusing and defocusing
that were seen in the simple geometric model (Fig. 5) are recorded here for the more
sophisticated model of evolution.
The simulations using GEOPETII take into account a variety of effects not included
in the simpler geometric model but reach the same basic conclusion: The body of the diapir
is cooler than surrounding formations with isotherms being 2-3 km deeper than in sediments
far from the diapir, and heat flux is focused towards the diapiric crest.
Of course, the pseudowell spacing selected here (about 3.3 km) indicates that fine-scale
details of diapiric shape evolution are not occurring as portrayed as with the geometric model.
Such a shortcoming can be remedied by including more pseudowells, at the expense of
considerably more computer time - that has not been considered worthwhile in the evaluation
of heat flux focusing and defocusing investigated here, but which is needed when hydrocarbon
migration and accumulation are the primary aims (Bagirov and Lerche, 1997b).
MUD DIAPIRS AND THERMAL ANOMALIES 215

:. .r . .
u 13 11 'IS
km

9000
19>00

27000
3SOOO

' S/X)O !
S<OOO
6J(I)O

21000 -lII1ttt!ltilllWll. nooo


-11111• •
2. 000
c 81000
~TJOl. ~ aJiIlCTJ~JTY (Ct!llltloV'~) 'lOOOO
27000
~ ~-I~
.~~~
m~~ __________________________________________ ~

14. 0

2S
km

9000

I 19000
2/000
3SOOO

' 5000 I
S<OOO
6J(I)O

nooo
81000

90000

Figure 7. Thermal conductivity variation in space and time around Abikh diapir across the 2-D section: C,
beginning of Quaternary; D, present-day.

DISCUSSION AND CONCLUSION

The problem of mud diapiric activity and associated thermal anomalies in the South
Caspian Basin has long been of interest. The temperature of water and "frothy" mud, coming
from relatively deep zones (as inferred from mineral content in waters), through gryphons and
salses associated with mud volcanoes and diapirs, is always cool to ice-cold despite
expectations that extruded deep material should be warm. At the same time the heat fluxes
inferred from measurements on the bodies of mud diapirs and mud volcanoes are typically an
order of magnitude or more higher than regional values, even though diapiric temperatures are
low. The results of all the numerical experiments reported here indicate that these observations
are accounted for readily by a low thermal conductivity in the diapir compared to surrounding
sediments causing focusing and defocusing of heat. This effect alone causes focusing of heat
towards the diapiric crest and produces isotherms 2-3 km deeper near the crest than in the
regional domain. In addition, adiabatic expansion of gas near the crest of a gas-charged diapir
adds to the cooling effect as well as making mud diapir thermal conductivities even lower than
without the expansion, exacerbating the already significant cooling effect. Near the base of
216 BAGIROV AND LERCHE

-
19 ?3 2S

"?7'"
''''
!f«<I

.-~
,,'"
6_
nooo
8\000
?'"
A
TOftRR1'lP.E CC)
"""
~.~
~~.,,~~~----------------------------------------~
?SO.' 300.0 3'50.'
,,'t===\O·i:~~~~I~"ffil·'mmm'IDl\Ot·'===2OO·0
~~~
ce)

-
11 19 23 2S

"''''
2_
!f«<I

'-1
,,'"
"'"
r_
.,'"
"""
2OO .0 ?SO.o JOO.' !SO.,
' .•r-__---=\O~.~;~==~=tmTTT1mTTT1mTTT1l\Orrt·"":'"O:-:c-:::-:::
ce)
Figure 8. Temperature across 2-D section: A, at 5.0 MYBP; B, at 1.8 MYBP.

the diapir, the opposite effect occurs: heat is defocused away from the diapiric base thereby
increasing the basal temperature relative to regional. This effect provides for an increased
generation of hydrocarbons from kerogen relative to regional, and an increased conversion to
gas. Consequently the supply of gas to the diapir increases, causing both an increased
buoyancy of the diapir and an additional lowering of the diapiric thermal conductivity. The
consequences of such thermal anomalies for hydrocarbon (oil and gas) generation, transport
and accumulation are profound, and are discussed in detail in Bagirov and Lerche (1997b).

ACKNOWLEDGMENTS

The work reported here was supported by the Industrial Associates of the Basin
Modeling Group at the University of South Carolina.
MUD DIAPIRS AND THERMAL ANOMALIES 217

Zl 25
km

-
18000
vooo

'5000 j
~

&3000

moo
111000
27000 9OOClO

~ ~------------------------------------------------~
0, 0 SO,O 100, 0 1SO,O 2OJ.O 250.0 3/».0 l'O.O

1C===:Jv~v~~~~~~llllllllllIlIImllillmlillmlll~~::=::=::=
::::: (Cl

u 13 1] 1.9 23 25

,-
km

2/000

l6000

<SOO> j
~

&3000

moo
111000

9OOClO

0•• 50.0 100, 0 150.0 200.0 250 •• 3/».0 l'O.O

Figure 8. Temperature across 2-D section: C, at 1.3 MYBP; D, at present-day.

REFERENCES

Bagirov, E., and Lerche, I., 1996, Hydrate hazards in the South Caspian Basin: Izvestiya Akad. Nauk
Azerbaijani, Nos. 1-6, 1993/1994, p. 116-124.
Bagirov, E., and Lerche, I., 1997a, Evolution of the Abikh Diapir, in Ali-zadeh, A., Guliev, I.S., Bagirov E., and
Lerche, I., eds., South Caspian Basin Risks and Hazards: Geology, Geochemistry and Diapirs, Nafta
Press, Azerbaijan Acad. Sciences, Baku, p. 304-358.
Bagirov, E., and Lerche, I., 1997b, Gas-charging of mud diapirs and fault leakage, in Ali-zadeh, A., Guliev, I.S.,
Bagirov, E., and Lerche, I., eds., South Caspian Basin Risks and Hazards: Geology, Geochemistry and
Diapirs, Nafta Press, Azerbaijan Acad. Sciences, Baku, p. 359-418.
Bredehoeft, J.D., Djevanshir, R.D., and Belitz, K.R., 1988, Lateral fluid flow in a compacting sand-shale
sequence: South Caspian Basin. Am. Assoc. Petroleum Geologists Bull., v. 72, no. 1, p. 416-424.
Brown, J.M., 1990, The nature and hydrogeologic significance of mud diapirs and diatremes for accretionary
systems: Jour. Geophys. Res., v. 95, no. B6, p. 8969-8982.
Brown, J.M., and Westbrook, G.K., 1988, Mud diapirism and subcretion in the Barbados Ridge accretionary
complex: the role of fluids in accretionary processes: Tectonics, v. 7, no. 3, p. 613-640.
Gambarov, J.G, Jafarov, Z.F., Mamedov, P.Z., and Shykhaliev, J.A., 1993, The seismostratigraphic and
structural-formational analysis of the regional proflles (cross-sections) data of the South Caspian Mega-
Trough, in Structural Formational and Seismostratigraphical Investigations of the South Caspian Mega-
Trough Sedimentary cover: AZNIIGEOPHYSIKA, Baku, p. 94-116.
218 BAGIROV AND LERCHE

Guliev, I.S., 1996, Model of mud volcanism, in Guliev, I.S., Bagirov, E., Lerche,l., Nadirov, R., and Tagiyev,
M., eds., The South Caspian Basin: Geologic Risks and Probable Hazards: Nafta Press, Azerbaijan
Acad. of Sciences, Baku, p. 506-530.
Ivanov, V.V., and Guliev, I.S., 1986, Attempt at an explanation of physico-chemical modeling of mud
volcanism: Bull. Moskovskoye Obshestvo Ispytateley Prirody, Geology Branch, v. 61, no. I, p. 72-79.
Lerche, I., 1990, Basin analysis: quantitative methods, v. I, Academic Press, San Diego, 657 p.
Lerche, I., Bagirov, E., Nadirov, R., Tagiyev, M., and Guliev,I.S., eds., 1996, Evolution of the South Caspian
Basin: geologic risks and probable hazards: Nafta Press, Azerbaijan Acad. Sciences, Baku, 625 p.
Lerche, I., and Petersen, K., 1995, Salt and sediment dynamics: CRC Press, Boca Raton, 322 p.
Nadirov, R.S., Bagirov, E., Tagiyev, M.F., and Lerche, I., 1996, Flexural plate subsidence, sedimentation rates,
and structural development of the super-deep South Caspian Basin: Marine Petrol. Geology, v. 14, no.
4, p. 383-400.
Tagiyev, M.F., Nadirov, R.S. Bagirov, E.B., and Lerche, I., 1996, Geohistory, thermal history and hydrocarbon
generation history of the north-west South Caspian Basin: Marine Petrol. Geology, v. 14, no. 4, p. 363-
382.
Tissot, B., and Welte, D.H., 1978, Petroleum formation and occurrence: Springer Verlag, New York, 538 p.
EFFECT OF OIL AND GAS SATURATION ON SIMULATION
OF TEMPERATURE HISTORY AND MATURATION

H. S. Poelchau 1• C. Zwach3, Th. Hantschel 2, and D.H. Welte2,

lInstitute of Chemistry and Dynamics of the Geosphere


Research Center Jiilich, Germany
2Integrated Exploration Systems (IES), Jiilich, Germany
3Norsk Hydro AlS, Stabekk, Norway

ABSTRACT

Calibration ofthermal histories in basin modeling usually relies on matching temperature


and vitrinite reflectance distribution in wells. The four main variables that can be adjusted to
improve such a match are heat flow, surface temperature, maximum depth of burial (or eroded
overburden), and thermal conductivity of the individual layers of rocks. This last parameter
can be used to match the fine structure of the vertical temperature distribution, thermal gradient
changes and heat anomalies.
Thermal conductivity values depend, among other variables, on the type of fluid filling
the pores of the sediment. Although water usually is assumed as the pore fluid, we have
experimented with using decreased conductivity resulting from high gas saturation in the pore
space of specific formations in our modeling. One such case study is the Alberta Deep Basin
in western Canada, where a large part of the lower Cretaceous section is thought to be gas
saturated, underneath a water-saturated seal. Another case study comes from northwestern
Siberia, where the largest gas accumulations on earth have been discovered.
Results show that the thermal effect of gas in pores, as opposed to water, is significant
and cannot be neglected in basin modeling. Gas saturation can explain frequently observed
sudden increases in vitrinite reflectance gradients or so-called "kinky" reflection profiles. The
gas effect also can be used to model heat anomalies in past geologic periods where
hypothetical increased heat-flow events cannot be justified.

INTRODUCTION

Basin modeling is concerned mainly with the reconstruction of the temperature history
of the sediments. Of primary interest is the evolution of petroleum source rock maturation

219
220 POELCHAU, ZWACH, HANTSCHEL, AND WELTE

which is related directly to heating variations with geologic time. Temperature history also
is important for estimating secondary effects, such as oil cracking to gas, as well as diagenetic
reactions changing the rock fabric, porosity, and permeability.
Temperature history modeling involves many interacting processes and depends mainly
on the knowledge (or estimation) offour sets of variables: depth and time of burial, heat-flow
history, surface temperature, and the distribution of time variant thermal properties of the
sediments involved. These include thermal conductivity and heat capacity, which also are
influenced by various interactions.
Using the appropriate values for thermal conductivity in a simulation may be as much an
art as a science. It involves knowing the lithology of each of the formations as well as its
variations, mixtures, and interbedding. Usually, in most simulation programs, thermal
conductivity is adjusted automatically for the changing in-situ parameters, such as temperature.
However, the type of fluid occupying the pore space of the rock also is a major contributing
factor which can change thermal conductivity. This problem is addressed in this paper by
looking at possible influences from a theoretical standpoint and by testing these considerations
with a simulation program, in our situation PetroMod(tm) produced by IES in Jiilich.
Basin modeling is used here as a tool for experimenting with geologic processes in
geologic settings approximated from reality and integrated from a large variety of data sources
(Poe1chau and others, 1997). Most variables involved in the modeling process have varying
degrees of uncertainty. This can be used to adjust and calibrate the model. For the thermal
history, all four variables mentioned can be varied within geologically reasonable limits. The
experiments with these variations are called sensitivity analyses. For our case histories we
have calibrated the models to match measured temperatures and vitrinite reflectances by
adjusting the available variables for the situation of water-filled pores. Then, in order show
the sensitivity to changes in pore fluids we replaced water with gas and ran the simulation
again. These comparisons are shown in the case histories.
Theoretical considerations: parameters affecting bulk thermal conductivity
Thermal conductivity is defined as the ability of a rock or substance to conduct heat. It occurs
as the coefficient? in the heat-flux equation (Fourier's law):

q= -A. grad T
Thermal conductivity usually is reported in units ofW m-1K 1 or as mcal S-I cm- I KI (=0.4184
Wm-1K 1).
The parameters influencing the effective or bulk thermal conductivity of rocks in a
sedimentary basin are rock composition on one hand and in-situ conditions on the other.
Composition includes the types of mineral grains and cements, the amount of porosity, and the
pore filling fluids. In-situ conditions are mainly temperature and, to a lesser degree, pressure.

Mixing Formulae

The literature is full of various approaches to estimate the effective or bulk thermal
conductivity of rocks or mineral mixtures based on the thermal conductivity of individual
components (Brigaud, Chapman, and Le Douaran, 1990; Krupiczka, 1967; Luo, Wood and
Cathles, 1994; Palciauskas, 1986; Robertson, 1979; Woodside and Messmer, 1961a;
Zimmerman, 1989). Each of the proposed methods works well for the particular set of data
the author is working with but less well for other sets. Therefore, we used the old mixing
formula for bulk thermal conductivity based on the weighted geometric mean:
EFFECT OF OIL AND GAS SATURATION SIMULATION 221

where fl ..3 .. are the fractions of the components oflitho10gy and pore fluids

and Al...3.their respective conductivities.

This is similar to the approach used by Brigaud, Chapman, and Le Douaran (1990) or Ungerer
and others (1990). It has the advantage of being simple, easy to visualize, and requires no
additional measurements such as grain size or grain shape that more specialized formulae use.
It also is the method applied for calculating effective thermal conductivity in the basin
modeling program (petroMod(tm» which was used for our case history studies.

Mineral Grain and Rock Matrix Conductivities

The conductivity of the rock matrix is determined by the mineral composition and the
conductivities of the respective components. This is one important starting point for the
mixing calculations. Figure 1 shows the broad range of variation for the most abundant matrix
constituents (rock types with 0% porosity) which are used in the modeling program.
Mineral thermal conductivity determined in the laboratory is not always the best value
to use for calculation of bulk rock thermal conductivity applied in basin modeling. Experience
during extensive use of the PDI and PetroMod(tm) modeling systems has shown that matrix
values for sandstone of approximately 3 W m-1K 1work better in basin modeling than the high
quartz values of7 or 8 W m-1K 1 listed in the literature (Brigaud, Chapman, and Le Douaran,
1990; Clauser and Huenges, 1995; Woodside and Messmer, 1961 b). The reasons are mu1tifo1d.
On one hand, the type of sandstone in most basins is not an orthoquartzite but may be
somewhat sha1y and contains feldspars. Both constituents have thermal conductivities lower
than mineral quartz. On the other hand, grain size seems to have a strong positive correlation
with thermal conductivity. Experimental data on artificial quartz sands (Midttomme and
Roaldset, 1998) show recalculated matrix values about 3 W m-1K 1 for samples in the sand size
range. For shales, Blackwell and Steele (Blackwell and Steele, 1989) have shown that in-situ
conductivities are lower by as much as 25% than those reported in the literature based on lab
measurements of clay minerals.

Thermal Conductivity of Fluids

Thermal conductivity of fluids is lower than that of most rock solids (Fig. 1) The
exception is coal which is slightly less conductive than water. Liquids have much higher
conductivities than gases, and oil tends to take intermediate values between gas and water. The
contrast between gas and water conductivities (0.03 and 0.6 W m-1K 1, respectively) is larger
than that between coal (0.5) and salt (5.7) which represent the least and most conducting
lithologies in sedimentary basin modeling. Increasing salinity of water tends to decrease
thermal conductivity [Landolt-Bomstein data (Siedler and Peters, 1986)].
222 POELCHAU, ZWACH, HANTSCHEL, AND WELTE

3.5

Figure 1. Thennal conductivities of abundant solids and fluids. Values for rocks are matrix conductivities for
0% porosity at 20°C. Data from IES (Wygrala, 1989), Blackwell and Steele (1989), Clark (1966), CRC
Handbook (Weast, 1974).

Porosity

For a rock with a given mineral composition, changes in porosity produce the greatest
change in thermal conductivity, more than the normal temperature variation. Porosity reduces
bulk thermal conductivity (Fig. 2) because the conductivity of the fluid filling the void space
is lower than that of most solids. The effect depends, of course, on the type of pore filling fluid
and on the saturation. In the modeling program, porosities are calculated according to
empirical depth and compaction relationships, optimized to give observed layer thicknesses
and checked against measured porosities.

Temperature

Conductivity of most solids generally decreases with rising temperature; only materials
with low thermal conductivity at ambient conditions (less than 1 W m-1K-l) tend to become
slightly more conductive (Somerton, 1992; Tikhomirov, 1968). Fluid thermal conductivity
increases with temperature. The effect is relatively greater for gases than for water. This
indicates that in a porous rock the increase of thermal conductivity of the fluids can, to some
degree, counteract the conductivity decrease with temperature of the minerals.
EFFECT OF OIL AND GAS SATURATION SIMULATION 223

3.5

3.0 \h.
.~
., ~ r--.... .....
~

~
-i:
., ~ ~ ............... ~
2.5

-
,~
~
"ti 2.0

---
1''',
c:
o . ~ ~<!. dSfn ... ~ ~ !!!.!.~ ater
' .... ~;,~- ~ I--!-~
~ 1.5
co Oil
...E ~ .. ..
/]e'-
~ 1.0 ·'-I!::Jlo -....'/0;
tone-;
~~
I-
-!-{}c t"- .. , .. r-- .. _
0.5

1&278.98
---
S
- ~s
"- .. -
0.0
o 0.1 0.2 0.3 0.4 0.5
Porosity
Figure 2. Bulk thennal conductivity of sandstone as function of porosity and pore fluid (at ambient T and p),
calculated with geometric mixing formula.

Pressure

Some experimental data (Kappelmeyer and Haenel, 1974) show that the pressure effect
is negligible after the first few kPa where bridging of microcracks or improvement of grain
contacts takes place. Therefore, lithostatic pressure effects, other than compaction and loss of
porosity, are ignored in this analysis. However, the effect of fluid pressure on the thermal
conductivity of gas can be considerable (data in Gallant (1968) and Landolt-Bomstein
(Todheide, Hensel, and Franck, 1968)). For the first 1000 m of burial the pressure effect
increases methane thermal conductivity by 25 to 30% whereas the temperature effect
contributes only about a 10% increase. Below 2000 m, at higher temperatures, the pressure
effect diminishes, and methane thermal conductivity increases more slowly.

Partial Water Saturation

For the simulations we generally have applied the geometric mixing formula. We also
have applied it to calculate the bulk thermal conductivity of partially saturated rocks, that is
where the pore water has been partly replaced by natural gas. This can be a problem, as
pointed out by Somerton (1992):
"Conductivity of the wetting phase will have the greatest effect on thermal conductivity
of a multi fluid-saturated rock. Since water is generally the wetting fluid in subsurface
reservoir rocks and furthermore, since water has the highest thermal conductivity of any fluid
which might occupy the pore spaces in reservoir rocks, water saturation will have a
disproportionately large effect on conductivity of the fluid-saturated rock. Thus, application
of mixing laws to obtain an effective composite value of fluid conductivity is probably not
valid." (Somerton, 1992).
Clauser and Huenges (1995) presented data from a thesis (Reibelt, 1991) showing a
strong deviation of measured thermal conductivity from the geometric mixing formula for
water saturation (Fig. 3). As they describe it, the curve reflects two independent processes:
first, the initial wetting of the pores at the grain contacts rapidly removes the dry rock
224 POELCHAU, ZWACH, HANTSCHEL, AND WELTE

2
a 0.2 0.4 0.6 0.8 1
(Air) Sw (Water)

Figure 3. Thermal conductivity vs. water saturation of sandstone with 18% porosity. Experimental data (squares
with shaded error envelop) from Reibelt (1991) in Clauser and Huenges (1995). For comparison curve derived
from geometric mixing formula.

resistance. This causes a steep increase of thennal conductivity between 0 and 10% water
saturation. Secondly, the gradual filling of the rest of the pore space (replacing air or gas with
water or oil) results in a much reduced gradient.
In his book, Somerton (1992) modifies the fonnula of Krupiczka (1967) to simulate the
effect of partial water saturation on thennal conductivity. The resulting curve does show the
same effect as Clauser and Huenges' (1995) data with a strong thennal conductivity increase
for low saturation values. However, the calculated values are too low to match the
experimental curves of Clauser and Huenges (1995). Of course, the equation can be modified
to fit this data set, but this would be just another example of a fonnula that fits only one data
set.
These data indicate that in natural systems with irreducible water saturation, the effect
of gas may not produce as large a contrast as assumed in the simulations using the simple
geometric mixing fonnula. However, where gas (or oil) is the wetting phase, the approach
used here should be valid.

Combined Effects of Subsidence

As sediments are buried in a basin they experience increasing temperature and pressure
as well as a progressive loss of pore space. These simultaneous processes affect sediment
physical parameters in various and differing ways. For bulk rock thennal conductivity one has
to consider and combine mainly three overlapping effects changing with increasing
temperature and depth of burial (Fig. 4): (1.) Mineral (rock matrix) conductivity decreases with
depth as a function of temperature; pressure has only minor influence within the nonnal
geological environment. (2.) Pore fluid conductivity increases with temperature and, in the
situation of gas, with pressure during shallow burial. (3.) The proportion of pore fluid relative
to the solid mineral matter decreases with compaction because of decreasing porosity.
The effective thennal conductivity of a fonnation therefore evolves in different ways
depending primarily on the parameters temperature, pressure, pore fluid composition, and
porosity (Fig. 4). The initial trend is towards increasing conductivity because of rapidly
EFFECT OF OIL AND GAS SATURATION SIMULATION 225

Thermal Conductivity [W rrr 1K·l) Porosity


0.00 1.00 2.00 3.00 0.0 0.2 0.4
0 15
<G.
I",,,, ~
I
I~
"\\
2000 Q)
c: I 'R\ 6S~
I.r:: '"
.£ I \
l!!
~

.. I l!!
Q)

ii. E
I \
..E
115 ~
C Qjl
I
4000 Oil
I
3:, I ~

j
I I 165
I I
I
6000
0.00 1.00 2.00 3.00 0.0 0.2 0.4

Figure 4. Thermal conductivity changes of sandstone with burial calculated for influence of porosity,
temperature, and pore filling fluids. Geothermal gradient is given as 30 KIkm and porosity change (right) by
exponential formula: 0 = 0.4 exp(-0.0002048 z). Note large difference between sandstone filled with methane
and sandstone filled with water. Note also that thermal conductivity of gas filled sandstones keeps increasing
to greater depth than that of water filled sandstones.

declining porosity and increasing proportion of the conductive mineral matrix. With
increasing burial, as porosity decrease slows down and temperature rises, the effect of
declining mineral conductivity takes over and the bulk conductivity shows a reversed trend
toward reduced values. However, if the pore fluid is gas instead of water, the conductivity
continues to increase and the reversal point of the curve can move to greater depth (or
temperature). This is because gas conductivity rises almost linearly with temperature whereas
pore water conductivity levels out. In a situation where the gas saturation increases slowly
with depth, the bulk conductivity curve will lie between the water and gas curves (Fig. 4) and
eventually approach the gas curve with an overall effect of decreasing conductivity with depth.
For oil saturation the effect would be less pronounced and the sandstone curve would lie closer
to the water-filled sandstone curve.

Test Model

The effect of gas in pore space on thermal conductivity and temperature history was
tested with a modified code of the program package PetroMod(tm) (IES 1993) taking into
account hydrocarbon saturation changing with geologic time and differing in space. A simple
structural trap model was constructed to simulate temperature history with and without gas fill.
The properties of the model are summarized in Table 1. The burial history of the model is
shown in Figure 5. Figure 6 illustrates the thermal insulation effect ofa "gas bubble" in the
structural high of the section, representing a spatially and temporally limited gas field. The
change in isotherms and maturation ofthe source rock is clearly visible. The gas field causes
the temperature to rise in the layers underneath, including the source rock. It also hinders heat
from moving above the reservoir and depresses the isotherms in the shallow layers. Porosity
distribution and thermal conductivities with depth at the location of the structure are shown
in Figure 7, including the resulting temperature and vitrinite reflectance differences for water
and gas filled pore space.
226 POELCHAU, ZWACH, HANTSCHEL, AND WELTE

Geologic Age (Ma)


100 75 50 25

' -_ _ _-1 1000

----::----1 Depth
(m)

2000

Figure 5. Burial history of test model. Sequence of sedimentation and subsidence for location at crest of structure
is shown in black lines. White lines with gray shaded bands are isotemperature lines for simulation with gas
accumulation.

Table 1: Description of the Test Model

20 Grid points with 1 km spacing


(section length =19 km)
10 events of 10 m.y. each

Start of sedimentation 110 Ma


Age of Structure 20 Ma

Heat flow: 103 m W m- 2

Gas saturation in reservoir: 100% at crest during last 20 m.y.

Lithology: Source rock (layer #1) - carbonaceous shale


Reservoir (layer #6) - sandstone
Cap Rock (layer #7) - shale
Others - sand and siltstone.
No lateral facies changes

Water depth - 10m throughout

Surface Temperature: 21 to 13 °C (oldest to youngest)


EFFECT OF OIL AND GAS SATURATION SIMULATION 227
Water Gas Accumulation
am
9 9

8 8

1000 +-----'-~

2000

3000
m
Depth ===:;:,.:.,;.,:.,:...:...:..,.:....:..,.:.:..:..:::.:.:..::;.,:.,:::..:..:.:..:;.,:.,:~===i:;:]
- l km
Figure 6. Comparison of two simulations of schematic hydrocarbon trap to show effects of water vs gas
saturation on temperature. Left: All sediments water filled. Right: Gas filling in structurally high position of
layer 6. Black lines denote layer boundaries in sand-shale sequence; dashed white lines are isotherms showing
simulated present day temperature field. Gas was generated in layer 1 and has migrated upward into sandstone
reservoir (layer 6) beneath shale cap rock (layer 7). Gas saturation in trap was assumed to be present for last 20
m.y.

Case Histories

Gas fields usually have a limited extent in a basin but can have a marked local effect on
the temperature history and distribution. This effect should not be ignored, especially when
concerned with large gas accumulations. One such case history is the megagiant Urengoy gas
field in NW Siberia. By contrast, some sedimentary basins, such as the Alberta Deep Basin
(Masters, 1979, 1984) or the San Juan Basin (Berry, 1959; Silver, 1968), have deeply buried
sections that are completely gas saturated beneath a water saturated zone. Clearly, gas
saturation must be considered for the calibration of the basin thermal history with the
consequence that estimates of input parameters such as heat flow would have to be lowered
in order to match measured vitrinite reflectance. Studies in the Alberta Deep Basin (Zwach
and others, 1994) have shown how different the reconstructed heat-flow history can be when
gas is (partially or completely) substituted for water as pore filling fluid. Conversely,
comparing water and gas filled models for the same heat flow history demonstrates greatly
increased vitrinite reflectance and much higher petroleum generation.

Urengoy Field, NW Siberia

The West Siberian Basin, located east of the Urals and west of the Siberian platform and
extending into the Arctic Ocean, contains mainly oil fields in its central part, whereas in the
northern part the world's largest gas accumulations have been located. Among these, the
228 POELCHAU, ZWACH, HANTSCHEL, AND WELTE

Urengoy gas field, west ofthe Pur River, is the largest known gas accumulation on Earth. The
entire West Siberian Basin contains about one-third of the conventional global gas reserves.
The major part of the gas is trapped at the top ofthe Pokur Formation (Aptian to Cenomanian)
in comparatively shallow depths between 800 to 1200 m. The gas consists of almost pure
methane. The source of the gas is controversial: it may be of bacteriogenic origin from the
Pokur Formation itself, or it has migrated up from more deeply buried source rocks below or
farther to the south (Littke and others, 1999).
The geology of the area is described in detail in Peterson and Clarke (1991) and
Rovenskaya and Nemchenko (1992). The post-Permian rift basin was filled with a thick
sequence of clastic sediments ranging from shales to shaly and silty sands. Source rocks are
mainly the Jurassic Bazhenov Shale and the Pokur Formation which contains coal seams and
dispersed coaly material. The highly porous and unconsolidated Pokur sands also serve as
reservoir for the large gas accumulations of the Urengoy and other gas fields. The main seal
for the Pokur field is the Turonian Kuznetsov Shale; however, there are many stacked pools
with individual thin shale seals.
In order to test the effect of the gas accumulation in basin modeling we have compared
two simulations for the same well (Urengoy 266) in the southern part ofUrengoy field, one
with water-saturated pore space, the other with gas saturation.
The assumed evolution of gas saturation in the Pokur reservoir is shown in the context
of the Tertiary burial history of well U266 (Fig. 8). As the layer properties in the simulation
model have to be averaged over considerable thicknesses we set the maximum saturation to
50%. This should be understood to indicate 50% of the sandstones being almost fully
saturated, rather than all of the sands having 50% gas saturation.

Porosity (%) Therm. Conduct. (W/m/K) Temperature (· C) Vitrinite Reflectance (%)


o 20 40 o 0.5 1.0 1.5 a 100 200 1.0 2 .0
0t=======~~~~~~~~~====~-=======~

1000

g
.c
ii
CI>
o
2000

3000 L--L______~~~________~_____

Figure 7. Comparison of some input and output parameters of simulation (at crest of structure). Porosity
influences thermal conductivity but is unchanged for gas case. Thermal conductivity is greatly reduced in layer
6 (just below lOOOm) because of gas filling. Temperature and vitrinite reflectance are both increased in layers
beneath gas reservoir, affecting maturity of source rock (at base of section).
EFFECT OF OIL AND GAS SATURATION SIMULATION 229

The two panels of Figure 9 show the difference in simulated temperature history between
water-filled and partially gas-filled Polrur Fonnation, overlaid on the Tertiary burial history
of well U266. The rise oftemperature for the gas situation in the layers below the Polrur is
clearly visible whereas the layers above it have the same temperature structure in both
situations.
The depth plot (present time) of the input parameters porosity and thermal conductivity
and the output parameters temperature and vitrinite reflectance (Fig. 10) gives an additional
view of the changes because of gas saturation. Note the reduction of thennal conductivity
resulting from partial gas saturation. The lowered conductivity acts similar to a thennal
blanket. Correspondingly, below the gas the temperature increases on the order of7 to 8°C and
vitrinite reflectance by as much as 0.08% Ro in the fonnations below the Cenomanian. The
effect would be greater for higher gas saturations.

Elmworth Field, Alberta Deep Basin

Another good test for the effect of gas can be made in the Elmworth field of the Alberta
Deep Basin. As a foredeep basin located just east of the Canadian Rockies (Fig. 11) and
consisting of a thick Mesozoic clastic section overlying Paleozoic carbonates and shales, the
Alberta Deep Basin contains gas in a rather unusual trapping mechanism (Masters, 1979).
Much of the Mesozoic section is gas saturated in the deep part of the basin and sealed by
water-bearing rocks above and updip.
Sedimentation and subsidence in the basin started to accelerate in Late Cretaceous and
became highest in the Early Tertiary just before the Laramide Orogeny started major uplift and
erosion of this area. Gas generation began during the last part of subsidence (ca. 60 Ma) and
continued during the time of deepest burial. Figure 12 shows the burial history and the
associated temperature evolution of well 10-35-71-13W6 near the Alberta-British Columbia
border.
Having done detailed basin analysis and modeling on this area (poelchau and Zwach,
1994; Zwach, 1995) we can assign gas saturation to each fonnation for each geological interval
(Fig. 13A) in order to calculate thermal conductivities adjusted to the gas content of the rocks
changing with time. A new simulation with this adjusted set of thermal conductivities
produces a strong thermal insulation effect, raising temperature in and below the gas saturated
sediments. Figure 13B shows the strongly increased maturation level relative to the base case
modeled with water-filled rocks.
The vitrinite reflectance profile in Figure 13B shows a fonn that occurs in sedimentary
basins which has been described by Law, Nuccio, and Barker (1989). The shape of the curve
is nonlinear with anomalous kinks or bends. The kinked maturity profiles may be the result of
unconformities, faults, recycled organic matter, and thermal conductivity contrasts associated
with lithological boundaries (Law, Nuccio, and Barker, 1989). However, the presence of gas
in the pore space beneath a caprock adds another possible explanation for the fonnation of
such maturation kinks. Given the right geological setting, the shape of the vitrinite reflectance
curve could be used as an indicator of fonner gas saturation even in sediments which are water
filled today or have lost their high porosity because of diagenesis.
Another typical problem occurs when estimating volumes of coal in fonnations such as
the Falher Member in the Alberta Deep Basin. Unrealistic high coal percentages, compared to
what well logs would suggest, may be required to calibrate the model. However, using gas
saturation instead of raising the coal content may achieve a good calibration.
230 POELCHAU, ZWACH, HANTSCHEL, AND WELTE

Geologic Time (Ma)


60 50 40 30 20 10 a
a a

1000

_ 2000
§.
J:
a.
Q)
o 3000

0.5
4000
0.4

0 0.3
5000 0 0.2

D 0.1

SOOO
Figure 8. Burial history diagram for Tertiary history of well U266 showing assumed evolution of gas saturation
in Pokur reservoir used in simulation.

CONCLUSIONS

Case histories and test models show that it is important in reconstructions of the thermal
history of a sedimentary basin to consider the effect of gas as a pore-filling phase where there
is good reservoir porosity. In general, models including gas saturation require a history with
lower heat-flow values than for water-filled models. There also is a possible feedback effect

1000

2000

~~~=h~~~ 1m)
4000

5000

~ ________________ ~~ __________ ~ 0000

Water filled pore space Partial gas saturation In Pokur Fm.

Figure 9. Comparison of simulated temperature history of well U266 (in the framework of Tertiary burial history
diagram) between water-filled and partially gas-filled Pokur Formation. Note relative rise of temperature in
layers below Pokur in right pane\.
EFFECT OF OIL AND GAS SATURATION SIMULATION 231
Porosity (%) Thermal Conductivity Temperature (0C) Vitrinite Reflectance
(Wlm/K) (%)
a 10 a 0.5 1.0 1.5 50 100 150 1.0 1.5

1000 t===:::::Jc:;:====::j:::=======t~=j::;~\==============

2000 I - -......L------+- - - -
I..c:
.,
0..
c
3000t---~--------~----------4_-r~~~~~-------~

4000

5000~~----------4_----------Lr~~~~~~~--~~--

Figure 10. Depth plot (present time) of input parameters porosity and thermal conductivity and output parameters
temperature and vitrinite reflectance for well U266. Note reduction of thermal conductivity resulting from partial
gas saturation and corresponding increase of temperature and vitrinite reflectance in formations below
Cenomanian. Effect would be greater for higher gas saturations.

Study Area

a
-....200km
Figure 11. Geological index map of western Canada showing location of Alberta Deep Basin and case study area
just south of Peace River Arch (PRA). Elmworth field lies approximately in middle of study area. After Osadetz
(1989). Contours indicate thickness of sediment fill (in km) in Western Canada Sedimentary Basin.
232 POELCHAU, ZWACH, HANTSCHEL, AND WELTE

Geo!. Time (Mal


250 200 1 50
1000

1000
Depth
(m)
2000

3000

4000

H!iP 14JUl8
5000
Figure 12. Burial history diagram of well 10-35-71-13W6 with temperature evolution shown as gray shaded
bands and white dashed lines. Sedimentation and subsidence was fastest during Late Cretaceous and Early
Tertiary when gas generation began. Uplift during Laramide orogeny removed more than 1500 m of overburden.

gas satura1ton Simulated Vitrinite Rellectance


o 1 0 1.0 2.0 3.0

-soo

o
-500

500
o depth
(ms.s.)
present 23
depth
(ms.s.) 1000
500

1500
1000

2000
1500

2000 40
1 non
~near
60 (>.gelMa -
GaOl. 0 reservoir rocks
ogas .1
saturation • source rocks

Figure 13. Well IO-35-71-13W6, Elmworth Field, Alberta Deep Basin. Left: Gas saturation (x-axis) in pore
space, derived from simulation, for various stratigraphic levels (z-axis) changing with geologic time (y-axis).
This is simulation input used to compare against water-filled rocks. Source rocks are shown in dark gray,
reservoirs in light gray. Right: Comparison of maturation trends for rocks filled with gas as shown on left panel
(thick dashed line) and base case with water. Note "kink" produced in curve for gas. After Zwach (1995).
EFFECT OF OIL AND GAS SATURATION SIMULATION 233

of gas emplacement on gas saturation: higher amounts of gas in the pore space in effect trap
heat flow from the base because of increased insulation. This leads to higher temperatures
which in turn would result in an enhanced rate of gas generation.
The gas effect on lowering thermal conductivity is significant only for rocks with good
porosity (about 20% or more). However, even where present porosity is low, we may be able
to see a fossil gas accumulation as an expression of anomalously high or kinky vitrinite
reflectance distribution.
To consider gas in modeling, several additional parameters must be available for
calibration of the thermal history model: the emplacement of gas in the pore space obviously
depends on the generation and migration of gas in the sediments. Inadequate data on the actual
saturation and the spatial gas distribution, timing, migration pathways, caprock leakage,
permeabilities and inhomogeneities therefore can cast a shadow of uncertainty on the
calibration of the models. It is obvious from these findings that a fitting of the calculated to the
measured vitrinite reflection data becomes rather complicated when considering the gas
emplacement and its influence on heat conductivity.

ACKNOWLEDGMENTS

Ernst Huenges pointed out the problem of water saturation in the mixing formula and
supplied supporting data. Dan Merriam and Andrea Forster supported and encouraged the
writing of the manuscript. The paper was improved by reviews of Dan Merriam and two
anonymous reviewers.
The Urengoy project benefitted considerably from the financial support and scientific
interest by Ruhrgas AG, Essen, Veba Oel AG, Gelsenkirchen, and Wintershall AG, Kassel.
Financial support from the Bundesministerium fUr Bildung, Wissenschaft, Forschung und
Technologie (BMBF), Kennziffer GAS 1000, is gratefully acknowledged.
The Alberta Deep Basin study was supported by Canadian Hunter and the
German-Norwegian Cooperation Program.

REFERENCES

Berry, F. A. F., 1959, Hydrodynamics and geochemistry of the Jurassic and Cretaceous systems in the San Juan
basin, northwestern New Mexico and southwestern Colorado: unpubl. doctoral dissertation, Stanford
Univ., 466 p.
Blackwell, D. D., and Steele, J. L., 1989, Thermal conductivity of sedimentary rocks: measurement and
significance, in Naeser, N. D., and McCulloh, T. H., eds., Thermal History of Sedimentary Basins -
Methods and Case Histories: Springer-Verlag, p. 14-36.
Brigaud, F., Chapman, D. S., and Le Douaran, S., 1990, Estimating thermal conductivity in sedimentary basins
using lithologic data and geophysical well logs: Am. Assoc. Petroleum Geologists Bull., v. 74, no. 9, p.
1459-1477.
Clark, S. P., Jr., ed., 1966, Handbook of physical constants: Geol. Soc. America Mem. 97, 587 p.
Clauser, C., and Huenges, E., 1995, Thermal conductivity of rocks and minerals; in Ahrens, T. J., ed., Rock
Physics and Phase Relations - A HandbookofPhysical Constants: Reference Shelf, v. 3: Am. Geophys.
Union, Washington, D.C., p. 105-126.
Gallant, R. W., 1968, Physical properties of hydrocarbons, v. 1: GulfPubl. Co., Houston, TX, 223 p.
Kappelmeyer, 0., and Haenel, R., 1974, Geothermics with special reference to application: Gebriider Borntrager,
Berlin, 240 p.
Krupiczka, R., 1967, Analysis of thermal conductivity in granular materials: Intern. Chem. Engineering, v. 7,
no. 1, p. 122-144.
234 POELCHAU, ZWACH, HANTSCHEL, AND WELTE

Law, B. E., Nuccio, V. F., and Barker, C. E., 1989, Kinky vitrinite reflectance well profiles: evidence of
paleopore pressure in low-permeability, gas-bearing sequences in Rocky Mountain Foreland Basins: Am.
Assoc. Petroleum Geologists Bull., v. 73, no. 8, p. 999-1010.
Littke, R., Cramer, B., Gerling, P., Lopatin, N. V., Poelchau, H. S., Schaefer, R. G., and Welte, D. H., 1999, Gas
generation and accumulation in the west Siberian Basin and the atmospheric methane balance: Am. Assoc.
Petroleum Geologists Bull., submitted, .
Luo, M., Wood, J. R., and Cathles, L. M., 1994, Prediction of thenna1 conductivity in reservoir rocks using fabric
theory: Jour. Appl. Geophysics, v. 32, no. 4, p. 321-3~4.
Masters, 1. A., 1979, Deep basin gas trap, western Canada: Am. Assoc. Petroleum Geologists Bull., v. 63, no.
2, p. 152-181.
Masters, J. A., 1984, Lower Cretaceous oil and gas in Western Canada, in Masters, J. A., ed., Elmworth - case
study of a Deep Basin gas field: Am. Assoc. Petroleum Geologists Mem. 38, p. 1-33.
Midttomme, K., and Roaldset, E., 1998, The effect of grain size on thermal conductivity of quartz and silts:
Petroleum Geoscience, v. 4, no. 2, p. 165-172.
Osadetz, K. G., 1989, Basin analysis applied to petroleum geology in Western Canada, in Ricketts, B. D., ed.,
Western Canada Sedimentary Basin. A Case History: Can. Soc. Petroleum Geologists, Calgary, p.
287-306.
Palciauskas, V. V., 1986, Models for thermal conductivity and permeability in normally compacting basins, in
Burrus, 1., ed., Thermal Modeling in Sedimentary Basins: Ed. Technip, Paris, p. 323-336.
Peterson, 1. A., and Clarke, J. W., 1991, Geology and hydrocarbon habitat of the West Siberian Basin: Am.
Assoc. Petroleum Geologists, Studies in Geology, v. 32, 96 p.
Poelchau, H. S., Baker, D. R., Hantschel, T., Horsfield, B., and Wygrala, B., 1997, Basin simulation and the
design of the conceptual basin model, in Welte, D. H., Horsfield, B., and Baker, D. R., eds., Petroleum and
Basin Evolution: Springer-Verlag, Heidelberg, p. 5-70.
Poelchau, H. S., and Zwach, C., 1994, Basin simulation and diagenetic models: Albian Cadotte Sandstone,
Alberta Deep Basin, Canada: Berichte des Forschungszentrums Jiilich, v. JiiI-2882, p. B1.1-142.
Reibelt, M., 1991, Untersuchung des Einflusses der Oberflachenbeschaffenheit und der Fluidsattigung von
Gesteinen auf die Messung der Wiirmeleitfahigkeit mit einer Halbraumlinienquelle: unpubl. Diplomarbeit,
Techn. Univ. Berlin, Berlin, 111 p.
Robertson, E. C., 1979, Thermal conductivity of rocks: u.S. Geol. Survey Open-File Rept. 79-356, 31 p.
Rovenskaya, A. S., and Nernchenko, N. N., 1992, Prediction of hydrocarbons in the West Siberian Basin: Bull.
Cent. Rech. Explor. Prod. Elf Aquitaine, v. 16, p. 285-318.
Siedler, G., and Peters, H., 1986, Properties of sea water, in Siindermann, J., ed., Oceanography:
Landolt-Bornstein Numerical Data and Functional Relationships in Science and Technology. New Series.
Group 5: Geophysics and space research, v. 3: Springer, Berlin, p. 233-258.
Silver, C., 1968, Principles of gas occurrence, San Juan basin, Natural Gases in North America. Pt. 2 Natural
gases in rocks of Mesozoic age: Am. Assoc. Petroleum Geologists Mem. 9, p. 946-960.
Somerton, W. H., 1992, Thermal properties and temperature-related behavior of rock/fluid systerus:
Developments in Petroleum Science, v. 37: Elsevier, Amsterdam, 260 p.
Tikhomirov, V. M., 1968, Heat conductivity of rocks and their relationship with density, saturation and
temperature: Neftyanoe Khozyaystvo, v. 46, no. 4, p. 36-40.
Todheide, K., Hensel, F., and Franck, E. U., 1968, Wiirmeleitfahigkeit von Gasen, in Schafer, K., ed.,
Transportphiinomene II - Kinetik - Homogene Gasgleichgewichte: Landolt-Bornstein Zahlenwerte und
Funktionen. II. Band: Eigenschaften der Materie in ihren Aggregatzustanden, v. 5b: Springer, Berlin, p.
39-71.
Ungerer, P., Burrus, J., Doligez, B., Chenet, P. Y., and Bessis, F., 1990, Basin evaluation by integrated
two-dimensional modeling of heat transfer, fluid flow, hydrocarbon generation, and migration: Am. Assoc.
Petroleum Geologists Bull., v. 74, no. 3, p. 309-335.
Weast, R. C., ed., 1974, CRC Handbook of chemistry and physics: CRC Press, Cleveland, Oh, variously
paginated.
Woodside, W., and Messmer, J. H., 1961a, Thermal conductivity of porous media. I. Unconsolidated sands:
Jour. Applied Physics, v. 32, no. 9, p. 1688-1699.
Woodside, W., and Messmer, J. H., 1961b, Thermal conductivity of porous media. II. Consolidated rocks: Jour.
Applied Physics, v. 32, no. 9, p. 1699-1706.
Wygrala, B. P., 1989, Integrated study of an oil field in the southern Po basin, northern Italy: unpubl. doctoral
dissertation, Univ. Koln, Berichte Kernforschungsanlage Jiilich, no. 2313, 217 p.
EFFECT OF OIL AND GAS SATURATION SIMULATION 235

Zimmennan, R. W., 1989, Thenna1 conductivity of fluid-saturated rocks: Jour. Petroleum Sci. & Eng., v. 3, no.
3, p. 219-227.
Zwach, C., 1995, Diagenesis and temperature history ofthe Cadotte Sandstone, Alberta Deep Basin, Canada:
Integration of reservoir quality analysis and basin modeling: Berichte des Forschungszentrums Jiilich
Jiil-3082, 173 p.
Zwach, c., Poelchau, H. S., Hantschel, T., and Welte, D. H., 1994, Simulation with contrasting pore fluids: Can
we afford to neglect hydrocarbon saturation in basin modeling?: Conf. on Basin Modelling, London,
Geological Soc., Petroleum Group, p. 81-84.
CONTRIBUTORS

Phillip A. Armstrong, Department of Geology and Geophysics, University of Utah, Salt Lake
City, UT 84112, USA

Stefan Bachu, Alberta Geological Survey, Edmonton, Alberta T5K 2G6, Canada

E. Bagirov, Geological Institute, Azerbaijan Academy of Science, Baku, Azerbaijan

Graeme R. Beardmore, Department of Geological Sciences, Southern Methodist University,


Dallas, TX 75275, USA

David D. Blackwell, Department of Geological Sciences, Southern Methodist University,


Dallas, TX 75275, USA

David S. Chapman, Department of Geology and Geophysics, University of Utah, Salt Lake
City, UT 84112, USA

T.A. Dodd, BPNorge UA, PO Box 197,4033 Forus, Norway; present address: BPExploration
pIc, Houston, TX USA

William D. Everham, Department of Geological Engineering and Science, Michigan


Technological University, Houghton, MI 49931, USA

W. Fjeldskaar, RF-Rogaland Research, PO Box 2503, 4004 Stavanger, Norway

Andrea Forster, GeoForschungsZentrum Potsdam, Telegrafeilberg, D-14473 Potsdam,


Germany

Grant Garven, Department of Earth and Planetary Sciences, Johns Hopkins University,
Baltimore, MD 21218, USA

William D. Gosnold, Jr., Department of Geology and Geological Engineering, University of


North Dakota, Grand Forks, ND 58202, USA

Th. Hantschel, Integrated Exploration Systems GmbH (IES), Bastionstr. 11-19, D-52428
Jiilich, Germany

237
238 CONTRIBUTORS

Jacqueline E. Huntoon, Department of Geological Engineering and Science, Michigan


Technological University, Houghton, MI 49931, USA

Alan Jessop, 333 Sliver Ridge Crescent NW, Calgary Alberta T3B 3T6, Canada

Christopher Jessop, Jessop Scientific Software, 333 Silver Ridge Crescent NW, Calgary,
Alberta T3B 3T6, Canada

H. Johansen, Geologica a.s., PO Box 8034, 4003 Stavanger, Norway

Ian Lerche, Department of Geological Sciences, University of South Carolina, Columbia, SC


29208, USA

Guichang Lin, Department of Geology and Geophysics, Louisianna State University, Baton
Rouge, LA 70803, USA

Jacek A. Majorowicz, Northern Geothermal Consultants, 105 Carlson Close, Edmonton,


Alberta T6R 2J8, Canada

Richard J. McMullen, Jr., Mobile Technology Company, PO Box 650232, Dallas, TX 75265-
0232, USA

Daniel F. Merriam, Kansas Geological Survey, University of Kansas, Lawrence, KS 66047,


USA

Richard K. Nishimori, Mobile Technology Company, PO Box 650232, Dallas, TX 75265-


0232, USA

Jeffrey A. Nunn, Department of Geology and Geophysics, Louisianna State University, Baton
Rouge, LA, 70803, USA

Harald S. Poelchau, Institut fUr Chemie und Dynamik of the Geosphare - 4,


Forschungszentrum Jiilich, D-52425 Jiilich, Germany

M. Thompson, BP Norge UA, PO Box 197, 4033 Forus, Norway: present address: BP
Exploration Operating Company Ltd., Middlesex, United Kingdom

D.H. Welte, Institut fUr Chemie und Dynamik of the Geosphare - 4, Forschungszentrum
Jiilich, D-52425 Jiilich, Germany

Libo Zhang, Department of Geology and Geophysics, Louisiana State University, Baton
Rouge, LA 70803, USA

C. Zwach, Norsk Hydro NS, PO Box 200, N-1321 Stabekk, Norway


INDEX

Abikh mud dispir, 204 DTS (see Distributed optical fiber


AkcessBasin™ software package, 120 Temperature Sensing)
Alberta Basin, 83, 84, 87, 88, 89, 95, 229 dynamic evolution modeling, 207
Alberta, Canada, 63
analytical models, 110 Eastern Mobile Belt (New Zealand), 158
Anadarko Basin, Oklahoma, 18 Elmworth Field, Canada, 229
Arkansas, 120 empirical correction factors, 47
Arkoma Basin, 120 equilibriam temperature, 3
erosion, 134
Basin analysis, 152, 161, 164,207 exhumation, 157, 160, 172
basin hydrodynamics, 95 experimental data, 223
basin modeling, 219
best-fit model, 197 Fluid flow, 3, 71, 88,118,125, 167
Black Hills, South Dakota, 111, 113 Fourier's law, 42, 119
BHTs (also see bottom-hole
temperatures), 65, 76, 86, 155, 161 Gamma-ray logs, 14,24
bottom-hole temperatures (also see gas saturation, 225, 232
BHTs), 2, 36,45, 55, 107, 154, 172 GEOPETH, 207
burial history, 182, 190, 194,230 geothermal gradient, 65, 92, 101
geothermal gradient map, 67
Canada, 83 geothermal regime, 89, 95
Cherokee Basin, 37, 38 Gjallar Ridge, Norway, 132, 141
climate warming, 103 gradient logs, 17
commercial temperature data, 37 gradient-type models, 185
convection, 9 gravity modeling, 165
cratonic basins, 75 Great Plains, 100, 108, 114
crustal thinning, 146 groundwater flow, 85, 109, 113

Darcy velocity, 82, 125 Heat capacity, 90


diapir geometry, 204 heat flow, 36
dimensional analysis, 91 heat-flow model, 85, 185
Distributed optical fiber Temperature heat transport, 2
Sensing (DTS), 6 Homer plot, 92, 155
drillstem-test temperatures (DSTs), 36, Hot Water Belt, 100
47 hydrocarbons, 118
hydrocarbon maturation, 144
hydrological model, 72

239
240 INDEX

hydrogeological regimes, 87 Rayleigh number, 83


hydro stratigraphic model, 121 rock matrix conductivities, 221
Hylton oil field, Kansas, 56
Sedimentary basin, 2, 36, 82, 95, 118,
Industrial temperature data, 63 152
insulating shales, 123 seismic refraction data, 133
intracratonic basin, 178 sensitivity models, 187
intrusion modeling, 165 shale conductivity, 31
isostatic subsidence, 136 Shiwassee County, Michigan, 181
Siberian Platform, 227
Kansas, 37, 38 signal and noise, 47
sonic logs, 17
Lateral groundwater flow, 77 sonic velocity, 31
Little Ice Age, 103 South Caspian Basin, 204, 215
logging speed, 7 South Dakota, 100, 104, 106, 113
stretching factors, 136
Magmetic intrusion, 140 subcrustal stretching, 141
magmatic underplating, 137 subsidence, 224
Maui field, New Zealand, 158 subsidence analysis, 136
modeled formation temperature, 49
modeled heat flow, 141 Taranaki Basin (New Zealand), 152, 160,
modeled temperature, 46 172
modeling thermal conditions, 2 tectonic subsidence, 136, 141
modeled thermal gradient, 71 temperature depth plots, 49
Michigan, 178 temperature distributions, 125
Michigan Basin, 178, 180 temperature gradient, 31, 105, 172, 179
mixing formulae, 220 temperature gradient logs, 31
mud diapirs, 204 temperature history, 220
mud volcanoes, 208 temperature maps, 53
temperature logs, 3
Nebraska, 100, 104, 112 temperature logging tools, 5
Nemaha Anticline, 41, 44 thermal conductivity, 2, 43, 68, 75, 90,
New Zealand, 152 103, 118, 127, 157, 182
numerical model, 118, 183 thermal conductivity of fluids, 221
numerical simulations, 118 thermal disturbance, 7
thermal equilibrium, 6
Ozark Dome, 120 thermal gradient, 2, 127
thermal gradient log, 32
Pacific Plate, 170 thermal history simulations, 163
paleo water-depths, 135 thermal insulation, 126
partial water saturation, 223 thermal logs, 14
Peclet number, 82 thermal model, 42, 138, 164, 182
pore fluid, 225 thermistor probe, 102
porosity, 222 thermal regime, 4, 140
porosity-depth relations, 134, 157, 158 "true" formation temperature, 6, 36
Postle-Hough field, Oklahoma, 18 T -z profiles, 103
precision temperature logs, 6, 10, 32, 65
pressure, 223
INDEX 241

Underplating model, 138 Well-log temperature, 45


uniform basement heat flow, 75 West Ranch field, Texas, 11
Urengoy field, NW Siberia, 227 West Siberian Basin, 227
U.S. Midcontinent 39 Western Canadian Sedimentary Basin,
62,231
Vering Basin, Norway, 132 Willows-Beehive Bend gas field,
Vigrid Syncline, 141 California, 26
vitrinite reflectance data, 182, 185,229

You might also like