You are on page 1of 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/314100394

Interrelation between surface and basement heat flow in sedimentary


basins

Article  in  AAPG Bulletin · October 2017


DOI: 10.1306/12051615176

CITATIONS READS

10 546

3 authors:

Alban Souche Daniel Walter Schmid


Bergverk AS University of Oslo
29 PUBLICATIONS   316 CITATIONS    97 PUBLICATIONS   1,767 CITATIONS   

SEE PROFILE SEE PROFILE

Lars Ruepke
GEOMAR Helmholtz Centre for Ocean Research Kiel
111 PUBLICATIONS   3,560 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

elasto-plastic deformation around inflating magma chambers View project

Tectonics View project

All content following this page was uploaded by Daniel Walter Schmid on 06 September 2019.

The user has requested enhancement of the downloaded file.


Interrelation between surface AUTHORS

and basement heat flow in Alban Souche ~ GeoModelling Solutions,


Hardturmstrasse 120, 8005 Zürich,
Switzerland; present address: Physics of
sedimentary basins Geological Processes, Department of
Geosciences, University of Oslo, Postboks
Alban Souche, Daniel Walter Schmid, and Lars Rüpke 1048 Blindern, 0316 Oslo, Norway; alban.
souche@geo.uio.no
Alban Souche is a postdoctoral researcher at
the University of Oslo, Norway. His work
ABSTRACT focuses on the development of numerical tools
The results of petroleum system models (PSM) critically depend for basin modeling. His interest is in exploring
on the computed evolution of the temperature field. Because and quantifying the coupling between
physical processes active in geological systems.
PSM typically only resolve the sedimentary basin and not the en-
tire lithosphere, it is necessary to apply a basement–heat-flow Daniel Walter Schmid ~ GeoModelling
boundary condition inferred from well data, surface–heat-flow Solutions, Hardturmstrasse 120, 8005 Zürich,
measurements, and an assumed tectonic scenario. The purpose of Switzerland; Physics of Geological Processes,
this paper is to assess the use of surface–heat-flow measurements Department of Geosciences, University of
to calibrate basin models. We show that a simple relationship Oslo, Postboks 1048 Blindern, 0316 Oslo,
Norway; daniel.schmid@geo.uio.no
between surface and basement heat flow only exists in thermal
steady state and that transient processes such as rifting and sediment Daniel Walter Schmid received his Ph.D. in
geology from Eidgenosgische Technische
deposition will lead to a decoupling. We study this relation-
Hocheschule (ETH), Zurich, Switzerland. He
ship in extensional sedimentary basins with a one-dimensional,
works as a consultant for GeoModelling
lithosphere-scale finite element model. The numerical model was Solutions, which he combines with a part-time
built to capture the large-scale dynamic evolution of the litho- researcher position at the University of Oslo,
sphere and simultaneously solve for transient thermal processes Norway. His main research interest is the
in basin evolution, such as sedimentation, compaction-driven fluid flow, development of more realistic basin and
and seafloor temperature variations. Our analysis shows that several petroleum system models.
corrections need to be applied when using surface–heat-flow infor- Lars Rüpke ~ GEOMAR, Helmholtz Centre
mation for the calibration of basement heat flow in PSM. Not doing so for Ocean Research Kiel, Wischhofstraße
can lead to significant errors of up to 30°C–50°C (86°F–122°F) at 1-3, D-24148 Kiel, Germany; lruepke@
typical petroleum-reservoir and source-rock depths. We further show geomar.de
that resolving sediment-blanketing effects in basin modeling is crucial, Lars Rüpke is a professor at the GEOMAR
with the thermal impact of sediment deposition being at least as im- Helmholtz Centre for Ocean Research in Kiel,
portant as rifting-induced basement–heat-flow variations. Germany, where he leads the Seafloor
Modeling group. Hismain research interests are
marine hydrothermal systems, sedimentary
basin evolution, and porous transport
INTRODUCTION phenomena. He received his Ph.D. in
geophysics in 2004 from Kiel University
Deciphering the thermal evolution of sedimentary basins from and was a postdoctoral researcher at Oslo
observables, such as the seismic stratigraphy and well data, is one University’s Center for the Physics of
Geological Processes from 2005 to 2007.

Copyright ©2017. The American Association of Petroleum Geologists. All rights reserved. Green Open ACKNOWLEDGMENTS
Access. This paper is published under the terms of the CC-BY license.
We would like to thank the reviewers
Manuscript received September 8, 2015; provisional acceptance February 1, 2016; revised manuscript
received June 3, 2016; revised manuscript provisional acceptance September 8, 2016; 2nd revised manuscript D. K. Higley, K. E. Peters, K. B. Trivedi, and
received November 3, 2016; final acceptance December 5, 2016. S. G. Henry and the AAPG Editors M. L. Sweet
DOI:10.1306/12051615176

AAPG Bulletin, v. 101, no. 10 (October 2017), pp. 1697–1713 1697


and B. J. Katz for their constructive of the primary objectives of basin modeling. Various key pro-
comments. We would also like to thank the cesses in sedimentary basins are temperature controlled. On
Exploration and Production Division of Eni in a basin scale, examples include source-rock maturation (Pepper
Milan, Italy, for the stimulating discussions and Corvi, 1995), quartz cementation (McBride, 1989),
that eventually led to this paper.
and smectite to illite transformation (Hower et al., 1976). These
are relevant for petroleum system analysis, because they affect the
EDITOR’S NOTE generation of hydrocarbons, fluid pressure development, and
A color version of Figure 9 can be seen in the reservoir quality. On the lithosphere scale, the temperature field
online version of this paper. controls postrift subsidence (McKenzie, 1978), strain partitioning
(e.g., Huismans and Beaumont, 2011; Brune et al., 2014), and
mantle melting (e.g., Koopmann et al., 2014). Constraining the
temperature evolution in the subsurface is therefore crucial for
general understanding of the geological evolution of a region of in-
terest and the assessment of the hydrocarbon potential in particular.
Two main data sources exist that can be used to constrain
temperature: well data and surface–heat-flow measurements. Well
data, where available, provide a direct measure of the present
temperature state in the subsurface. However, usually only
bottomhole temperatures and a few additional measurements
in potential reservoir units are collected. This introduces bias,
because the thermal material properties of reservoir rocks are
different from other rocks in the system, which results in
a different local temperature gradient. Well temperatures may
also not be representative of basin geological history. Other
temperature data, such as vitrinite reflectance (e.g., Burnham
and Sweeney, 1989) and apatite fission tracks (e.g., Hendriks et al.,
2010), may give additional constraints. However, they only provide
a “blurred” picture, because they depend on temperature as well as
time and do not yield a unique temperature evolution.
Surface–heat-flow measurements are simple and inexpensive
and may offer additional insights. This is also true offshore, where
gravity-driven probes can be used to penetrate the first few meters
of sediment (Bullard, 1954; Beardsmore and Cull, 2001). The
question is how much information about the present and past
temperature field of a basin can be extracted from surface–heat-
flow measurements.
Modeling the thermal evolution of sedimentary basins in-
volves solving the transient heat equation, which requires that
initial and boundary conditions be specified. Petroleum system
models (PSM) are designed to resolve sedimentation processes
within a specific region of interest where the heat equation is
solved at basin scale. A crustal (basement)–heat-flow boundary
condition must be prescribed at the bottom of the model, which
can be approximated from lithospheric-scale geodynamic models
and calibrated by well data. Several geodynamic models can be
used for this purpose, ranging from the classic McKenzie model or
modifications of it (McKenzie, 1978; Jarvis and McKenzie, 1980;
Waples, 2001) to fully dynamic margin models (e.g., Huismans
and Beaumont, 2011; Brune et al., 2014). However, these

1698 Interrelation between Surface and Basement Heat Flow in Sedimentary Basins
lithospheric-scale geodynamic models often neglect Table 1. List of Symbols Used in This Study
sedimentation processes. Sediments usually have
Symbol Definition and Unit
lower thermal conductivity than crustal rocks, which
changes the steady-state geotherm, thereby causing Q Heat flow (W/m2) (with [W] = [kg$m2/s3])
a reduction in basement heat flow and a slowdown of Qb Basement heat flow (W/m2)
postrift cooling (Zhang, 1993; Theissen and Rüpke, Qs Surface heat flow (W/m2)
2010). Furthermore, the sedimentation rate may Qtr Transient correction term (W/m2)
impact the thermal transient of the system and Qx Referring to either Qs or Qb in equation 10
contribute to a depression of the geothermal gra- S Radiogenic heat from the basin (W/m2)
dient near the surface (De Bremaecker, 1983; ter k Thermal conductivity (W/[m$K])
Voorde and Bertotti, 1994; Wangen, 1995; Rüpke r Density (kg/m3)
et al., 2013). Finally, as we will show below in more cp Heat capacity (J/kg/K) (with [J] = [kg$m2/s2])
detail, radiogenic heating within the sediments changes ðxÞb Bulk property of x term
the lithospheric geotherm and depresses basement ðxÞs Solid property of x term
heat flow. These processes are often collectively re- ðxÞf Fluid property of x term
ferred to as sediment blanketing. Their cumulative u Porosity (—)
u0 Surface porosity (—)
effect is quite complex and results in a nontrivial re-
l Compaction constant (m)
lationship between surface and basement heat flow.
vs Solid velocity (m/s)
This has important consequences for basin models
vf Fluid velocity (m/s)
in that surface–heat-flow measurements may be
VD Darcy flux (m/s)
difficult to use for model calibration. In addition,
Hr Solid radioactive heat production (W/m3)
basement–heat-flow curves are not easily transfer-
T Temperature (°C) (with [°C] = [K] - 273)
able from lithosphere-scale geodynamic models to
T0 Surface temperature (°C)
basin-scale PSM. DT Temperature difference from geotherms calculated
Direct access to the sediment–basement interface with different basement heat flow (°C)
is generally not possible, thus requiring alternative t Time (s)
ways to infer the temperature boundary condition at z Depth (m)
this interface. One possibility is to try to use surface– zb Basement depth (m)
heat-flow measurements to constrain the basement–
heat-flow condition. In thermal steady state, the
If we could accurately estimate the radioactive heat
heat flow measured at the surface of a sedimentary
production in a sedimentary column, then we could
basin is the sum of the basement heat flow and the
use surface–heat-flow measurements to estimate
cumulative heat generated within the sediments
basement heat flow. Figure 1 presents calculated
(Turcotte and Schubert, 2014). Under the assump-
values of S for different amounts of radioactive heat
tion that the decay of radioactive elements is the only
production in basins with differing lithologies and
heat source within the sedimentary matrix, the cu-
thicknesses. For reasonably deep sedimentary basins
mulative radiogenic heat, S, in the sediments can be
and frequently assumed radioactive-heat-production
expressed as
values, the differences between surface and basement
ðh heat flow are significant.
 
S= Hr 1 - uðzÞ dz (1) The condition of thermal steady state is, however,
0 fairly exceptional. Transient thermal perturbation
can be induced by various factors at the basin scale, such
where u is porosity, z is depth, and Hr is radioactive
as sedimentation, erosion, rift-related heat input, or
heat production. All variables are explained in Table 1
pore-fluid circulation (e.g., Benfield, 1949; Hutchison,
for equations 1–11. To recover the basement heat
1985; Wangen, 1994, 1995; Person et al., 1996;
flow, Qb , from surface heat flow, Qs , we therefore
Souche et al., 2014). The question is to what extent
need to subtract S from Qs :
these transient effects disturb the balance between
Qb = Qs - S (2) basement and surface heat flow. We therefore

SOUCHE ET AL. 1699


 
 b ¶T  f f  s s ¶T
rcp + u rcp v + ð1 - uÞ rcp v
¶t ¶z
 
¶ ¶T
- k = Hr ð1uÞ (4)
¶z ¶z
where
 b  f  s
rcp = u rcp + ð1 - uÞ rcp (5)

A Lagrangian mesh is used to track the displacement


of the solid material (sediments and lithospheric
units); equation 3 can be expressed in terms of the
material derivative ðDt
D
= ¶t¶ + vs $=Þ, where the ad-
vection term simplifies to account only for Darcy
flow, VD :

VD = uðv f - vs Þ (6)
 
Figure 1. Heat-flow contribution (S) from the sediment ra-  b DT  f ¶T ¶ ¶T
dioactive heat source vs. basin thickness for different basin-fill rcp + rcp VD - k = Hr ð1 - uÞ(7)
Dt ¶z ¶z ¶z
lithologies and different amounts of matrix radioactive heat
production (Hr). See Table 2 for lithology-specific data. Qb =
The diffusion and advection terms in equation 6 are
basement heat flow; Qs = surface heat flow.
split and solved separately using a fractional step ap-
proach, where diffusion is solved implicitly and fluid
introduce a transient–heat-flow correction term, Qtr , advection is solved explicitly using a backward char-
and an augmented version of equation 2: acteristic Euler scheme. Darcy’s flux is calculated from
fluid mass conservation in sediments that compact as
Qb = Qs - S + Qtr (3) a function of burial depth. The potential buildup of
In the following parts of the paper, we numeri- overpressure and its effect on matrix porosity is ne-
cally solve the thermal evolution of the entire glected in our model. This assumption is reasonable for
lithosphere with coupled basin formation and the study of surface heat flow that is mostly sensitive to
sedimentation processes to assess the contribution the compaction of highly porous (and permeable)
of Qtr . This analysis will identify the dominant surface sediments in the shallow part of the basin. Heat
sediment-blanketing effects that need to be con- flow can be extracted from the temperature field at any
sidered when using surface–heat-flow measure- depth by Fourier’s law:
ments for basement–heat-flow calibration and when
¶T
transferring basement–heat-flow evolution curves QðzÞ = -kðzÞ (8)
from lithosphere-scale–geodynamic to basin-scale ¶z
PSM. This definition of the heat flow provides an effec-
tive estimate coherent with surface measurement
techniques performed with heat-flow probes (Bullard,
1954; Beardsmore and Cull, 2001).
METHODS AND MODEL SETUP

Governing Equations Lithospheric-Scale Model

We employ a one-dimensional diffusion–advection All the numerical results presented in this study were
temperature equation to describe the lithosphere determined using a lithospheric-scale model defined
temperature field. It assumes thermal equilibrium by a sedimentary basin, an upper crust, a lower crust,
between the solid and the fluid phase in porous and a lithospheric mantle. The model extends to the
media: base of the lithosphere, also referred to as “lithosphere

1700 Interrelation between Surface and Basement Heat Flow in Sedimentary Basins
depth,” where the bottom thermal boundary is set to
1300°C (2372°F). Two scenarios of lithosphere
depth (120 and 90 km [75 and 56 mi]) are compared
in the Results section.
In the case of lithospheric thickening, the depth
of the thermal boundary extended to the base of the
lithosphere. In the case of rifting, the depth of the
thermal boundary was the defined lithosphere
depth level. Upwelling of the lithospheric mantle
unit was then balanced by introducing asthenospheric
material at the base of the model. Asthenospheric
material was assumed to have the same properties as
the lithospheric mantle with an initial temperature of
1300°C (2372°F). Figure 2. Left: porosity–depth trend for shale, siltstone, and
The thermal and material properties of each unit sandstone lithologies. Right: effective conductivity based on the
are presented in Table 3. The mesh resolution was porosity–depth trend for water-filled sediments with geometric
averaging between the matrix/water conductivities.
approximately 100 m (328 ft) throughout the do-
main and refined to 2 m (6 ft) in the sedimentary part.
as opposed to more elaborate models such as the
As a consequence, the surface heat flow calculated
temperature-dependent Sekiguchi model (Sekiguchi,
from our numerical model is an approximation based
1984; Whittington et al., 2009). The Sekiguchi model
on the mean conductivity and the linear temperature
tends to give lower rock conductivity values than
gradient over the uppermost two meters. The time
those chosen in this study and leads to relatively lower
resolution was adaptive and ranged from 500 to
heat flow throughout the lithosphere. We systemati-
15,000 yr, depending on the sedimentation rate.
cally compared our numerical results with both
conductivity models and have observed only a relative
change in the absolute value of the heat flow; the
Basin-Fill Porosity and Conductivity
trends in heat-flow evolution were not affected.
We investigated three different basin lithologies Therefore, using either a constant matrix conductivity
(sandstone, siltstone, and shale) with distinct com- or temperature-dependent model does not change the
paction laws and matrix thermal conductivities (Figure conclusions presented here regarding the effect of
2; Table 2). For simplicity, we assumed the same sedimentation processes on heat-flow evolution.
matrix density for all (Table 3). The porosity–depth
trend of each lithology follows Athy’s law (Athy,
1930) and is expressed as: RESULTS
ð- lz Þ
u = u0 e (9)
Steady-State Heat Flow in Isostatic Equilibrium
where u0 is the surface porosity, l is the compaction
constant, and z is the burial depth below the seafloor. In the Introduction section, we demonstrated how
The effective conductivity of the sediments was surface and basement heat flow are related under
calculated using the geometric average between the thermal steady-state conditions. We did not consider
conductivity of the pore water and the sediment
Table 2. Porosity–Depth Parameters for the Different Litholo-
matrix. The effective radioactive heat production was
gies (Hantschel and Kauerauf, 2009)
calculated based on the radioactive heat production
of the sediment matrix (Hr ) scaled by the solid vol- Lithology Surface Porosity u0 Compaction Constant l (m [ft])
ume ratio ð1 - uÞ. The resulting trends of porosity and
Shale 0.70 1205 (3953)
effective conductivity are shown in Figure 2.
Siltstone 0.55 1961 (6433)
The conductivity model used in this study is
Sandstone 0.41 3226 (10,584)
relatively simple, with constant matrix conductivity

SOUCHE ET AL. 1701


Table 3. Thermal and Material Properties Used for the Different Units of the Model

Density Heat Capacity Conductivity Radioactive Heat Initial Lower Coefficient of Thermal
Unit (kg/m3) (J/kg/K) (W/m/K) Production (W/m3) Boundary (km [mi]) Expansion (1/°C)

Water 1000 4210 0.56 — — —


Sediments 2700 837 2.2*, 2.5†, 3.5‡ 0 to 1.5 · 10-6 0 0
(matrix)
Upper crust 2700 837 2.5 1 · 10-6§ 15 (9) 3.28 · 10-5¶
Lower crust 2950 837 2.5 4 · 10-7§ 30 (19) 3.28 · 10-5¶
Mantle 3280 837 3.5 2 · 10-8§ 90 to 120 (56 to 75) 3.28 · 10-5¶
lithosphere

*Shale; †Siltstone; ‡Sandstone; §After Hasterok and Chapman (2011); ¶After Parsons and Sclater (1977).

the issue of isostatic equilibrium, which requires that relates basin and crust geometry to the stretching
the weight of every lithospheric column be identical factor.
at the isostatic compensation level. In a system in local Figure 5A, C shows how surface and basement
isostatic equilibrium, the thickness of the lithospheric heat flow vary as a function of basin thickness as-
mantle, the crust, and the basin are not independent. suming thermal steady state and local Airy isostasy
Figure 3 shows a sketch of Airy isostasy for a litho- when the depth of the lithosphere is set to 120 km
spheric column with and without a sedimentary basin. (75 mi). The plot in Figure 5A essentially reproduces
In the latter case, the crust was thinned by a factor b Figure 1, but now we can also analyze the absolute
value of Qb and Qs . We observe a decrease of the
(so-called stretching factor) and the created accom-
basement heat flow for increased basin thickness
modation space was filled with sediments.
caused by reduced radiogenic heat from the thinned
The average sediment density varies with infill
crust and the presence of the overlying sediments with
lithology as well as basin depth. Figure 4 shows how
a lower thermal conductivity. The effect of reduced
basin and crustal thicknesses are related for the dif-
radiogenic heat from a thinned crust is especially
ferent sediment lithologies that we consider. It also
significant, causing a reduction in basement heat flow
by up to 20% for a 10-km (6-mi)-deep basin.
Basement heat flow is progressively reduced with
increasing basin thickness, but surface heat flow
shows different behavior. Radiogenic heat within the
sediments can compensate for the reduced basement
heat flow. For reasonable values of radioactive heat
production in the sediments and relatively large basin
depth, the surface heat flow may be higher than at the
top of an unstretched crust. Radiogenic heat within
the sediments has, however, only limited impact on the
basement heat flow, which is slightly reduced for higher
degrees of radioactive heat production. Basement heat
flow is further dependent on the thermal conductivity
of the sediments (i.e., the basin infill) and decreases
with decreasing conductivity values (Figure 5C).
To provide a more intuitive measure of the
computed heat-flow values, we calculate the tem-
Figure 3. Configuration of the crust and lithosphere before and perature difference predicted by a model that assumes
after deposition of sediments. Airy isostasy was assumed with the correct basement–heat-flow Qb as the bottom
compensation depth at the base of the lithosphere. boundary condition and a model that uses Qs instead.

1702 Interrelation between Surface and Basement Heat Flow in Sedimentary Basins
situation becomes progressively more complex when
additional processes and transient conditions are
explored.

Effects of Sedimentation and


Compaction-Driven Fluid Flow

In environments characterized by relatively fast sed-


imentation rates, the burial of sediments (initially at
surface temperature) dominates over thermal dif-
fusion, which results in depression of the geother-
mal gradient. In such high-sedimentation settings,
compaction-driven fluid flow may also be important
(Hutchison, 1985) and can counteract the depression
of the geothermal gradient. To investigate this in-
terplay, we assess the transient thermal budget of the
Figure 4. Crustal geometry vs. beta factor for different basin-fill basin by integrating both the heat carried by the fluid
lithologies calculated from isostatic equilibrium under thermal and the solid phase within the sediments.
steady-state conditions. We modeled two scenarios where sediments are
added at relatively high sedimentation rates either on
Put differently, this is the error involved in using top of the basement (crust model 1) or in a preexisting
measured surface–heat-flow values as the bottom sedimentary basin of 5 km (3 mi) depth (crust model
boundary condition (top basement). A corresponding 2, see Table 4). Both models were initialized at
analytical expression is given, for example, in Wangen thermal equilibrium. Note that we focus here on the
(2010): effects of sedimentation and do not consider isostatic
ðz effects. Hence, the entire model subsides with the
Qx 1 added sediments and effectively becomes larger. We
TðzÞ = T0 + Qx du
kðuÞ used a basin fill of shale lithology without radioactive
ðz
0
ðzb ! heat source to isolate the effect of sedimentation and
1  
+ Hr 1 - uðwÞ dw dv (10) compaction-driven pore-fluid flow.
kðvÞ
0 v Figure 6A, C shows heat flow when 2 km (1 mi)
of sediments were added on the top of the crust model
ðz
Qs Q 1 1 and 2 and no fluid flow was accounted for in
DTðzÞ = TðzÞ - TðzÞb =S du (11)
kðuÞ the temperature equation. The sedimentation rates
0 ranged from 500 to 2000 m/m.y (1640 to 6562 ft/m.y.).
Figure 5B, D shows the resulting temperature dif- Figure 6B, D shows the corresponding models when
ferences from using the wrong value for the basement the pore-fluid movement was taken into account.
heat flow. For sedimentary basins deeper than 8 km After the sedimentation phase, we ran the model for
(5 mi), the temperature difference at an assumed up to 20 m.y. to investigate the thermal relaxation of
reservoir depth of 3 km (2 mi) is larger than 10°C this system.
(50°F) for reasonable radioactive heat sources in The first 1 m.y. was modeled as steady state with-
the sediments. This difference may exceed 15°C–20°C out sedimentation. Crust model 1 and 2 resulted in
(59°F–68°F) for units buried at source-rock depth different initial heat flows because of the different
(assumed at 4 to 5 km [2 to 3 mi] depth) and may overall geometry. Once sedimentation started, the
exceed 30°C (86°F) in basins that are deeper surface heat flow was rapidly reduced. The reason
than 12 km (7 mi). These results show that detailed for this is that at high sedimentation rates the solid
information on basin infill and geometry are necessary advection term in equation 7 becomes more im-
to correctly relate surface to basement heat flow. The portant and depresses the geothermal gradient. The

SOUCHE ET AL. 1703


Figure 5. (A) Surface (Qs ) and basement (Qb ) heat flow at steady state for a shale basin fill with various amounts of matrix radioactive
heat production (Hr ). (B) Temperature difference (DT) obtained assigning Qs instead of Qb at the base of the basin using the values shown
in (A). (C) The Qs and Qb at steady state for a shale, siltstone, and sandstone basin fill with fixed matrix Hr of 1 · 10-6 W/m3. (D) The DT
obtained by assigning Qs instead of Qb at the base of the basin using the values shown in (C).

most extreme case is when sediments were deposited Depositing the same amount of sediment in the
at 2000 m/m.y (6562 ft/m.y.). in crust model 1 crust model 2 results in less reduction for both the
without considering the effect of fluid flow (Figure surface and the basement heat flow (Figure 6C). For
6A). Here the heat flow drops by approximately 50% the surface heat flow, the reduction is explained by
to 25 mW/m2. Once sedimentation stopped, heat flow compaction of the 5 km (3 mi) of sediments initially
recovered, and after 10 m.y. a new steady state was present in the basin. Their continued compaction leads
asymptotically approached. That heat-flow value is less to increased thermal conductivity and thermal gradient
than the initial heat flow because the same tem- across the unit, which balances part of the surface–heat-
perature boundary conditions were applied at the flow depression from sedimentation. For the basement
top and bottom of a lithospheric column that grew heat flow, the lesser reduction compared with the crust
throughout these experiments. Depending on sedi- model 1 is explained by greater basement depth.
mentation rate, the surface heat flow was reduced by Figure 6B, D shows that the upward flow of pore
10% to 50%. fluids in compacting sediments is an important process

1704 Interrelation between Surface and Basement Heat Flow in Sedimentary Basins
Table 4. Depth to the Base of Each Unit Defining the Crust 120 (Figure 8, left panels) and 90 km (75 and 56 mi)
Model 1 and 2 (Figure 8, right panels), which correspond to the
depths of the prescribed bottom temperature of the
Sediment UC LC ML
model (1300°C [2372°F]). The depth of the LAB
Model (km [mi]) (km [mi]) (km [mi]) (km [mi])
eventually upwells from the initial lithosphere depth
Crust model 1 0 15 (9) 30 (19) 120 (75) during the synrift stretching phase and slowly returns
Crust model 2 5 (3) 15 (9) 30 (19) 120 (75) to its original depth during the postrift thermal-
cooling phase.
Abbreviations: LC = lower crust; ML = mantle lithosphere; UC = upper crust.
Figure 8A, B shows the temperature evolution of
during sedimentation that counteracts the reduc- the crust and basin during the synrift and postrift (up
tion of the surface heat flow. The reason is that the to 30 m.y.). The isotherms in the basement are char-
fluid-advection term in equation 7 contains the acterized by a significant rise during the synrift, re-
Darcy flux, which points in the opposite direction flecting heating resulting from lithospheric thinning
as the burial of the sediments. This relatively and burial along the geothermal gradient caused by
warmer upward fluid flow affects mainly the sur- sedimentation. The rising basement temperature is,
face heat flow and does not have a significant effect however, not reflected by an increase in basement heat
on basement heat flow. flow (Figure 8C, D, blue curve). Instead of rising, the
The differences between the results obtained basement heat flow rapidly decreases in the early
with or without considering pore-fluid flow are synrift. This somewhat counterintuitive behavior results
shown in Figure 7. The surface heat flow is strongly from sediment-blanketing effects (i.e., the burial of
influenced by pore-fluid flow, and the effect increases thermally nonequilibrated sediments to depth and the
with increasing basin thicknesses until a maximum is lower sediment conductivity) and the thinning of
reached for basin thicknesses larger than the com- the crustal radiogenic layer. Our model results are
paction constant l of the sediments (Figure 7B). opposite to the prediction of the McKenzie model
shown in Figure 8C, D (green curve). The transient
effect is even more pronounced in the surface–
Rifting-Induced Heat-Flow Variations heat-flow curve, which shows a larger depression
than the basement heat flow (Figure 8C, D, red
During rifting, the lithosphere–asthenosphere bound- curve).
ary (LAB) rises, which causes a transient increase in During the transition from synrift to postrift,
heat flow. Here we study the formation of a rift basin surface heat flow increases rapidly. This is a result of
that forms by Airy isostasy in response to thinning of a slowdown in sedimentation, which accommodates
the crust and the lithosphere. The thinning is assumed the postrift cooling subsidence. Although Qs is lower
to be uniform within the crust and the lithosphere, than Qb during synrift, the situation appears re-
whereas it is ignored in the sediments. The thermal versed in all but the earliest postrift. To facilitate the
effects of sedimentation and compaction-driven pore- comparison between the surface and the basement
fluid flow are modeled during the evolution of the heat flow, we subtract the contribution of sediment
rifting process. radiogenic heat, which results in the dashed curve
in Figure 9A, B. This corrected surface–heat-flow
curve lies consistently below the basement–heat-
Rift Basin Fully Filled with Sediments flow curve, emphasizing the blanketing effect of
We first study the end-member scenario where the sediment deposition. Toward the end of the simu-
accommodation space is fully filled with sediments lation runs, the corrected surface–heat-flow curve
during rifting, as opposed to the McKenzie model converges toward the basement curve, indicating that
that does not account for the thermal effects of the shallow system has almost completely equili-
sedimentation (McKenzie, 1978). The model results brated. Note that the time scale of the surface–heat-
shown in Figure 8 were obtained with a rift duration flow transient (<5–10 m.y.) is shorter than that of the
set to 5 m.y. and a stretching factor b = 2. Two thermal relaxation of the lithosphere after rifting
lithosphere depth scenarios were investigated at (~60 m.y., Jarvis and McKenzie, 1980).

SOUCHE ET AL. 1705


Figure 6. Surface– and basement–heat-flow evolution when 2 km (1 mi) of sediments were added on the top of crust model 1 and 2
(shale lithology without radioactive heat production). The sedimentation started at 1 m.y. and lasted until 2 km (1 mi) of sediments were
deposited. The model was run for 20 m.y. (A) Crust model 1 without fluid flow, (B) crust model 1 with fluid flow, (C) crust model 2 without
fluid flow, and (D) crust model 2 with fluid flow.

A detailed analysis of the transient effects in these larger amount of sediments accumulated during
systems is shown in Figure 9B, C, where Qtr is plotted rifting. Increasing the rift duration at constant b
(solved using equation 3) for the 5 and 10 m.y. rift reduces the amplitude of the transient–heat-flow
event simulation runs. Following a gradual buildup, effect but reproduces the overall behavior.
Qtr reaches maximum values at the synrift and postrift An estimate of the temperature effect of the
boundary, before rapidly dropping during the postrift. difference between Qs and Qb is provided in Figure
In a model with the base at 120 km (75 mi), most of 9E, F, where we compute the temperature difference
the transient disappeared only 2 to 3 m.y. after rifting between a model with the correct heat-flow condition
ceased. In a model with the base at 90 km (56 mi), at the basement–sediment interface and one where
the transient effects last longer in the shallow parts of we use Qs - S, i.e., neglect the transient term Qtr .
the system, i.e., 5 to 10 m.y., which is because of the The differences reach more than 20°C (68°F) at the

1706 Interrelation between Surface and Basement Heat Flow in Sedimentary Basins
Figure 7. (A) Fluid-flow contribution on surface heat flow (Qs ) and basement heat flow (Qb ) vs. the sedimentation rate for the crust
model 1 and 2. (B) Fluid-flow contribution on Qs vs. the initial basin thickness for a sedimentation rate of 1000 m/m.y (3281 ft/m.y.).
The fluid-flow contribution is the ratio of the heat-flow variation calculated with and without fluid flow at the end of sedimentation (see
Figure 6).

end of rifting in the 120-km (75-mi)-deep model flow during synrift (basement and surface heat flow)
and up to 50°C (122°F) in the 90-km (56-mi)-deep is strongly attenuated by the presence of sediments,
model. even for relatively thin sedimentary cover. There-
fore, the presence of sediments should always be
Rift Basin Partially Filled with Sediments considered when evaluating the evolution of the
Our previous results only considered the sedi- basement heat flow during rifting.
mentary end-member scenarios where the accom-
modation space is either fully filled with sediments Surface-Temperature Variations
(100% of sediments) or sediment starved following
McKenzie model approach (0% of sediments). To So far, we have kept the top and bottom boundary
assess different sedimentary scenarios during rifting, condition in our lithospheric-scale thermal model
we also run models where the total thickness of the fixed. However, it is clear that the top boundary
sediments was set to a given percentage of the ac- condition also varies through time. Daily, seasonal,
commodation space created during rifting (0%, and climatic changes may influence surface temper-
20%, 40%, 60%, 80%, or 100%). ature, more so with direct subaerial exposure than in
Figure 10 presents the evolution of the basement systems with water cover that buffers the air tem-
heat flow and the surface heat flow corrected for the perature variations (Beardsmore and Cull, 2001). It is
sediments’ radiogenic heat during synrift and postrift. noteworthy that shallow seas, such as the North Sea,
The difference between these two quantities defines show significant seasonal variations in sea bottom
the thermal transient of sedimentation, as illustrated temperatures (Skjoldal, 2006), resulting in a transient
in detail in Figure 9. As expected, the largest thermal thermal regime in the shallow sediments. The prop-
transient occurs for the fully filled end-member agation of a surface-temperature signal into the
scenario. For lower proportions of sediments, we subsurface is affected by attenuation and phase shift.
observed that the thermal transient is less pro- The depth at which a periodic signal has decayed to
nounced during rifting and can be practically ne- 1=e (0.37) of the surface amplitude is called the skin
glected when sediments represent less than 50% of depth and is often used as the depth of the thermal
the total accommodation space created. However, transient in the subsurface. The skin depth can be
an interesting observation is that the maximum heat estimated analytically for cyclic, sinusoidal surface-

SOUCHE ET AL. 1707


Figure 8. Model results for a rift duration of 5 m.y., b = 2, and a shale basin-fill lithology with radiogenic heat production set to 1 ·
10-6 W/m3. Lithosphere depth set to 120 (left panels) and 90 km (75 and 56 mi) (right panels). (A, B) Temperature (T ), (C, D) heat
flow, and (E, F) sedimentation rate as function of time. The surface (Qs ) and basement (Qb ) heat-flow values are compared with the
basement heat flow obtained using a McKenzie-type model in which no sedimentation occurs during rifting. S = cumulative
radiogenic heat in the sediments.

1708 Interrelation between Surface and Basement Heat Flow in Sedimentary Basins
Figure 9. Model results as presented in Figure 8 for (A, B) heat flow as function of time, (C, D) transient surface–heat-flow contribution
(Qtr ), and (E, F) temperature difference (DT) introduced by Qtr . Qb = basement heat flow; Qs = surface heat flow; S = cumulative radiogenic
heat in the sediments.

SOUCHE ET AL. 1709


Figure 10. Basement heat flow (Qb ) and surface heat flow corrected for radioactive heat source (Qs - S) for different proportions of
sediments in the accommodation space created during rifting. Lithosphere depth set to (A) 120 and (B) 90 km (75 and 56 mi). Model setup
parameters given in caption for Figure 8.

temperature variations, assuming a diffusive system cooling and warming that is induced by the back-
of constant rock conductivity and constant back- ground geotherm in the sediments. Although the
ground temperature (Table 5). Surface signal var- temperature anomaly decreases rapidly below the
iations with periods of a year or less only affect the skin depth (Figure 11A, B), the amplitude of the heat-
uppermost few meters, but signals with longer flow anomaly remains relatively large compared with
wavelengths can reach substantial depths and the background heat flow (Figure 11C, D). For shale,
affect the subsurface over long time periods. This the heat flow within the first 5 m (16 ft) of sediments
has been used by Huang et al. (1997) and several varies by several orders of magnitude compared with
later contributions to derive the temperature the background heat flow of 47 mW$m-2. At 5 m (16
evolution for the last 20,000 yr based on the ft) depth (Figure 11C), the heat flow remains dom-
onshore heat-flow well database of the Interna- inated by the surface-temperature signal and ranges
tional Heatflow Commission. from +38 to -28 mW$m-2 around the background
The depth at which Bullard-type probes pene- value. The heat flow is essentially at steady state
trate unconsolidated surface sediments in offshore below 10 m depth, where it only ranges from +1 to
basins is typically 5 (Bullard, 1954) to 11 m (16 to -4 mW$m-2 around the background value. For
36 ft) depth (Lewis et al., 1991), which appears to sandstone and siltstone, which have greater bulk
be within the skin depth range of 1–10 yr surface- conductivity, the temperature variations penetrate
temperature variations. Using our lithospheric- deeper into the sediments and result in a much larger
scale model, we reassess the effect of seasonal heat-flow anomaly (Figure 11D) compared with
surface-temperature variations on the temperature
and heat flow. The results are presented in Figure 11
Table 5. Periodic Surface Signal Propagation in a Diffusive
for the first 10 m (33 ft) of the sediments. The
System with a Rock Thermal Diffusivity of 10-6 m2s-1 and
calculated skin depths in our model, ranging from
Constant Background Temperature
1.5 to 2.7 m (4.9 to 8.8 ft), are lower than that
presented in Table 5 because of the nonuniform Period 24 hr 1 yr 101 yr 102 yr 103 yr 104 yr 105 yr
diffusivity of the sediments with depth in our
Skin depth 0.2 3.2 10 32 100 317 1002
model, whereas the expression of the analytical
(m)
solution assumes uniform rock diffusivity. We notice
an asymmetry of the skin depths calculated during The skin depth is the depth at which a periodic signal has decayed to 1/e (0.37).

1710 Interrelation between Surface and Basement Heat Flow in Sedimentary Basins
Figure 11. Temperature and
heat-flow variations at depth in-
duced by a sinusoidal surface-
temperature signal of period 1 yr
and amplitude –1°C (33.8°F). (A)
Temperature variations (shale
only), (B) temperature anomaly
from the background geotherm,
(C) heat-flow variations (shale
only), and (D) heat-flow anomaly
from the background heat flow.
The crust model 2 setup was used
(as defined in Tables 2, 4) with
5 km (3 mi) of sediments on the
top of the crust and no radioac-
tive heat production in the sedi-
ments. The mesh resolution was
refined to 1 · 10-2 m in the first
50 m (164 ft) of the sediments,
and the time step was reduced to
3 days. The maxima curves pre-
sented in (A)–(D) were obtained
after 10 annual cycles starting
from steady state, which was
enough to reach convergence.

shale. The amplitudes of the temperature and heat- surface heat flow to provide an estimate for the
flow anomalies in Figure 11 scale accordingly to the basement heat flow.
amplitude of the surface-temperature variations (set The transient differences, Qtr , between the base-
here to –1°C [33.8°F]). ment and surface heat flow (corrected for radiogenic
heat source within the sediments) can be relatively
large but are also relatively short lived in the sce-
DISCUSSION AND CONCLUSIONS narios that we study here. Sedimentation-driven
transient effects last for a few million years and
Surface and basement heat flow in sedimentary ba- have almost entirely disappeared 10 m.y. after
sins are generally different, and a simple relationship sedimentation stops. For rifting, the transient ef-
between the two only exists at thermal steady state. fects are only relevant for 2–3 m.y. if the initial
Even this, however, requires detailed knowledge of lithosphere depth is 120 km (75 mi) and 5–10 m.y.
the radiogenic heat source from the sedimentary if the initial lithosphere depth is 90 km (56 mi). It is
column. In all other cases, it is challenging to use the important to distinguish this basin-scale thermal

SOUCHE ET AL. 1711


relaxation from lithospheric-scale processes, be- United Kingdom, Cambridge University Press, 324 p.,
cause it shows a much larger characteristic time doi:10.1017/CBO9780511606021.
Benfield, A., 1949, The effect of uplift and denudation on
scale of approximately 60 m.y. for thermal relax-
underground temperatures: Journal of Applied Physics,
ation and related postrift subsidence (Jarvis and v. 20, p. 66–70, doi:10.1063/1.1698238.
McKenzie, 1980). Brune, S., C. Heine, M. Perez-Gussinye, and S. V. Sobolev,
The PSM only resolve the temperature field within 2014, Rift migration explains continental margin asym-
the sedimentary unit. As a consequence, a basement– metry and crustal hyper-extension: Nature Communi-
cations, v. 5, p. 4014–4023, doi:10.1038/ncomms5014.
heat-flow boundary condition has to be imposed at the Bullard, E., 1954, The flow of heat through the floor of the
top basement level. Here we have shown that two Atlantic Ocean: Proceedings of the Royal Society of
main problems exist in doing so. First, a correction for London A: Mathematical, Physical and Engineering Sci-
radiogenic heat within the sedimentary unit must be ences, v. 222, p. 408–429, doi:10.1098/rspa.1954.0085.
applied to relate information from surface to basement Burnham, A. K., and J. J. Sweeney, 1989, A chemical kinetic
model of vitrinite maturation and reflectance: Geo-
heat flow. Omitting this correction can result in mis- chimica et Cosmochimica Acta, v. 53, p. 2649–2657,
calculation of the temperature by up to 30°C (86°F) at doi:10.1016/0016-7037(89)90136-1.
the reservoir level. Second, the evolution of basement De Bremaecker, J.-C., 1983, Temperature, subsidence,
heat flow is strongly affected by sediment-blanketing and hydrocarbon maturation in extensional basins: A
effects and variations in the rift-related heat input. As finite element model: AAPG Bulletin, v. 67, no. 9, p.
1410–1414.
shown by Waples (2001), surface heat flow during the Hantschel, T., and A. I. Kauerauf, 2009, Fundamentals of
synrift phase may be characterized by an initial de- basin and petroleum systems modeling: Berlin Heidelberg,
pression caused by the thinning of the radioactive Springer-Verlag, 476 p.
crustal layer. Our model, which includes the effect of Hasterok, D., and D. Chapman, 2011, Heat production and
geotherms for the continental lithosphere: Earth and
sedimentation, exemplifies further the potential de-
Planetary Science Letters, v. 307, p. 59–70, doi:10.1016
pression of the surface heat flow during synrift as op- /j.epsl.2011.04.034.
posed to that predicted by the popular McKenzie Hendriks, B., P. Osmundsen, and T. Redfield, 2010, Normal
(1978) model. faulting and block tilting in Lofoten and Vesterålen
Finally, we have shown that variations in surface constrained by apatite fission track data: Tectonophysics,
v. 485, p. 154–163, doi:10.1016/j.tecto.2009.12.011.
and/or seafloor temperature have a strong effect on Hower, J., E. V. Eslinger, M. E. Hower, and E. A. Perry, 1976,
surface heat flow, causing further decoupling of it from Mechanism of burial metamorphism of argillaceous sedi-
basement heat flow. The heat flow within the thermal ment: 1. Mineralogical and chemical evidence: Geological
transient zone may deviate by several orders of mag- Society of America Bulletin, v. 87, p. 725–737, doi:10.1130
/0016-7606(1976)87<725:MOBMOA>2.0.CO;2.
nitude from the background value and is strongly af-
Huang, S., H. N. Pollack, and P. Y. Shen, 1997, Late Qua-
fected by surface-temperature variations well below ternary temperature changes seen in world-wide conti-
the characteristic skin depth. Offshore surface– nental heat flow measurements: Geophysical Research
heat-flow measurements usually only sample the Letters, v. 24, p. 1947–1950, doi:10.1029/97GL01846.
uppermost meters of sediments, which render Huismans, R., and C. Beaumont, 2011, Depth-dependent
extension, two-stage breakup and cratonic underplating
these measurements prone to seasonal variations in
at rifted margins: Nature, v. 473, p. 74–78, doi:10.1038
bottom-water temperature. Shallow seas, such as /nature09988.
the North Sea, exhibit several degrees of seasonal Hutchison, I., 1985, The effects of sedimentation and compaction
bottom-water temperature variations, which may on oceanic heat flow: Geophysical Journal Interna-
dominate transient heat flow within the first 10 m tional, v. 82, p. 439–459, doi:10.1111/j.1365-246X
.1985.tb05145.x.
(33 ft) of sediments. Jarvis, G. T., and D. P. McKenzie, 1980, Sedimentary basin
formation with finite extension rates: Earth and Planetary
Science Letters, v. 48, p. 42–52, doi:10.1016/0012-821X
(80)90168-5.
REFERENCES CITED Koopmann, H., S. Brune, D. Franke, and S. Breuer, 2014,
Linking rift propagation barriers to excess magmatism at
Athy, L. F., 1930, Density, porosity, and compaction of volcanic rifted margins: Geology, v. 42, p. 1071–1074,
sedimentary rocks: AAPG Bulletin, v. 14, no. 1, p. 1–24. doi:10.1130/G36085.1.
Beardsmore, G. R., and J. P. Cull, 2001, Crustal heat flow: Lewis, T., W. Bentkowski, and J. Wright, 1991, Thermal state
A guide to measurement and modelling: Cambridge, of the Queen Charlotte Basin, British Columbia: Warm,

1712 Interrelation between Surface and Basement Heat Flow in Sedimentary Basins
in G. J. Woodsworth, ed., Evolution and hydrocarbon from western Norway: Geofluids, v. 14, p. 58–74, doi:
potential of the Queen Charlotte basin: Ottawa, Ontario, 10.1111/gfl.12042.
Canada, Geological Survey of Canada, p. 489–506. ter Voorde, M., and G. Bertotti, 1994, Thermal effects of
McBride, E. F., 1989, Quartz cement in sandstones: A review: normal faulting during rifted basin formation, 1. A finite
Earth-Science Reviews, v. 26, p. 69–112, doi:10.1016 difference model: Tectonophysics, v. 240, p. 133–144,
/0012-8252(89)90019-6. doi:10.1016/0040-1951(94)90268-2.
McKenzie, D., 1978, Some remarks on the development of Theissen, S., and L. Rüpke, 2010, Feedbacks of sedimentation
sedimentary basins: Earth and Planetary Science Letters,
on crustal heat flow: New insights from the Vøring Basin,
v. 40, p. 25–32, doi:10.1016/0012-821X(78)90071-7.
Norwegian Sea: Basin Research, v. 22, p. 976–990.
Parsons, B., and J. G. Sclater, 1977, An analysis of the variation
Turcotte, D. L., and G. Schubert, 2014, Geodynamics:
of ocean floor bathymetry and heat flow with age: Journal
Cambridge, United Kingdom, Cambridge University
of Geophysical Research, v. 82, p. 803–827, doi:10.1029
/JB082i005p00803. Press, 636 p.
Pepper, A. S., and P. J. Corvi, 1995, Simple kinetic models Wangen, M., 1994, Numerical simulation of thermal con-
of petroleum formation. Part I: Oil and gas generation vection in compacting sedimentary basins: Geophysical
from kerogen: Marine and Petroleum Geology, v. 12, Journal International, v. 119, p. 129–150, doi:10.1111
p. 291–319, doi:10.1016/0264-8172(95)98381-E. /j.1365-246X.1994.tb00918.x.
Person, M., J. P. Raffensperger, S. Ge, and G. Garven, 1996, Wangen, M., 1995, The blanketing effect in sedimentary
Basin-scale hydrogeologic modeling: Reviews of Geo- basins: Basin Research, v. 7, p. 283–298, doi:10.1111
physics, v. 34, p. 61–87, doi:10.1029/95RG03286. /j.1365-2117.1995.tb00118.x.
Rüpke, L. H., D. W. Schmid, M. Perez-Gussinye, and Wangen, M., 2010, Physical principles of sedimentary basin
E. Hartz, 2013, Interrelation between rifting, faulting, analysis: Cambridge, United Kingdom, Cambridge Uni-
sedimentation, and mantle serpentinization during versity Press, 527 p., doi:10.1017/CBO9780511711824.
continental margin formation—Including examples from Waples, D. W., 2001, A new model for heat flow in exten-
the Norwegian: Geochemistry Geophysics Geosystems, sional basins: Radiogenic heat, asthenospheric heat, and
v. 14, p. 4351–4369, doi:10.1002/ggge.20268.
the McKenzie model: Natural Resources Research, v. 10,
Sekiguchi, K., 1984, A method for determining terrestrial heat
p. 227–238, doi:10.1023/A:1012521309181.
flow in oil basinal areas: Tectonophysics, v. 103, p. 67–79,
Whittington, A. G., A. M. Hofmeister, and P. I. Nabelek,
doi:10.1016/0040-1951(84)90075-1.
2009, Temperature-dependent thermal diffusivity of the
Skjoldal, H. R., 2006, Update report on North Sea con-
ditions—2nd quarter 2006: ICES/EuroGOOS North Sea Earth’s crust and implications for magmatism: Nature,
Pilot Project–NORSEPP ICES/EuroGOOS Planning v. 458, p. 319–321, doi:10.1038/nature07818.
Group for NORSEPP: Bergen, Norway, Planning Group Zhang, Y. K., 1993, The thermal blanketing effect of sediments
on the North Sea Pilot Project, 26 p. on the rate and amount of subsidence in sedimentary
Souche, A., M. Dabrowski, and T. Andersen, 2014, Modeling basins formed by extension: Tectonophysics, v. 218,
thermal convection in supradetachment basins: Example p. 297–308, doi:10.1016/0040-1951(93)90320-J.

SOUCHE ET AL. 1713


View publication stats

You might also like