You are on page 1of 17

FetKin: Coupling

kinematic restorations and


temperature to predict
thrusting, exhumation histories,
and thermochronometric ages
Ariel Almendral, Wilmer Robles,
Mauricio Parra, Andrs Mora,
Richard A. Ketcham, and Michael Raghib
ABSTRACT
FetKin is a C++ program for forward modeling thermochronological ages on a two-dimensional geological cross section.
Modeled ages for various thermochronometers are computed from
timetemperature histories that result from coupling the modeled
kinematics of deformation obtained from commercial software
for balanced reconstructions (2DMove) and a finite element
computation of temperatures. Additional capabilities include the
ability to accommodate (1) a smooth change of topological relief;
(2) the influence of variation in rock physical properties; and
(3) multikinetic modeling of fission-track ages and length distributions, as well as apatite and zircon (U-Th)/He and muscovite
40
Ar39 Ar systems. A joint first-order analysis of the impact of
erosion parameters and material properties improves age predictions and allows for a more complete analysis of observed cooling
ages based on their modeled thermal histories. Thus, this paper
presents a new software tool that has been developed as a basic
support for the methodological approach used to build the
kinematic restorations shown in this volume, which are the basic
input for petroleum systems modeling and prediction in the
Colombian Eastern Cordillera and Llanos foothills basin.

INTRODUCTION
Knowledge about the timing of deformation of geological structures is critical to understanding virtually all geological processes
Copyright 2015. The American Association of Petroleum Geologists. All rights reserved.
Manuscript received August 1, 2011; provisional acceptance December 10, 2012; revised manuscript
received August 2, 2013; revised manuscript provisional acceptance October 2, 2013; 2nd revised
manuscript received March 7, 2014; final acceptance July 7, 2014.
DOI: 10.1306/07071411112

AAPG Bulletin, v. 99, no. 8 (August 2015), pp. 15571573

1557

AUTHORS

Ariel Almendral Instituto Colombiano


del Petrleo, Piedecuesta, Santander,
Colombia; present address: Norsk
Regnesentral/Norwegian Computing Center,
P.O. Box 114, Blindern, NO-0314 Oslo,
Norway; Ariel.Almendral.Vazquez@nr.no;
aalmendral@gmail.com
Ariel Almendral is a senior researcher at
Norwegian Computing Center in Oslo,
Norway, where he develops geostatistical
software for the O&G industry. He has also
contributed to the development of the Petrel
framework in the area of three-dimensional
geology and modeling. His Ph.D. work from
the University of Oslo (2005) dealt with
numerical computation of financial
derivatives.
Wilmer Robles Instituto Colombiano del
Petrleo, Piedecuesta, Santander, Colombia;
wilmerangel@gmail.com
Wilmer Robles is an external consulting
geophysicist at Instituto Colombiano del
Petrleo-Ecopetrol, with an M.Sc. degree in
geophysics from the Universidad Nacional
de Colombia. His interests include fault-bend
folding, thermochronology, and the
acquisition, processing, and interpretation of
geophysical data.
Mauricio Parra Instituto Colombiano del
Petrleo, Piedecuesta, Santander, Colombia;
Jackson School of Geosciences, The
University of Texas at Austin, Austin,
Texas 78759; m.parra@iee.usp.br
Mauricio Parra is an external research
consultant for Instituto Colombiano del
Petrleo-Ecopetrol. He received a Ph.D. in
geosciences from the University of Potsdam
in 2009, followed by a postdoctoral stay at
the Department of Geosciences, University
of Texas at Austin. He has worked in the
tectonic evolution of Colombian orogens
using thermochronometry and basin
analysis.
Andrs Mora Instituto Colombiano del
Petrleo, Piedecuesta, Santander, Colombia;
andres.mora@ecopetrol.com.co
Andres Mora is a senior geologist at Instituto
Colombiano del Petrleo-Ecopetrol. He

earned a B.Sc. degree in geology from the


Universidad Nacional de Colombia in 1999
and a Ph.D. in geological sciences from the
University of Potsdam (summa cum laude)
in 2007. His research interests include
structural geology, thermochronology, and
basin analysis.
Richard A. Ketcham Jackson School of
Geosciences, The University of Texas at
Austin, Austin, Texas 78759; ketcham@jsg
.utexas.edu
Richard Ketcham is an associate professor in
the Jackson School of Geosciences at the
University of Texas at Austin. He received his
B.A. degree in geology and computer
science from Williams College in 1987 and
his Ph.D. in geological sciences from the
University of Texas, Austin in 1995. His active
research interests include thermochronology
and geological applications of highresolution x-ray computed tomography.
Michael Raghib Instituto Colombiano del
Petrleo, Piedecuesta, Santander, Colombia;
michael.raghib@gmail.com
Michael Raghib is an external senior
research consultant at Instituto Colombiano
del Petrleo-Ecopetrol. He received his Ph.D.
in applied mathematics from the University
of Glasgow in 2006, followed by
postdoctoral stays at the Princeton
Environmental Institute at Princeton
University and the Applied Mathematics
group (T-5) at Los Alamos National
Laboratory. His interests include
mathematical biology and geosciences.
ACKNOWLEDGEMENTS

This study was a part of the project


Cronologa de la deformacin en las
Cuencas Subandinas, a project of the
Instituto Colombiano del Petrleo, the
research division of Ecopetrol S.A. We thank
W. F. Jaimes for his help with validation of
examples of varying rock properties using
the Finite Element software ANSYS.

1558

including those related to petroleum systems. Usually, determination of the timing of deformation relies on features directly
observed on folds and faults. This is particularly the case with
growth strata (Anadn et al., 1986; Medwedeff, 1989; Suppe et al.,
1991; Vergs et al., 1996; Suppe et al., 1997; Novoa et al., 2000;
Shaw et al., 2004; Strayer et al., 2004) where the geometry, rates,
and kinematics of folding can be directly assessed when preservation and exposure permit. However, in many situations, the
growth strata record has been erased by erosion, and it becomes
necessary to use alternative, indirect methods to assess the timing
and rates of deformation for individual structures. This is the case
of the settings like the Colombian Eastern Cordillera where restorations have to rely on limited to absent growth strata, which
poses a challenge for accurate petroleum system modeling in
mature fold-and-thrust-belt settings. Therefore, an alternative
approach incorporating thermochronometric methods has to be
developed. If those methods are correctly used, a more precise
prediction of the timings of oil migration, generation, and trapping
can be achieved.
One approach is to evaluate cooling histories resulting from
hot material at depth being exhumed toward the cool surface of
the earth. By keeping track of this cooling history with the help
of thermochronometers, it is possible to obtain approximate estimates of the times over which rocks remained at temperatures
cooler than a certain threshold temperature. The basic data
obtained from thermochronology are, therefore, cooling ages.
Ideally, from cooling ages, it is possible to obtain exhumation histories or the timing when thrusting began in a given fault block or
structural domain (Gans et al., 1991; Stwe et al., 1994; Ehlers
and Chapman, 1999; Ring et al., 1999; Ehlers et al., 2003;
Ehlers, 2005; Stockli, 2005, Eichelber et al., 2013). Simple calculations are possible by assuming a paleogeothermal gradient
(Sobel and Strecker, 2003), but this is not straightforward in situations where there are rapid shortening rates or strong contrasts in
local topographic relief. In such cases, isotherms markedly deviate
from the simplest planar shape, resulting in perturbed patterns that
are not easy to predict (Mancktelow and Grasemann, 1997; Huerta
and Rodgers, 2006; Lock and Willett, 2008).
Given the complications engendered by even these limited
departures from an ideal situation, it is easy to imagine that it is
even more difficult to derive deformation rates from exhumation
histories in yet more complex, realistic geological contexts.
Prime examples are thrust belts with complex structural geometries, such as the Canadian Rockies (Price, 1981), Taiwan
(Suppe, 1980a, b), and the sub-Andean basins of South America
(Baby et al., 1997; Martnez, 2006; Mora et al., 2006; Espurt et al.,
2008).

Coupling Kinematic Restorations and Temperature

The many challenges in deriving wellconstrained shortening rates from exhumation rates
may explain why such studies are quite rare and
why it is, in fact, very unusual to have thermochronometric data fully incorporated into balanced cross
sections and the subsequent construction of kinematic
restorations. However, in such geological settings,
cooling ages could represent the best data available
to derive past deformation rates. In previous studies
where thermochronological data were used to produce kinematic restorations (Mora et al., 2008,
2010), derivation of shortening rates relied on planar
isotherms and simplified assumptions concerning
past topographic relief. This approach, though useful
in interpreting the deformation history in certain simple structural settings, might become misleading in
situations where isotherms bend due to mass movement and where the sample path experiences reburial
or lateral movement due to complex and interconnected structures (Batt and Braun, 1999; Batt and
Brandon, 2002; Bollinger et al., 2004; Herman
et al., 2010).
A more detailed computational model that calculates heat conduction and advection due to development of structures and/or erosion and landscape
evolution can enable more precise interpretation of
measured ages. Early and simple examples tying thermal computations to thermochronometric data include
models of erosion and landform periodicity (Stwe
et al., 1994) and detachment faulting (Ketcham,
1996). As computational power has increased, such
efforts have become more ambitious and powerful.
The PeCube software package (Braun, 2003) simulates development of complex landscapes using a
three-dimensional thermal solution for erosion, which
can also incorporate simple structures in the subsurface. To allow simulation of more complex and
detailed structures, Lock and Willett (2008) linked a
thermal solver to a commercial structural reconstruction package (2DMove by Midland Valley,
Ltd.) to examine the thermal development along cross
sections of the Taiwan thrust. For a survey of additional related resources, see Ehlers et al. (2005).
FetKin expands upon the approach of Lock and
Willett (2008) by being designed so that it can be
incorporated into a section balancing workflow using
any commercial balancing software (i.e., 2DMove,

LithoTect, Dynel, etc.) and thermochronological


field data for calibration. It enables a more detailed
interpretation of data and geology that provides a
quantitative and accessible means to support or refute
the expert but subjective interpretation of the geologist. FetKin enables calculation of the changes in the
thermal configuration of an area under rapid, complex deformation and evolving topographic relief
(understood as in Huerta and Rodgers, 2006) and,
therefore, makes the calculation of past exhumation
rates easier and more precise. The workflow
associated with FetKin is iterative, based on twodimensional balanced cross sections, using as a starting point a kinematic restoration obtained classically,
that is, without the constraints of a detailed thermal
model (i.e., assuming planar isotherms). This provides an initial velocity field that is subsequently fed
into the finite element method (FEM) thermal model
that advects the isotherms according to the geological
displacements. In addition, FetKin enables prediction
of cooling ages at any user-specified position for
different thermochronometers in a given structural
cross section. Thus, with the possibility of forward
modeling ages, FetKin may provide hints on where
to sample when looking for specific timing information like activation of faults, duration of the deformation, and rate of motion, provided that previous
thermochronological data are available.
FetKin is written in C++ with an objectoriented approach to facilitate its maintenance and
future development. At the time of writing of this
paper, the program is invoked from the command line
with the input being an extensible markup language
(XML) file. This text file contains a set of keywords
that specify initial and boundary conditions, sample
data, topography data, and kinematic data in the form
of a sequence of files containing the interpreted grid
evolution in time. The FetKin first performs a quality
control test of the input grids to detect inconsistencies
such as large velocity jumps or nonmatching grid
points. Once the moving grid is accepted as a valid
kinematic model, the program proceeds to compute
temperatures in a time-stepping loop. Finally, ages
for specified thermochronometers are determined
from the timetemperature paths for the input samples. An important feature is that velocity is an input
parameter that the user can vary at each geological
ALMENDRAL ET AL.

1559

time step. In particular, we include the concept of


grid-timing or geological time for each particular grid
in the time-stepping process so we may simulate scenarios of tectonic activity or quiescence. Moreover, a
set of user-defined topographies can be incorporated
into the computations to account for the topographic
effect on the temperature. The timetemperature
paths produced by FetKin can be used as an initial
guess in HeFTy (Ketcham, 2005) for inverse modeling and data fitting.
The calculated thermochronometric ages in
FetKin use state-of-the-art algorithms and kinetics
for various thermochronometers. These include modeling of apatite fission-track (AFT) annealing that
accounts for compositional variation in annealing
kinetics (e.g., Ketcham et al., 2007), a detail
neglected in most modeling efforts that can have a
pivotal effect on data interpretation (e.g., Mora et al.,
2010).
For visualization purposes, several MATLAB
routines have been implemented which automatically
load the finite element grids, the 2DMove sections,
and the sample locations to produce a report sequence
for each time step. This permits a more complete and
efficient analysis of the effect of material properties,
erosional relief changes, and sedimentation into the
calculated age profile.
In this paper, we show how critical it is to incorporate in a standard kinematic restoration workflow,
completely or partially based on thermochronology,
the behavior of the isotherms during deformation. In
a companion paper (Mora et al., 2015, this issue),
we show the precise workflow used to apply FetKin
with actual geological data during a kinematic restoration. In an additional paper (Snchez et al., 2015,
this issue), cross sections where the FetKin approach
was used are the base for petroleum system modeling
in Petromod.

affecting the interpretation of observed ages (Stwe


et al., 1994; Mancktelow and Grasemann, 1997).
(2) Lateral advection of rocks causes lateral temperature variations. (3) Local contrasts in material properties are usually encountered associated with fault
planes. These factors call for a more advanced modeling of the thermal regime. We have, thus, implemented
the well-studied linear diffusionadvection equation in
two dimensions (2-D) that has consistently proven
useful in modeling the temperature distribution within
the earths crust. This equation needs to be solved
numerically in time and space to provide a complete
description of the temperature field. In this section,
we introduce our notation and then proceed to explain
the connection with a kinematic model.
The Two-Dimensional DiffusionAdvection
Equation
The FetKin solves numerically the transient advectiondiffusion equation in two dimensions (Carslaw
and Jaeger, 1986):


T
T
T
T
T
vx
vy
= k
+ k
+ H (1)
c
t
x
y
x x y y

where Tx; y; t is the temperature at location x; y at


time t, x; y is the (space-dependent) rock density,
kx; y defines the rock thermal conductivity tensor,
cx; y is the specific heat, v = vx ; vy is the velocity
of the moving grid, and H is the radioactive heat
production. For this equation to be well defined,
one needs to specify initial and boundary conditions.
The boundary conditions on the top and bottom
of the grid are defined as Dirichlet-type fixedtemperature conditions; see Braun et al. (2006) for a
similar notation:
Tt; x; y = Sx = T S y = T MSL + y ymax
Tx; y = ymin = T Basal

FETKINS TEMPERATURE MODEL


One-dimensional thermal models have proven to
have shortcomings in translating low-temperature
cooling ages into exhumation rates. (1) Topography
is known to vertically bend the low-temperature isotherms for moderate exhumation rates (>0.5 mm/yr),
1560

Coupling Kinematic Restorations and Temperature

(2)
(3)

where Sx is the topography line, T MSL is the


mean sea level (MSL) temperature, is the surface
temperatureelevation gradient (temperature change
with elevation), ymax is the MSL level, ymin is the base
level, T Basal is the basal temperature and T S y is the
top surface temperature. We assume no heat flux on
the left-hand and right-hand sides of the grid, that is,

Tx = 0 (Neumann-type boundary conditions). For


the initial condition,
Tt = 0; x; y = T 0 x; y

(4)

we use the steady-state solution of equations 13,


assuming no velocity, no heat production, and
negligible temperature gradient in the x-direction.
This can be summarized by means of the linear
relationship:
T 0 x; y = T S y

y ymin
y
y
+ T Basal max
ymax ymin
ymax ymin

(5)

This initial condition assumes no heat production, as


well as all subsequent modeling in this paper.
Although heat production can be an important variable for crustal-scale thermal modeling, for petroleum
systems where the depths of interest are relatively
shallow, the difference between heat generated within
the crust and heat that flows up from below is a
second-order consideration. Although it is not
reported in this paper, a Neumann-type (given by a
temperature flux instead of a constant temperature)
boundary condition for the bottom boundary is also
available in FetKin.

time-integration is of first-order accuracy in time


and is unconditionally stable, that is, no time-step
restriction is required for stability. We believe that
these choices of a first-order method in space and in
time suffice for our purposes of a first-order description of thermochronological ages.
In a typical geological setting, the computational
grid is bounded in the lateral direction by xmin and
xmax and in the vertical direction by ymin and the topography surface. The advection component incorporates
the direction of the velocity field at a given point of
the computational grid, and the magnitude of this
velocity affects the thermal regime. A sufficiently
dense cloud of particles that move in time according
to some kinematic law (fault parallel flow, trishear,
flexural slip, etc.) is used to define a velocity field;
see Figure 1 for a schematic view, where the particles
from the moving grid are represented as filled squares.
The purpose of the cloud of particles is manifold.
By construction, this set of points is implicitly used in
the algorithms of 2DMove to describe a deformation
rule. A sufficient number of frames or cross sections
throughout the tectonic activity will describe in detail

Adding Kinematics to the Temperature Model


We have adopted the methodology set forth by Lock
and Willet (2008), coupling the thermal solution with
a kinematic model generated from 2DMove, to find
the temperature distribution that honors a given
deformation mechanism. The temperature is computed on an Eulerian or computational grid with the
help of the FEM. This is a classical choice given its
versatility in coping with deformed geometries. The
nodes of the FEM grid are evenly positioned along
the lateral direction. In the vertical direction, the
nodes are evenly distributed from the topography line
to the bottom of the grid. The nodes are used to build
triangular elements where linear basis functions are
defined. The temperature field is obtained as a linear
combination of linear basis functions whose coefficients are found by solving a linear system where
the matrix and right-hand side vector are assembled
from element matrices. The time derivative has been
approximated by the implicit Euler scheme. This

Figure 1. Interpolation from moving grid (blue squares) onto


closest quadrature points in finite element method grid (red
discs). The closest moving grid point with respect to a quadrature
point could be located in a different triangle of the computational
grid; therefore, we omit in this figure some connecting arcs.
ALMENDRAL ET AL.

1561

the velocity field associated with the deformation and


will, therefore, be a good approximation for the
advection component. In addition to the velocity,
information on the thermal conductivity, the density,
and the specific heat can be also extracted from the
particles in the moving grid (see equation 1). The
velocity vector and physical properties from the closest particle are assigned to the closest quadrature
point in the computational grid, represented as discs
in Figure 1. Note that due to the nature of the FEM
method, properties are sampled at quadrature points
because the integration of inner-products of hat functions can be done exactly by quadrature rules (see
Zienkiewicz et al., 2005). We believe that this
approach is sufficient for this first-order method. An
important by-product of this strategy is that the set
of samples can be restored, yielding a set of trajectories in time for subsequent timetemperature history
computation. Numerical tests have been conducted
to verify that our results coincide with published
numerical examples from Lock and Willet (2008),
as shown in the section Numerical Experiments.
SETTING UP A FETKIN PROJECT AND
IMPLEMENTATION DETAILS
In this section, we outline the main steps to prepare a
FetKin project, with focus on the following aspects:
(1) construct a valid velocity field, (2) build a set of
topography lines with corresponding times, (3) assign
times to the moving grids, (4) define the sample locations, and (5) specify the thermochronometers for age
and length calculations. A brief summary of the spatial discretization method and time-integration is also
included.
Constructing 2DMove Moving Grids
The two primary tools used in 2DMove for tracing
positions throughout a geological history are lines,
usually corresponding to stratigraphic boundaries,
and points, which may be on lines or between them.
Our experience from the use of FetKin on real sections from the Colombian foothills has suggested that
a number of rules must be followed for proper construction of a moving grid that can be successfully
1562

Coupling Kinematic Restorations and Temperature

processed. Preparing a set of grids using 2DMove is


time-consuming, so one normally bypasses a number
of important details that are left to internal 2DMove
algorithms to sort out. However, for our purpose with
FetKin, we need to go further in the number of conditions we demand from a set of grids to be eligible to
run coupled with a thermal model. These rules help
in creating robust moving grids in 2DMove: (1) assign
a unique name to each line; (2) keep the same number
of lines for all grids; and (3) keep the same number of
points in each line.
Once the grids have been created, they can be
exported to an ASCII format and tested for consistency
using the batch subprogram QCSECTIONS. The
importance of using this tool lies mainly in detecting
possible human errors as well as locating and repairing
them. Note that consistency in the number of lines and
points is an essential step for achieving a consistent
velocity description. In the case of sedimentation,
where new lines are added to a cross section, the line/
point consistency must still be respected. In this case,
a dummy line can be placed outside the computational
grid that will have no effect on the temperature computation until it enters the grid. During the balancing
process, it becomes practically impossible to strictly
maintain a constant number of points. This is part of
the nature of the geological interpretation and workflow. However, if the geologist is aware that her/his
cross sections are going to be part of the FetKin
workflow, she/he can resample the edited lines in each
balancing step, preserving the geometry suggested, to
restore the initial number of points.
Changing Topography and Computational
Grid Adaption
The need for a more realistic approach to numerically
model the effect of a nonstatic topographic relief was
addressed in Braun (2003). The method implemented
in PeCube is based on an interpolation of the solution
obtained at each time step over a deformed, current
grid caused by a displacement of the top nodes.
PeCube makes use of an amplification factor in order
to augment or reduce the relief height. In FetKin,
we specify a set of topography lines with its corresponding time. The topography within a given time
interval is not available and has, therefore, to be

interpolated to deduce missing data. More precisely,


for a given time interval where the topography at time
t init has height yinit and at time t final the height is yfinal ,
the height at any t tinit ; t final ; is interpolated with
the help of the formula:
yt = yinit + tyfinal yinit

(6)

Here, t is a function given by




init
1 exp tt
topo


t =
t init
1 exp tfinaltopo

(7)

where topo is a positive parameter used to model


transitions between topographic lines. High values
of topo will give an almost linear interpolation
between two topography lines; whereas, low values
will speed up erosion of peaks for the most ancient
topography.
Moving Grid Timing and Time Interpolation
As mentioned earlier, FetKin can model periods of
tectonic activity and quiescence. When the practitioner constructs the sequence of sections, he/she is
only taking into consideration the displacement. In
order to assign times to each particular event, one
can enter a list of section versus time via a text file.
For the episodic example in the numerical experiments, such a list takes the form of Table 1.
The timing in Table 1 represents geological
times. For the episodic tectonic example, sections 21
and 22 are identical to simulate no motion over the
period 22 to 12 Ma. Next, the sections labeled from
1 until 21 simulate motion in the period from 12 to
10 Ma. Finally, the sections labeled 0 and 1 are identical to simulate negligible velocity in the interval
from 10 to 0 Ma. The number of time steps does not
need to be 22. Sections not defined for a time step
are interpolated linearly according to the closest moving grids in time. The sections interpolation ensures a
smoother transition of velocities across sections.

timetemperature development can be derived from


the moving grid. Given a data point, FetKin finds
the closest point from the final grid (most recent)
and traces this point back. To obtain a smooth age
curve, it is most advantageous to simply include the
samples in the moving grid as markers and do a full
restoration of these markers. Occasionally, sample
trajectories fall slightly outside of the computational
grid where temperatures have not been computed.
This happens, for example, when generating topographies using a numerical method that tends to smooth
out corners by adding artificial numerical diffusion.
A cure for this is to vertically offset the topography
by a minor increment so that those shallow samples
give meaningful temperatures. Another situation in
which problematic age holes may arise is where the
terrain has a hill slope angle greater than the angle
of shallow velocity vectors. In this situation, points
will necessarily drop out from the fixed grid during
a recent period and will reenter the grid at an older
stage, producing jumps in the age profile. Hence, as
a rule of thumb, hill slope angles should not exceed
the dip of the steepest fault.
Thermochronometer Calculations
The design of FetKin allows incorporation of any
number of thermochronometers. For the models presented in this paper, apatite fission-track ages, as well
as length distributions, are calculated using the
Ketcham et al. (2007) calibration. Apatite (U-Th)/He
ages are calculated based on the Wolf et al. (1998)
and Farley (2000) calibration for Durango apatite.
Other thermochronometers computations are also
available in FetKin. Zircon fission-track (ZFT) and
40
Ar39 Ar in muscovite (MAr) ages are calculated
according to the implementation by Braun (2003) that
uses the parameters from Brandon et al. (1998) for
ZFT and Hames and Bowring (1994) for MAr, based
on the framework by Dodson (1973).
Time Integration and the Finite Element
Method

Sample Data
Data can be sampled at arbitrary locations in the computational grid, provided sufficient information on

Equations 14 have been discretized in time by


the implicit Euler scheme in time and the FEM in
space. We now briefly outline the method; for a
ALMENDRAL ET AL.

1563

practical introduction, see Lynch (2004). Let u1 and


u0 be the temperature for the next time step and the
current time step, respectively. For simplicity, denote
by F the operator that only involves spatial derivatives in equation 1. The implicit Euler scheme can
be compactly written as
c

u1 u0
= F u1
t

(8)

The objective is to compute u1 implicitly, after


rearranging terms to arrive at an equation of the form
u0 = Gu1 . The next step, and most difficult, is to
discretize this equation in space. The method of
choice for spatial discretization is the FEM. First,
one needs a weak interpretation of the second-order
differential equation. Next, the region of interest, that
is, our computational 2-D grid, is triangulated. On
each triangle, three linear 2-D basis functions define
the so-called test functions (a linear function that
equals one on one vertex and vanishes on the other
two vertices). The numerical solution is set up as a
linear combination of these test functions, where the
unknown coefficients represent the temperature values at each node. This leads to a linear system on
the form u0 = Gu1 , where the matrix G is assembled
from element matrices whose entries are inner products of test functions for each element. The boundary
conditions are embedded in the matrix and the righthand side. The solution to this system of equations
yields the sought temperature for the next time step.
One then proceeds iteratively until all time steps have
been computed. We mention that no automatic timestep detection is done by FetKin. This entails that
rapidly changing topography surfaces demand a
smaller time step, which must be imposed by the user.

NUMERICAL EXPERIMENTS
We now conduct several numerical tests for
well-known geologically sound examples. The
purpose is to demonstrate the impact of the following
factors on the structure of the forward-modeled
cooling age: (1) time and velocity of deformation,
(2) rock properties, and (3) topography evolution.
For each of these factors, we have considered two
extreme scenarios, running from 22 Ma to the
1564

Coupling Kinematic Restorations and Temperature

present. The first scenario assumes a period of


tectonic quiescence from 22 to 12 Ma, followed by a
tectonically active period from 12 to 10 Ma, and concludes with a quiescent period from 10 until 0 Ma.
For future reference, we denote this situation as
episodic tectonic. The second scenario assumes
continuous motion throughout the complete running
time from 22 to 0 Ma. This situation will be referred
to as continuous tectonic. Summaries of these two
cases are shown in the columns of Figures 2 and 3,
respectively.
Example 1: Fault-Bend Fold on a Ramp with
Flat Topography and Homogenous Physical
Properties
We start by comparing results from FetKin with published results. This example tests the thermokinematics of a fault-bend fold as described in
Suppe (1980a, b, 1983). This model was used in
Lock and Willet (2008) for the fold-and-thrust-belt
example of the Taiwan Western Foothills, focusing
on the determination of the longterm deformation
rate. Note that unlike the example in Lock and
Willet (2008) where a heat flux is specified as a bottom boundary condition, here a Dirichlet-type condition as a basal temperature T Basal is assigned on the
bottom of the grid. The results are very similar for
these two boundary conditions because of the large
distance from the grid bottom to the detachment.
The computational grid has 80 km (49 mi) laterally and 30 km (18 mi) in depth. A single 30 dipping
fault ramp extends from 0 to 15 km (0 to 9 mi). The
top of the ramp is positioned at the central (zero) x
position and at a depth of 5 km (3 mi), the bottom is
placed at a depth of 15 km (9 mi). The rock layers
ride over the ramp from right to left, at a slip rate of
8 mm/yr (0.31 in./yr). A fold is created above
the upper kink of the ramp with lateral horizontal
transport of material. The rock-uplift rate equals
the denudation rate in this case because no relief is
being created. The material properties are assumed
constant throughout the grid and are as follows:
specific heat 1000 J kg1 K1 , thermal conductivity
2.5 Wm1 K1 , density 2500 kg m3 , basal temperature 600C (1112F), MSL temperature 0C (32F)
and no heat production. Finally, the grid spacing in

ALMENDRAL ET AL.

1565

Figure 2. Examples 1 to 3 assuming episodic tectonics: (A) 12 Ma, (B) 11 Ma, (C) 10 Ma, (D) 0 Ma, (E) AFT and AHe. The shaded part above the topography line indicates that this
portion has been eroded away. The first column from left to right corresponds to the homogeneous lithology, the second column alternates shales (green) with sands (white) 5-km
(3-mi) thick. These two cases have instantaneous removal of the topography via erosion; the last column shows the case where topography evolves on the same time scales as the
tectonic deformation events. The circles marked with A, B, and C are samples located on the surface at 3, 5, and 10 km (1.8, 3, and 6 mi). The points E at (3, 4) and
F (10, 4), marked with squares, are reference points. For this case, there is no relevant influence of the paleotopography on the ages.

1566

Coupling Kinematic Restorations and Temperature

Figure 3. Examples 1 to 3 explore the same situation as in Figure 2, but assume continuous tectonics: (A) 22 Ma, (B) 15 Ma, (C) 8 Ma, (D) 0 Ma, (E) AFT and AHe ages. The shaded
part above the topography line indicates that this portion has been eroded away.

Table 1. Timing Table for Episodic Tectonic in Examples


1 to 3
Grid Label

Timing
(Ma)

22 12

11.9

11.8

11.7

20

10.1

Table 2. Apatite Fission Track Ages for Samples A, B, and C


at Lateral Coordinates 3, 5, and 10, Respectively*

21 22
10

both directions is taken as 1 km (0.62 mi), and the


model has been run for 22 m.y. An inherited age of
8 m.y. was chosen arbitrarily so that the differences
between fully reset and inherited ages could be
clearly visualized.
The first column of Figures 2 and 3 shows the
results for the episodic and continuous tectonics,
respectively, together with the corresponding
(U-Th)/He and AFT ages. The cooling ages have a
distinctive U-shape, where younger ages are located
right above the ramp. The shape of the U varies
depending on the thermochronometer and on the
duration of the tectonic scenario. For the following
discussion, we use samples A, B, and C, located at
lateral coordinates 3, 5, and 10 km (1.8, 3, and
6 mi) and marked with circles, and samples E and F,
at coordinates (3, 4) and (10, 4) marked with
squares, as reference points. Note the significant difference in age for sample A in Figures 2-E1 and
3-E1, caused by its earlier cooling in the continuous
model. Summaries of calculated AFT ages (using
F-apatite kinetics) for all selected samples can be
found in Tables 2 and 3.
Example 2: Fault-Bend Fold on a Ramp with
Flat Topography and Nonhomogenous
Physical Properties
This example shows how physical properties can be
assigned to layers in the FetKin moving grid and
how these properties influence the relative positions
of the isotherms. The moving medium is considered
to be a sequence of alternating layers of sands and
shales, with a uniform thickness of 5 km (3 mi) each.
Note that these exaggerated layer dimensions are
chosen to demonstrate that AFT ages may feature significant lateral variations if material properties feature
sufficient contrast (Figures 2 and 3). The thermal
parameters within each layer are assumed constant,

Sample/
Example

1epi

1cont

2epi

2cont

3epi

3cont

A
B
C

12.2
11.45
11.61

18.03
12.28
12.37

16.80
11.87
17.29

21.01
12.11
16.27

12.44
11.93
11.73

17.2
17.89
19.34

*The notation epi is for the episodic tectonic activity in the interval 12 to 10 Ma
and cont stands for continuous tectonics during the period 22 to 0 Ma. The
abbreviation 2cont is read as example 2 with continuous tectonics.

Table 3. AFT Ages for Samples D and E at Coordinates


(2, 4) and (8, 4), Respectively*
Sample/Example

1epi

2epi

D
E

6.99
6.87

6.61
1.87

*The notation epi is for the episodic tectonic activity in the interval 12 to 10. The
abbreviation 2epi is read as example 2 with episodic tectonics.

but their values vary from layer to layer according


to mean depth and lithology. Material properties for
this example were computed with formulas in
Demongodin et al. (1991), Allen and Allen (2005),
and Nemcok et al. (2005) and are summarized in
Table 4.
Snapshots are displayed in the second column of
Figures 2 and 3. Compared to the case of homogenous materials, the temperature field in this case is
influenced by different layer properties, which results
in a different temperature field from 22 to 12 Ma.
This difference is most clearly seen in A2D2 as a
downward shift in the 120C (248F) isotherm,
which explains the older ages observed for samples
A and C, as compared to Example 1, and the narrowing of the U shape; see also Table 2. We note that
apatite samples are rarely found in shales, so this
example is extreme. However, another scenario (not
reported in this paper) that assumed shales with 10%
sand gave similar results.
Material properties have a posttectonic thermal
influence as well. To highlight this effect, we use
samples D and E, located at the same present-day
depth, which remain in the AFT PAZ after thrusting.
Their ages differ by about 5 m.y. (see Table 3) as a
result of the different steady-state temperatures in
ALMENDRAL ET AL.

1567

Table 4. Rock Properties for Example 2


Rock Type

Initial Mean
Depth (km)

Density
kg m3

Specific Heat
J kg1 K1

Thermal Conductivity
W m1 K1

Diffusivity
km2 myr1

Sandstone3
Shale2
Sandstone2
Shale1
Sandstone1

25
17.5
12.5
7.5
2.5

2649
2720
2622
2696
2238

973
837
993
854
1330

2.19
1.28
3.24
1.46
3.14

26.82
17.77
39.21
19.97
33.26

the sands and shales caused by their contrasting


material properties and geometric arrangement.
These differences cannot be accounted for in the
case of Example 1 with homogenous materials, where
a uniform cooling rate for these samples gives
approximately the same age. This post-tectonic effect
of materials is large in this case of episodic tectonics,
in contrast to our second scenario of continuous tectonics where this effect is not apparent.
Example 3: Fault-Bend Fold on a Ramp with
Changing Topography and Homogenous
Physical Properties
We now show how FetKin incorporates changing
relief. Here we have used the same moving grid from
Example 1. In the episodic case, topography creation
triggers erosion followed by 10 m.y. of only erosion.
For our second continuous case, a combination of
topography creation and erosion occurs throughout
the model execution. We mention that FetKin also
allows the definition of a series of intermediate topographies such that the missing ones are interpolated for
each time step. For details, see section Changing
Topography and Computational Grid Adaption.
The erosion causing topography change is simulated by a fluvial incision model (Whipple and Tucker,
1999). The elevation changes (h) in time and space
according to a model computed from the equation:
n
h
h
h
= vy + vx ke jxr jm
t
x
x

(9)

where xr is the distance from the drainage divide, k e is


an erodibility parameter, vx and vy are the velocity components of the particles close to the topographic boundary, and m and n are parameters that need to be
1568

Coupling Kinematic Restorations and Temperature

calibrated depending on the situation. The last component in this equation represents the erosion rate resulting
from a stream-power law. This law, although heuristic,
is useful to capture the interconnection between erosion
and orogeny formation, see Willett (1999) and Huerta
and Rodgers (2006). Empirical studies suggest that the
ratio mn is likely to be about 0.5 (Whipple and
Tucker, 1999; Whipple, 2004). Our tests assume
m = 0.5 and n = 1.
This model needs to be constructed in two steps.
First, we build the sequence of eroded topographies
using FetKin as a preprocessor only for topography
surface generation and not for temperatures. Once
the elevation model has been constructed, we use
2DMoves restoration algorithm to find the trajectories of the samples by restoring the samples at their
present topography. These samples are then added to
the moving grid as markers. We then prepare FetKin
for a second run that includes the previously computed topographies and the sample trajectories. The
deformed grid, the eroded topography, and the computed temperatures are displayed in the right-hand
columns of Figures 2 and 3. The resulting age
sequence now has a W-shape in the continuous case,
where older ages are located around the relief top, as
expected. For the episodic tectonic case, there is no
relevant influence of the paleotopography on the
ages. As proved by Stwe et al. (1994), the topography has a large impact on the temperature distribution
depending on the topography amplitude, the wavelength, and the denudation rate. For denudation rates
larger than 1 mm/yr (0.03 in./yr) (our episodic case
has 4 mm/yr [0.15 in./yr]), a topographic amplitude
of 3 km (1.8 km), and a wavelength of about 20 km
(12 mi), the apatite retention isotherm can be perturbed by up to 1 km (0.6 mi). This phenomenon is

observed in our episodic case, just after motion has


ceased. Fast erosion decreases the wavelength, which
added to overall cooling results in almost flat stationary isotherms, as exemplified in case D3 in Figure 2.
A comparison between the topography generated
in the episodic and continuous cases reveals a symmetric topography in the first case and an asymmetric
topography in the second case. As implemented in
our model, episodic tectonics create an almost symmetrical shape after 12 m.y. of running time, and
because no lateral change in erosion rate is assumed,
the final result is symmetrical. This is in clear contrast
with the continuous case, where a horizontal velocity
component is always present throughout the model
run, making erosion rates vary laterally. Even though
this is a simplified model that does not account for
material properties versus erosion rates, it illuminates
the influence of the topography creation process during tectonically active periods on the age profile.
The FetKin also allows straightforward production of ageelevation plots from model results
(Figure 4). For the episodic case, almost the whole
relief was eroded away, and so it is not shown. For
the continuous case, the slope in the age elevation
plot varies as a consequence of the asymmetry of the
topographic surface. At low-angle hillslope, the
topography cuts the rock units in such a way that
younger stratigraphic units or crustal levels may crop
out at topographically higher elevations. Therefore, in
certain cases, higher elevation does not correspond to

Figure 4. Ageelevation plot for Example 3 with changing


topography and continuous tectonics.

an older cooling age in the ageelevation plot. As we


advance uphill, this trend reverses to result in the
familiar positive slope.
Example 4: Fault-Bend Fold on a Listric Fault
This example considers a listric fault detaching at a
depth of 35 km (21 mi). This is a synthetic model
based on the geometries and depths to detachment
of an actual cross section reported by Mora et al.
(2008). The same model is tested in FetKin with
actual data in the companion paper by Mora et al.
(2015, this issue) because it appears to be the style
of intracrustal deformation that more closely resembles the deformation in the Colombian Eastern
Cordillera. The rates of thrusting, exhumation, depth
of erosion, and thickness of sedimentary rocks used
here are also derived from the study by Mora et al.
(2008). The fault moves at a rate of 0.5 km/m.y.
(0.31 mi/m.y.) throughout a running time of 30 m.y.
Again, fault-parallel flow is used to simulate the
motion of the hanging wall to the left of the fault.
The computational grid extends from 0 to 80 km
(0 to 49 mi), and the cell size is 1 km (0.6 mi) in each
direction. The modern topographic surface is
extracted from a digital elevation model of the
Farallones Anticline from the Cordillera Oriental
in Colombia, along the trace of the Mora et al.
(2008) cross section. The topography at the beginning of the run starts at sea level and amplifies linearly in time to reach the modern topography.
Samples have been placed along the topographic line
and restored using 2DMove for a full description of
the sample trajectory. Figure 5 displays the evolution
of the thermal field at 30 and 0 Ma. The AFT and
AHe ages are reset near the valley located at about
50 km (31 mi) and closer to the emerging fault
plane. One interesting feature of this profile is the
greater roughness of the AHe profile, as compared
to the AFT one, caused by the lower AHe closure
temperature, which is in some cases a desired property when capturing topographical effects using
low-temperature thermochronometers.
These numerical experiments can be used to
refine deformation rates in actual cross sections. For
example, Mora et al. (2008) constrained their kinematic restoration based on cooling ages and planar
ALMENDRAL ET AL.

1569

sophisticated kinematic modeling tools with the


time- and rate-resolving power of thermochronometric data.
CONCLUSIONS

Figure 5. Example 4, fault-bend fold on a listric fault: (A)


30 Ma, (B) 0 Ma, (C) AFT and AHe ages.

isotherms. Using FetKin, the shortening rates in the


initial kinematic restoration can be adjusted until the
model matches as many observations/variables as
possible. One of the powers of a kinematic approach
lies in the restoring properties of the acting deformation operators. Complicated sample paths can be
accurately traced, which is a key tool when determining the time in the past when samples reset as
compared to their computed forward ages. The
approach created by FetKin thus combines the most
1570

Coupling Kinematic Restorations and Temperature

We have introduced a software package that is able to


calculate isotherm advection in complex geological
settings. The new software FetKin is especially suitable for areas with rapid shortening rates (past and
recent), jagged topography, and can be used considering numerous and complex kinematic paths for a
growing structure. To do that, FetKin couples kinematic restorations and a 2-D temperature model to
model temperatures, from which it calculates cooling
ages for various thermochronometers. It includes the
possibility of testing realistic thermal parameters,
changing relief, and episodic versus continuous tectonic events. These factors have observable effects
on the age response. A careful modeling of each of
these features provides a better basis for interpreting
and understanding thermochronological data. The
deviations observed in age response for changing
materials warn us of the nonuniqueness of this complex problem and the uncertainty related to not
always having access to physical properties. One of
the main implications of a careful 2-D modeling is
that we may estimate exhumation rates, denudation
rates, and velocity of deformation by testing different
scenarios in actual structural cross sections. The procedure is shown in more detail in a companion paper
by Mora et al. (2015, this issue). In those cases, the
procedure includes finding a modeled solution that
fits more observations by trial and error, departing
from an initial simple kinematic restoration which
assumes planar isotherms.
Forward modeling as done by FetKin allows
for a more efficient and significant sampling of data.
We may identify regions that do not contain a topographical effect or that do not have significant local
influence of materials. FetKin is designed to work
closely with the kinematic algorithms of 2DMove,
providing a general framework for working with real
geological cross sections. However, the generalized
design and approach in FetKin should allow it to be
adapted to other 2-D cross section restoration programs, if desired.

REFERENCES CITED
Allen, P., and J. Allen, 2005, Basin analysis: Principles and applications: Oxford, Blackwell Science, 550 p.
Anadn, P., L. Cabrera, F. Colombo, M. Marzo, and O. Riba,
1986, Syntectonic intraformational unconformities in alluvial fan deposits, eastern Ebro basin margins (NE Spain),
in P. A. Allen and P. Homewood, eds., Foreland basins:
Oxford, Blackwell Publishing, Special Publication
of the International Association of Sedimentologists, v. 8,
p. 259271.
Baby, P., P. Rochat, G. Mascle, and G. Hrail, 1997, Neogene
shortening contribution to crustal thickening in the back arc
of the Central Andes: Geology, v. 25, p. 883886, doi:10
.1130/0091-7613(1997)025<0883:NSCTCT>2.3.CO;2.
Batt, G. E., and J. Braun, 1999, The tectonic evolution of the
Southern Alps, New Zealand: Insights from fully thermally
coupled dynamical modelling: Geophysical Journal
International, v. 136, no. 2, p. 403420, doi:10.1046
/j.1365-246X.1999.00730.x.
Batt, G. E., and M. T. Brandon, 2002, Lateral thinking: 2-D
interpretation of thermochronology in convergent orogenic
settings: Tectonophysics, v. 349, p. 185201, doi:10.1016
/S0040-1951(02)00053-7.
Bollinger, L., J. P. Avouac, O. Beyssac, E. J. Catlos, T. M.
Harrison, M. Grove, B. Goffe, and S. Sapkota, 2004,
Thermal structure and exhumation history of the Lesser
Himalaya in central Nepal: Tectonics, v. 23, no. 5, p. 119,
doi:10.1029/2003TC001564.
Brandon, M. T., M. Roden-Tice, and J. I. Garver, 1998, Late
Cenozoic exhumation of the Cascadia accretionary wedge
in the Olympic Mountains, northwest Washington State:
The Geological Society of America Bulletin, v. 110, no. 8,
p. 9851009, doi:10.1130/0016-7606(1998)110<0985:
LCEOTC>2.3.CO;2.
Braun, J., 2003, Pecube: A new finite-element code to solve the
3D heat transport equation including the effects of a timevarying, finite amplitude surface topography: Computers
and Geosciences, v. 29, no. 6, p. 787794, doi:10.1016
/S0098-3004(03)00052-9.
Braun, J., P. van der Beek, and G. Batt, 2006, Quantitative
thermochronology: Numerical methods for the interpretation
of thermochronological data: New York, Cambridge
University Press, 272 p.
Carslaw, H. S., and J. C. Jaeger, 1986, Conduction of heat in
solids: London, Oxford University Press, 520 p.
Demongodin, L., B. Pinoteau, G. Vasseur, and R. Gable, 1991,
Thermal conductivity and well logs: A case study in the
Paris basin: Geophysical Journal International,
v. 105, no. 3, p. 675691, doi:10.1111/j.1365-246X.1991
.tb00805.x.
Dodson, M. H., 1973, Closure temperature in cooling geochronological and petrological systems: Contributions to
Mineralogy and Petrology, v. 40, no. 3, p. 259274, doi:10
.1007/BF00373790.
Ehlers, T. A., 2005, Crustal thermal processes and the interpretation of thermochronometer data: Reviews in Mineralogy and
Geochemistry, v. 58, no. 1, p. 315350, doi:10.2138/rmg
.2005.58.12.

Ehlers, T. A., and D. S. Chapman, 1999, Normal fault thermal


regimes: Conductive and hydrothermal heat transfer surrounding the Wasatch Fault, Utah: Tectonophysics, v. 312,
no. 2, p. 217234, doi:10.1016/S0040-1951(99)00203-6.
Ehlers, T. A., S. D. Willett, P. A. Armstrong, and D. S.
Chapman, 2003, Exhumation of the Central Wasatch
Mountains 2: Thermo-kinematics of exhumation, erosion,
and thermochronometer interpretation: Journal of
Geophysical Research, v. 108, no. B3, 2173, doi:10.1029
/2001JB001723.
Ehlers, T. A., T. Chaudhri, S. Kumar, C. W. Fuller, S. D. Willett,
R. A. Ketcham, M. T. Brandon, D. X. Belton, B. P. Kohn,
A. J. W. Gleadow, T. J. Dunai, and F. Q. Fu, 2005,
Computational tools for low-temperature thermochronometer interpretation, in P. W. Reiners and T. A. Ehlers, eds.,
Low-temperature thermochronology: Techniques, interpretation and applications: Reviews in Mineralogy and
Geochemistry, v. 58, p. 589622.
Eichelber, N., N. McQuarrie, T. A. Ehlers, E. Enkelmann, J. B.
Barnes, and R. O. Lease, 2013, New constraints on the chronology, magnitude and distribution of deformation within
the central Andean orocline: Tectonics, v. 32, p. 122,
doi:10.1002/tect.20073.
Espurt, N., S. Brusset, P. Baby, W. Hermoza, R. Bolaos, D.
Uyen, and J. Dramond, 2008, Paleozoic structural controls
on shortening transfer in the Subandean foreland thrust system, Ene and southern Ucayali basins, Peru: Tectonics,
v. 27, TC3009, doi:10.1029/2007TC002238.
Farley, K. A., 2000, Helium diffusion from apatite: General
behavior as illustrated by Durango fluorapatite: Journal of
Geophysical Research, v. 105, no. B2, p. 29032914,
doi:10.1029/1999JB900348.
Gans, P. B., E. L. Miller, R. Brown, G. Housman, and G. S.
Lister, 1991, Assessing the amount, rate, and timing of tilting in normal fault blocks: A case study of tilted granites in
the Kern-Deep Creek Mountains, Utah: Geological Society
of America Cordilleran Section, 87th Annual Meeting
Abstracts with Programs, v. 23, no. 2, 28 p.
Hames, W. E., and S. A. Bowring, 1994, An empirical evaluation
of the argon diffusion geometry in muscovite: Earth and
Planetary Science Letters, v. 124, p. 161169, doi:10.1016
/0012-821X(94)00079-4.
Herman, F., P. Copeland, J. P. Avouac, L. Bollinger, G. Maho,
P. Le Fort, S. Rai, D. Foster, A. Pcher, K. Stwe, and P.
Henry, 2010, Exhumation, crustal deformation, and thermal
structure of the Nepal Himalaya derived from the inversion
of thermochronological and thermobarometric data and
modeling of the topography: Journal of Geophysical
Research, v. 115, B06407, doi:10.1029/2008JB006126.
Huerta, A. D., and D. W. Rodgers, 2006, Constraining rates of
thrusting and erosion: Insights from kinematic thermal
modeling: Geology, v. 34, no. 7, p. 541544, doi:10.1130
/G22421.1.
Ketcham, R., 1996, Thermal models of core complex evolution in
Arizona and New Guinea: Implications for ancient cooling
paths and present day heat flow: Tectonics, v. 15, no. 5,
p. 933951, doi:10.1029/96TC00033.
Ketcham, R., 2005, Forward and inverse modeling of
low-temperature thermochronometry data: Reviews in

ALMENDRAL ET AL.

1571

Mineralogy and Geochemistry, v. 58, no. 1, p. 275314,


doi:10.2138/rmg.2005.58.11.
Ketcham, R. A., A. Carter, R. A. Donelick, J. Barbarand, and
A. J. Hurford, 2007, Improved modeling of fission-track
annealing in apatite: American Mineralogist, v. 92, no. 5,
p. 799810, doi:10.2138/am.2007.2281.
Lock, J., and S. Willet, 2008, Low-temperature thermochronometric ages in fold-and-thrust belts: Tectonophysics,
v. 456, no. 3, p. 147162, doi:10.1016/j.tecto.2008.03.007.
Lynch, D. R., 2004, Numerical partial differential equations for
environmental scientists and engineers: A first practical
course: New York, Springer, 388 p.
Mancktelow, N. S., and B. Grasemann, 1997, Time-dependent
effects of heat advection and topography on cooling histories during erosion: Tectonophysics, v. 270, p. 167195,
doi:10.1016/S0040-1951(96)00279-X.
Martnez, J., 2006, Structural evolution of the Llanos foothills,
Eastern Cordillera, Colombia: Journal of South American
Earth Sciences, v. 21, no. 4, p. 510520, doi:10.1016
/j.jsames.2006.07.010.
Medwedeff, D. A., 1989, Growth fault-bend folding at Southeast
Lost Hills, San Joaquin Valley, California: AAPG Bulletin,
v. 73, no. 1, p. 5467.
Mora, A., W. Casallas, R. A. Ketcham, D. Gomez, M. Parra,
J. Namson, D. Stockli, A. Almendral, W. Robles, and
B. Ghorbal, 2015, Kinematic restoration of contractional
basement structures using thermokinematic models:
A key tool for petroleum system modelling: AAPG
Bulletin, v. 99, no. 8, p. 15751598, doi:10.1306
/04281411108.
Mora, A., M. Parra, M. R. Strecker, A. Kammer, C. Dimat, and
F. Rodrguez, 2006, Cenozoic contractional reactivation
of Mesozoic extensional structures in the Eastern Cordillera
of Colombia: Tectonics, v. 25, TC2010, doi:10.1029
/2005TC001854.
Mora, A., M. Parra, M. R. Strecker, E. R. Sobel, H.
Hooghiemstra, V. Torres, and J. Vallejo Jaramillo, 2008,
Climatic forcing of asymmetric orogenic evolution in the
Eastern Cordillera of Colombia: Geological Society of
America Bulletin, v. 120, p. 930949, doi:10.1130/B26186.1.
Mora, A., M. Parra, M. R. Strecker, E. R. Sobel, G. Zeilinger, C.
Jaramillo, S. Ferreira da Silva, and M. Blanco, 2010, The
eastern foothills of the Eastern Cordillera of Colombia: An
example of multiple factors controlling structural styles and
active tectonics: Geological Society of America Bulletin,
v. 122, p. 18461864, doi:10.1130/B30033.1.
Nemcok, M., S. Schamel, and R. Gayer, 2005, Thrustbelts:
Structural architecture, thermal regimes, and petroleum systems: Cambridge, Cambridge University Press, 554 p.
Novoa, E., J. Suppe, and J. H. Shaw, 2000, Inclined-shear
restoration of growth folds: AAPG Bulletin, v. 84, no. 6,
p. 787804.
Price, R. A., 1981, The Cordilleran foreland thrust and fold belt in
the southern Canadian Rockies, in K. R. McClay and N. J.
Price, eds., Thrust and nappe tectonics: London, Geological
Society Special Publication, v. 9, p. 427448.
Ring, U., M. T. Brandon, G. S. Lister, and S. D. Willett, 1999,
Exhumation processes, in U. Ring, M. T. Brandon, G. S.
Lister, and S. D. Willet, eds., Exhumation processes:

1572

Coupling Kinematic Restorations and Temperature

Normal faulting, ductile flow and erosion: London,


Geological Society Special Publication, p. 127.
Snchez, N., A. Mora, M. Parra, D. Garcia, M. Cortes, T. M.
Shanahan, R. Ramirez, O. Llamosa, and M. Guzman, 2015,
Petroleum system modeling in the Eastern Cordillera using
geochemistry and timing of thrusting and deformation:
AAPG Bulletin, v. 99, no. 8, p. 15371556, doi:10.1306
/04161511107.
Shaw, J. H., E. Novoa, and C. D. Connors, 2004, Structural controls on growth stratigraphy in contractional fault-related
folds, in K. R. McClay, ed., Thrust tectonics and hydrocarbon systems: AAPG Memoir 82, p. 400412.
Sobel, E. R., and M. R. Strecker, 2003, Uplift, exhumation and
precipitation: Tectonic and climatic control of late
Cenozoic landscape evolution in the northern Sierras
Pampeanas, Argentina: Basin Research, v. 15, no. 4,
p. 431451, doi:10.1046/j.1365-2117.2003.00214.x.
Stockli, D. F., 2005, Applications of low-temperature thermochronometry to extensional tectonic settings, in P. W.
Reiners and T. A. Ehlers, eds., Low-temperature thermochronology: Techniques, interpretation and applications:
Reviews in Mineralogy and Geochemistry, v. 58,
p. 411448.
Strayer, L. M., S. G. Erickson, and J. Suppe, 2004, Influence of
growth strata on the evolution of fault-related folds:
Distinct-element models, in K. R. McClay, ed., Thrust
tectonics and hydrocarbon systems: AAPG Memoir 82,
p. 413437.
Stwe, K., L. White, and R. Brown, 1994, The influence of eroding topography on steady-state isotherms: Application to fission track analysis: Earth and Planetary Science Letters,
v. 124, p. 6374, doi:10.1016/0012-821X(94)00068-9.
Suppe, J., 1980a, Imbricated structure of western foothills belt,
South Central Taiwan: Petroleum Geology of Taiwan,
v. 17, p. 116.
Suppe, J., 1980b, A retrodeformable cross section of northern
Taiwan: Proceedings of the Geological Society of China,
v. 23, p. 4655.
Suppe, J., 1983, Geometry and kinematics of fault-bend folding:
American Journal of Science, v. 283, no. 7, p. 684721,
doi:10.2475/ajs.283.7.684.
Suppe, J., G. T. Chou, and S. Hook, 1991, Rates of folding and
faulting determined from growth strata, in K. R. McClay,
ed., Thrust tectonics: London, Chapman and Hall, p. 105121.
Suppe, J., F. Sabat, J. A. Muoz, J. Poblet, E. Roca, and J.
Vergs, 1997, Bed-by-bed fold growth by kink-band migration: Sant Lloren de Morunys, eastern Pyrenees: Journal
of Structural Geology, v. 19, p. 443461, doi:10.1016
/S0191-8141(96)00103-4.
Vergs, J., D. W. Burbank, and A. Meigs, 1996, Unfolding: An
inverse approach to fold kinematics: Geology, v. 24, no. 2,
p. 175178, doi:10.1130/0091-7613(1996)024<0175:
UAIATF>2.3.CO;2.
Whipple, K. X., 2004, Bedrock rivers and the geomorphology of
active orogens: Annual Review of Earth and Planetary
Sciences, v. 32, p. 151185, doi:10.1146/annurev.earth.32
.101802.120356.
Whipple, K. X., and G. E. Tucker, 1999, Dynamics of the streampower river incision model: Implications for height limits of

mountain ranges, landscape response timescales, and


research needs: Journal of Geophysical Research, v. 104,
no. B8, p. 17,66117,674, doi:10.1029/1999JB900120.
Willett, S. D., 1999, Orogeny and orography: The effects of
erosion on the structure of mountain belts: Journal of
Geophysical Research, v. 104, no. B12, p. 957981, doi:10
.1029/1999JB900248.

Wolf, R. A., K. A. Farley, and D. M. Kass, 1998, Modeling


of the temperature sensitivity of the apatite (UTh)/He thermochronometer: Chemical Geology, v. 148, p. 105114,
doi:10.1016/S0009-2541(98)00024-2.
Zienkiewicz, O. C., R. L. Taylor, and P. Nithiarasu, 2005, The
finite element method for fluid dynamics: Oxford, Elsevier
Butterworth-Heinemann, 400 p.

ALMENDRAL ET AL.

1573

You might also like