RapleeLong080108 - Stanford School of Earth, Energy

advertisement
1
2
3
4
5
Structural geometry of Raplee Ridge, UT Revealed Using Airborne Laser Swath
6
Mapping (ALSM) Data
7
8
Hilley, G. E. (*), Mynatt, I., and Pollard, D. D.
9
Department of Geological and Environmental Sciences, Stanford University, Stanford,
10
CA 94305-2115
11
12
(*) Corresponding Author: hilley@stanford.edu
13
14
15
16
17
18
19
20
In preparation for Journal of Structural Geology
21
1
22
Abstract:
23
Kinematic models are often used relate the geometry of strata deformed in fault-
24
related folds to the location and geometry of faults in the subsurface. Recently, two-
25
dimensional mechanical analyses that assume a viscous rheology have demonstrated that
26
it is possible to use such an approach to infer loading conditions and fault geometries
27
based on observations of surface deformation. In this contribution, we present a method
28
that uses a three-dimensional boundary element model to infer the geometry of an
29
underlying fault and the loading conditions that are most consistent with layers deflected
30
by fault-related folding. The model assumes that deformation accrues in a homogeneous,
31
isotropic, linear-elastic half-space as a frictionless elliptical fault slips in response to
32
remote loading.
33
We apply this model to the Raplee Monocline in southwest Utah, where Airborne
34
Laser Swath Mapping (ALSM) topographic data detail the geometry of layers deflected
35
within this fold. The spatial extent of five stratal surfaces were mapped using the ALSM
36
data, elevations were extracted from the high-resolution topography, and points along
37
these layers were used to infer the underlying fault geometry and remote strain
38
conditions. First, we compared elevations extracted from the ALSM data to publicly
39
available Digital Elevation Models (DEMs), such as the National Elevation Dataset 10-m
40
DEM (NED-10) and 30-m DEM (NED-30). We found that while the spatial resolution of
41
the NED dataset was far coarser than the ALSM data, the elevation values extracted at
42
points spaced ~50 m apart from each mapped surface yielded similar elevations. Second,
43
we used a single layer in the Raplee Monocline to infer the geometry of an underlying
44
fault and the remote loading that is most consistent with the deformation recorded by this
45
stratum. Using a Bayesian sampling method, we also assessed the uncertainties within,
46
and covariation between the fault geometric and loading parameters of the model. Next,
47
we considered elevations extracted along all five mapped layers when inferring fault
48
geometry, and found broad agreement with results obtained from considering a single
49
layer in isolation. Interestingly, modeled elevations matched those observed to within a
50
root-mean-squared error of 16-18 m, and showed little bias with position along the fold.
51
Thus, while our idealized isotropic elastic model is clearly oversimplified when
52
considering the complicated layered rheology of this fold, it provides an excellent fit to
2
53
the observations. This may suggest that comparable surface deflections can be produced
54
with a variety of rheological models; constraining factors such as the fault geometry
55
when performing these types of inversions may be required to ascertain the appropriate
56
representation of fold rheology.
57
3
58
Introduction:
59
Slip along blind reverse faults deform the surrounding rock mass, often resulting
60
in large deflections of the strata exposed at the surface and present in the subsurface
61
{e.g.`, \Allmendinger, 1998 #347; Narr, 1994 #346}. Forward mechanical models have
62
been used to understand how fault geometry and loading changes the deformation
63
surrounding these faults {e.g.`, \Allmendinger, 1998 #347}. In general, those studies
64
relating surface deformation to slip along faults in the subsurface either assume a fault’s
65
geometry and solve for the slip distribution acting along it {Johnson, 2002 #348}, assume
66
that slip along a fault is uniform and solve for the location and geometry of the fault {,
67
2007 #558}, or assume that long-term deformation recorded within folded strata conserve
68
their cross-sectional area and use the surface exposure of such strata to infer the location
69
and geometry of subsurface faults {Wolkowinsky, 2004 #560}. The latter approach has
70
been widely used to extrapolate deformed structural markers at the surface to fault
71
geometry at depth; however, such methods produce non-unique inferences of the
72
geometry of underlying structures {Mynatt, 2007 #558; Ziony, 1966 #559} and may
73
result in physically unrealistic particle displacements as the strata are required to deform
74
discontinuously {e.g.`, \Stone, 1993 #569; Stone, 2004 #570}.
75
approaches are unable to infer both the geometry of underlying faults as well as the
76
remote loading conditions that cause them to slip. For this reason, Johnson and Johnson
77
(2002) derived a complete mechanical description of the folding process in two
78
dimensions that uniquely relates surface deformation to fault geometry when the strata
79
deform according to a viscous rheology. This approach is advantageous in that the
80
observed and inferred structural geometries are uniquely related through the requirement
81
of continuity in the surrounding medium and realistic physical rheologies may be
82
specified.
83
dimensional elastic model to infer subsurface geometry and remote loading conditions
84
from surface deflections recorded in the geologic record.
In addition, these
More recently, Mynatt et al. {Thomas, 1993 #350} have used a three
85
The well-exposed Raplee Ridge Monocline in southeastern Utah is a ~N-S
86
oriented fold that is ~14-km-long by ~2-km-wide (Fig. 1). Here, the San Juan River has
87
incised through the fold in the last several Ma {Mynatt, 2007 #558}, exposing a thick
88
sedimentary package {Davis, 1999 #562`, and references therein}. Folding within the
4
89
ridge likely occurred during the Laramide phase of deformation in the area during latest
90
Mesozoic and early Cenozoic time. Many folds within the Colorado Plateau were formed
91
due to reactivation of high-angle structures that likely date back as far as the Precambrian
92
{Davis, 1999 #562`, and references therein}. While no fault is exposed at the Raplee
93
Ridge Monocline and no subsurface information exists here, the presence of steeply-
94
dipping beds along its west side along with the fact that this fold is analogous to many
95
other Laramide folds in the Colorado Plateau for which subsurface structures have been
96
imaged {Davis, 1999 #562`, and references therein} suggest that this fold has likewise
97
been formed above a east-dipping high-angle reverse fault.
98
The National Center for Airborne Laser Mapping (NCALM) collected high-
99
resolution topographic data that images the deformed strata in this structure (Fig. 2).
100
These data provide a high precision, dense array of points that define the fold’s geometry,
101
which we use to infer information about the subsurface geometry of faults that may be
102
responsible for the formation of the Raplee Ridge Monocline. Our previous work at
103
Raplee Ridge Monocline combined the high-resolution ALSM data with an elastic
104
boundary element model {Mynatt, 2007 #558} to infer the geometry of a fault that may
105
have produced the fold {e.g.`, \Gratier, 1991 #571}.
106
In this contribution, we compare the ALSM data to other more commonly
107
available topographic data to determine the necessary spatial resolution that is required to
108
infer the subsurface geometry of faults based on the deflections of exposed folded strata.
109
In addition, we document a method to infer fault geometry from exposed surface strata,
110
and expand this method to use a series of layers within a deformed fold to better constrain
111
the geometry of the underlying structure. We use a Markov-Chain Monte Carlo method
112
to provide an estimate of the variation within, and covariation between, model estimates
113
of fault geometry parameters that produce deformation similar to that recorded within the
114
fold. Given the assumptions embedded in our mechanical model, our approach places
115
broad constraints on the geometry of the underlying fault that may be improved using our
116
approach should additional prior information (such as a range in fault geometric
117
parameters) be available. Even using the assumption of a homogeneous, isotropic linear-
118
elastic rheology for the deforming strata, we find good agreement between observed and
119
predicted deflections for our best-fit model. We speculate that our inversion method,
5
120
used with physical models that consider more complicated rheologies, may produce
121
deflections that produce a similar level of agreement than those observed. If correct, this
122
indicates that when viewed in isolation from other information, such as direct measures
123
of physical rock properties and prior knowledge of factors such as fault geometry, surface
124
deflections may provide ambiguous information about the rock rheology at the time of
125
deformation.
126
127
Study Area:
128
Raplee Ridge Monocline is located in southeast Utah outside of the town of
129
Mexican Hat (Fig. 1). The details of the field site are reported in Mynatt et al. (in
130
review); herein we summarize the most important aspects of the field site. The fold is
131
composed of the Pennsylvanian to Permian Rico Formation, which consists of alternating
132
limestones, siltstones, and sandstones that record a general trend of marine regression up-
133
section. The occurrence of marine incursions during this generally regressional sequence
134
has resulted in the deposition of five limey sandstones and sandy limestone layers
135
encased within shales and siltstones that currently provide well-exposed stratigraphic
136
markers whose surfaces reveal the geometry of the fold (Figs. 1, 2d, and 8). These layers
137
are folded into an ~500-m high doubly plunging monocline whose fold axis trends
138
approximately N-S. Sediments within the monocline’s forelimb dip steeply (up to 40°) to
139
the west, while strata are gently dipping (< 5°) to the east within its backlimb (Figs. 1 and
140
2). Exposure of different statigraphic levels varies depending on the depth of incision
141
into the fold. The surface of the highest stratigraphic level used as part of this study, the
142
McKim limestone, is widely exposed throughout the fold due to the fact that erosion has
143
stripped virtually all of the overlying sediments from this surface over much of the fold
144
(Fig. 2). Within basins incised into the fold, four additional mapable layers define the
145
fold’s geometry including, from top to bottom, the Goodrich, Shafer, Mendenhall, and
146
Unnamed limestones (Mynatt et al., in review). Within these basins, individual bedding
147
surfaces can be identified and mapped; however, the exposure of these four lower units is
148
often restricted to limited areas of the fold hinge that have been excavated by erosion.
149
Overlying the McKim limestone, exposure of the Halgaito Tounge shale along the
6
150
peripheral edges and away from the fold constrains the extent of the fold by defining
151
those areas that are undeformed.
152
We have no direct constraints on the geometry of the underlying fault that formed
153
Raplee Ridge. Throughout the Colorado Plateau, similar monoclines are often associated
154
with Laramide contraction that occurred during late Mesozoic through early Cenozoic
155
time {Johnson, 2002 #348}. Where exposed, at least some of these structures appear to
156
result from movement along high-angle basement faults that may have first experienced
157
motion as early as the Proterozoic {e.g.`, \Willemse, 1996 #344`, and references therein}.
158
In some cases, the reverse faulting that deformed Colorado Plateau strata into monoclinal
159
geometries were subsequently reactivated during Basin and Range extension in the
160
Miocene {e.g.`, \Huntoon, 1975 #563}. However, where this can be shown to be the
161
case, normal-sense reactivation of the basement faults caused them to propagate through
162
the strata above. Thus, Laramide contractional deformation is often accommodated
163
within the Paleozoic and Mesozoic section by warping of the strata, whereas later
164
extensional deformation typically resulted in discrete offset of these units {Ziony, 1966
165
#559}. Within the Raplee Ridge Monocline, detailed field observations show that there
166
are no faults that link with the underlying basement structures; thus, we infer that the
167
deflected strata within the Raplee Ridge Monocline result from Laramide contraction,
168
rather than Miocene extension. As a result, the western monoclinal forelimb implies that
169
a reverse fault that formed the structure likely dips steeply towards the east.
170
171
Airborne Laser Swath Mapping (ALSM) data:
172
On February 24, 2005, the National Center for Airborne Laser Mapping
173
(NCALM) collected topographic data from the Raplee Ridge area using an Optech 25
174
kHz pulsed laser range finding system and associated Inertial Navigation Unit (INU) that
175
is corrected for drift using kinematic GPS observations taken onboard the aircraft. The
176
acquisition was performed to build a fold-scale geometric model to compare outcrop-
177
scale measurements of fracture characteristics to fold geometry (Mynatt et al., in prep.).
178
The combination of acquisition frequency and low elevation flight plan provided several
179
laser range positions per square meter from which a 1 m Digital Elevation Model (DEM)
180
was produced by kriging interpolation. The vegetation at the site proved sufficiently
7
181
sparse to obviate the need for spatial filtering to remove its contribution from the DEM
182
(Fig. 2). From this 1-m DEM, a shaded relief map of the fold was produced, which was
183
used in the field to map the extent of the uppermost bedding plane of six particular
184
exposed stratagraphic units throughout the fold (Fig. 2d). Once these surfaces were
185
identified and mapped on the georeferenced images, (x,y,z) points that defined the
186
geometry of each of the surfaces were extracted from the high-resolution DEM. In some
187
cases, the topographic surfaces that appear to define the tops of the mapped strata have
188
been modified by up to 1 m by surface processes; in addition, erosion along the edges of
189
the surfaces has distributed a small amount of colluvium (typically less than 1 m thick)
190
onto the tops of the exposed strata.
191
estimates to be ~ 2-5 m {Bayes, 1763 #215}.
Thus, we estimate the precision of elevation
192
The extraction of points from the mapped layers produced > 100,000 points.
193
Consideration of this large number of points would render intractable the fold-scale
194
inversion described below. For this reason, we decimated the high-resolution dataset to
195
provide (x,y,z) points for each of the surfaces that were spaced no less than 50 m apart.
196
We compared the ALSM measurements of these extracted points to more commonly
197
available datasets to assess the precision of these other data for use in our study (Figures
198
3 and 4). Specifically, we compared equivalent z values at the selected (x,y) points to
199
those extracted from the National Elevation Dataset (NED) 30-m resolution DEM and
200
10-m resolution DEM. The poor spatial resolution of both NED datasets relative to the
201
ALSM data would prohibit accurate identification and mapping of the extent of the
202
different stratigraphic levels within the fold (Fig. 3), although low-elevation air photos
203
that were precisely georeferenced might be employed for such a purpose. However,
204
considering the coarse resolution of both NED datasets, their extracted elevation values
205
compared favorably with those obtained from the decimated ALSM dataset (Fig. 4a, c).
206
To provide a quantitative metric of the differences between these three datasets, we
207
calculated the residual elevation by subtracting the elevation value at each (x,y) point
208
used in this study for both NED datasets from the elevation at corresponding (x,y)
209
locations in the ALSM dataset (Fig. 4b, d). Elevations from both NED datasets showed
210
little, if any bias in elevation relative to the ALSM data (mean residual value = 1.6 m and
211
–1.1 m for the NED 30-m and 10-m datasets, respectively, relative to the ALSM dataset).
8
212
Variation between these two datasets was modest (standard deviation in residuals from
213
NED 30-m and 10-m datasets was 4.1 and 4.2 m, respectively); however, as described
214
below, the average misfit for our best-fit inversions was several times this variation.
215
Thus, given the uncertainties in the fold-wide inferences produced by modeling carried
216
out in this study, elevations extracted using the NED 30-m and NED 10-m datasets
217
should supply adequate precision when using these data as inputs to the fault geometry
218
inversions described below, provided accurate locations of the stratigraphic layers can be
219
defined.
220
221
Methods:
222
223
Inversion Methods:
224
The two major goals of this study were to provide a detailed and accurate
225
depiction of the continuous structural geometry of Raplee Ridge Monocline that could be
226
compared to outcrop scale measurements of fracture characteristics, and to infer the
227
geometry of and loading conditions acting along the reverse fault underlying the
228
monocline. The finite extent of the Raplee Ridge Monocline necessitated the use of a
229
three-dimensional model that could capture the along-strike variations in fault and fold
230
geometry (e.g., Fig. 1 and 2). Some three-dimensional models that conserve the area and
231
length of strata exist {e.g.`, \Hilley, 2008 #223; Hilley, 2008 #561}; however, as outlined
232
above, such methods do not take advantage of the constraints provided by a full
233
mechanical analysis. Likewise, methods that use a complete mechanical analysis to
234
relate stratal deflections to the geometry of underlying strata assume two-dimensional
235
plane-strain conditions {e.g.`, \Pollitz, 2003 #567}, and thus cannot adequately capture
236
the three-dimensional nature of the Raplee Ridge Monocline.
237
In this study, we used the three-dimensional Boundary Element Model (BEM)
238
Poly3D to relate the deflection of stratal layers to underlying fault geometry and loading
239
conditions (Thomas, 1993). Poly3D idealizes the rheology of Earth’s brittle upper crust
240
as homogeneous, isotropic, and linear-elastic. Arbitrarily shaped planar elements are
241
embedded into this elastic material to represent features such as faults and fractures along
242
which displacement within the material is discontinuous. Slip or opening along these
9
243
elements is driven either by specifying the displacement discontinuity directly, allowing
244
the elements to slip under a prescribed remote loading, specifying a stress drop along the
245
elements, or any combination of these boundary conditions for the normal and two shear
246
components. For example, one may specify mixed-mode boundary conditions in which
247
shear tractions acting across an element induce slip parallel to the element, while
248
prescribing a zero displacement condition perpendicular to the element (no
249
opening/closing). A set of planar elements may be assembled to produce an arbitrarily
250
complex surface in three-dimensions along which any combination of displacement
251
discontinuity or stress drop boundary conditions may be specified. Finally, Poly3D can
252
calculate fault slip distributions along the fault(s), stresses/strains within the surrounding
253
medium, and displacements within a half or full-space problem. The former conditions
254
compute stresses, strains, and displacements given the presence of a traction-free surface
255
that represents the interface between the solid earth and atmosphere, while the latter may
256
be used to approximate faulting processes deep within the crust.
257
In this study, we assume that stratal deflections are produced by slip along a
258
single, elliptically shaped, frictionless planar fault underlying the monocline. Its elliptical
259
geometry approximates the three-dimensional extent of many faults documented in the
260
field {Metropolis, 1953 #182}, while ignoring geometric complexity that would be
261
difficult to constrain using measurements of stratal deflections from the ALSM data.
262
Likewise, we chose a planar fault geometry to mimic the geometry of older reactivated
263
basement faults {, 2007 #558}.
264
underlying this monocline is likely more complex than this simple geometry; however, as
265
we show below, even this simple idealization of the underlying fault geometry produces
266
deflections that closely mimic those observed at Raplee Ridge Monocline.
267
additional data become available that better constrain the at-depth geometry of the Raplee
268
Ridge fault, it is straightforward to assimilate this information into our modeling
269
approach. Given this idealization, the geometry and location of the fault underlying the
270
fold is defined by 9 parameters (Fig. 5): the down-dip length of the fault (H), the along-
271
strike width of the fault (W), the depth of the fault below the surface (D), the map-view
272
location of the fault relative to the coordinate system of the model (xo, yo), the dip of the
273
fault (), the strike of the fault (s), and the remote conditions that drive motion along the
We acknowledge that the geometry of the fault
10
Should
274
fault. In this study, slip along the fault results from a prescribed remote strain tensor (xx,
275
yy, xy), which we assume does not vary with depth (zz=xz=xy=0). We use this remote
276
strain to compute the stresses acting along the fault elements, and compute shear
277
displacements that result from this load.
278
perpendicular to the fault surface, as large amounts of opening or fault-zone contraction
279
are physically unrealistic at depths of > 1.5 km that likely typify the shallowest levels of
280
slip along this fault during late Mesozoic time (see below).
However, we prohibit opening or closing
281
Once the fault geometry and loading conditions are specified, Poly3D calculates
282
stresses, strains, and displacements at specified points in the surrounding rock mass.
283
These displacements can be used to calculate the deflection of initially flat surfaces, such
284
as the strata currently exposed in the Raplee Ridge Monocline. We reconstructed the
285
relative elevation and stratigraphic thickness of each of the six units modeled in this study
286
using the measured stratigraphy {, 2007 #558} as well as ALSM-based thickness
287
measurements from flat-lying portions of the stratigraphy exposed by downcutting of the
288
San Juan River (Mynatt et al., in review). The likely depth of each of these units was
289
then inferred by noting that during late Mesozoic time, approximately 1.2 km of
290
sediments locally overlaid the section under study (IAN-INSERT REF). Thus, we were
291
able to use this information to estimate the depth of the originally flat-lying strata at the
292
time they were deformed.
293
As these units underwent deformation, points originally located on the flat-lying
294
surfaces were displaced both vertically and horizontally. Thus, the current location of the
295
observed points along deflected layers do not record their initial positions faithfully,
296
especially in the current case in which surface displacements are large.
297
calculates the displacement of a given point that is defined prior to deformation; however,
298
the displaced points measured using the ALSM dataset record each point’s final, rather
299
than initial position. We calculate the appropriate displacement at each (x, y) observation
300
point on the deformed-state surfaces using a two-step process. First, we calculate a
301
regular grid of displacements for each stratigraphic level that results from a specified set
302
of fault geometric and remote loading parameters. These regularly spaced points are
303
deformed into an irregularly shaped mesh that represents the final configuration of each
304
deformed surface. This process results in a set of irregularly spaced points on the
11
Poly3D
305
deformed surface for which we know displacements that are required to restore the points
306
to a regular grid. Using the deformed configuration, we use a linear interpolation to map
307
these displacement values to the (x, y) locations of each of the ALSM measurements, and
308
use these interpolated displacements to restore each point to its initial location on the
309
undeformed surface. In a second step, we use identical input parameters with Poly3D to
310
deform these initial-state points into the final state to determine the geometry of the
311
deflected surface. This results in (x, y) coordinates of the new deformed state points that
312
match the locations of the observations, and allows us to directly compare the elevation
313
values observed to those that are predicted by the BEM.
314
The above approach is appropriate for comparing observations of the elevation of
315
each deflected surface to those predicted by Poly3D. However, in some instances, such
316
as our observations along the Halgaito Tounge surface, the rotation of bedding rather than
317
its absolute elevation helps to constrain the fault geometry. In this case, we employ a
318
similar approach to that used for elevation values; however, instead of calculating
319
displacements at each of the points, we instead calculate bedding-plane rotation. As
320
before, we first restore the observation locations on each of the surfaces for which
321
bedding rotations are measured to their original locations. However, we use the initial-
322
state locations to calculate rotation of these points as they are deformed and compare
323
these rotations to observations. This was used to enforce areas of no rotation (e.g., flat-
324
lying sediments as seen in front of the fold).
325
Using these methods, we calculate the elevation values that would be predicted by
326
a given set of fault geometry and loading parameters, and compare them to those
327
observed elevation values for each layer using the following misfit function:
328
2 
 obs pred 2 
m  obs
rj  rjpred  
z

z
i
i

 W r  
WRSS  W z 

  zobs 
  robs
 
 
i
j
i1 
j1



n
329
(1)
330
331
332

where zobs are the ALSM-derived elevations of the deflected strata at the (x, y) locations,
zobs are the standard deviations of the elevation measurements, zpred are the predicted
333
elevations from Poly3D, Wz is the weight given to the misfit between measured and
334
observed elevation values, robs are the observed rotations (only flat-lying portions of the
12
335
Halgaito Tounge member are used in areas away from the fold to constrain its extent),
336
robs are the standard deviations of the rotation measurements, rpred are the predicted
337
rotations from Poly3D, Wr is the weight given to the misfit between measured and
338
observed rotations, and WRSS is the Weighted Residual Sum of Squares misfit function.
339
As WRSS values decrease, the displacements and rotations calculated by Poly3D better
340
match those observed. Thus, we can find the best-fitting fault geometric and loading
341
parameters (hereafter referred to cumulatively as “model parameters”) by exploring
342
various values of these model parameters, calculating the misfit defined by Eq. 1, and
343
identifying the model parameters associated with the minimum value of the WRSS.
344
In addition to the best-fitting model parameters, we quantify the uncertainty in
345
these model estimates using a Bayesian sampling method. In this approach, the model
346
parameters are viewed as a joint probability density function (pdf), the number of whose
347
dimensions corresponds to the number of model parameters. Such a joint pdf may be
348
used to determine both the best-fitting set of model parameters and uncertainties within
349
and covariation between model parameters. These measures can be used to quantify the
350
uniqueness of the best-fitting solution, and provide information about those combinations
351
of model parameters that provide similar fits to the observed data. We calculate this joint
352
pdf of the model parameters using Bayes’ Rule {, 2007 #558}:
353
P(m | x) 
354
P(x | m)P(m)
(2)
n
 P(x | m )P(m )
j
j
j1
355
356
357
where x is a vector consisting of the observations, m is a vector consisting of the model

parameters, P(m|x) is the joint pdf of the model parameters given the observations,
P(x|m) is the probability of the data given the model parameters that can be derived from
358
a misfit function similar to Eq. 1, P(m) is a pdf that represents the probability of
359
occurrence of the model parameters in the absence of any observations, and the
360
denominator normalizes the density of P(m|x) to unity {, 2007 #558}. P(m|x) is often
361
referred to as the posterior density while P(m) is referred to as the prior density.
362
Bayes’ Rule is used to estimate P(m|x) in our analysis as follows. First, a set of
363
model parameters (m) is selected and from these choices, predicted elevations and
13
364
rotations are computed at all points for which we have measured values of these
365
quantities. The probability of observing the data given the model parameters, P(x|m) is
366
computed using a modified version of Equation 1 that takes into account the number of
367
Degrees Of Freedom (DOF; defined as the number of observations minus the number of
368
model parameters) in the model and the uncertainty associated with each elevation or
369
rotation observation {e.g.`, \Jaeger, 1969 #568`, and references therein}:
P(x | m)  exp  r2 
370
371
(3a)
where
372 
 r2 
WRSS
DOF
(3b)
373
374
375

Bayes’ Rule also requires the definition of the prior density P(m). This prior
density represents the probability that a set of model parameters occurs in the absence of
376
any data to which the model is compared. This prior density may represent some
377
quantitative a priori information about various aspects of the fault geometry or loading
378
conditions, or simply may be used to incorporate expert opinion into the statistical
379
analysis. For example, at a particular site, subsurface seismic data may resolve a range of
380
permissible fault geometric parameters, such as fault strike and dip.
381
information constrains the geometry of the fault in the absence of any mechanical
382
modeling or measurements of bedding deflection. We cast these ranges in the fault
383
geometric parameters in terms of the probability of each value’s occurrence as a pdf. By
384
repeating this process for all of the model parameter values, we construct the prior pdf
385
P(m) that represents our prior knowledge of reasonable ranges in fault geometry. In the
386
case of the Raplee Ridge Monocline, we do not have additional prior information that
387
constrains the geometry of the underlying fault, and so we simply specify a uniform
388
probability of P(m) for all values of the model parameters, m. Thus, Eq. 2 reduces to
This prior
389
P(m | x) 
390
P(x | m)
(4)
n
 P(x | m )
j
j1

14
391
By evaluating P(x|m) for all permissible values of m, we can use Eq. 4 to compute the
392
probability of occurrence of the model parameters, P(m|x).
393
The simple mechanical model used to compute P(m|x) requires nine parameters to
394
be specified, and hence, m consists of a nine dimensional parameter space. If each
395
dimension of the parameter space were discretized into only ten values, the number of
396
evaluations of Poly3D that would be required to compute P(m|x) would be 109, or one
397
billion different combinations of model parameters. Such a large number of evaluations
398
is computationally infeasible given current computing technology. Indeed, it is this
399
limitation that has prevented the direct application of Bayes’ Rule to all but the simplest
400
problems with few model parameters. To circumvent this difficulty, Markov-Chain
401
Monte Carlo (MCMC) methods have been developed that sample the underlying
402
distribution P(m|x) in a computationally efficient way to provide a numerical
403
approximation of this distribution.
404
performed by the MCMC sampler, which is designed to sample P(m|x) according to the
405
pdf’s underlying probabilities.
406
identified using only a small fraction of the evaluations that would be required to
407
exhaustively explore the entire parameter space.
408
In these methods, sparse sampling of P(m|x) is
In this way, the portions with high P(m|x) can be
The sampler employed in this study is the Metropolis-Hastings MCMC method
409
{Mynatt, 2007 #558}.
In this algorithm, The Metropolis-Hastings sampler uses a
410
selection-rejection criterion to guide sampling through the parameter such that P(m|x) is
411
approximated. We first use an initial, randomly selected choice for m—the exact choice
412
for this starting point becomes less important as the simulation proceeds and the sampler
413
instead selects points based on the underlying distribution P(m|x). After this choice has
414
been made, a set of random numbers is drawn from the interval [-1, 1], one for each
415
dimension of m. These numbers are then scaled by a specified, arbitrary constant and
416
added to the previous choices for the initial model parameters, moving the samples a
417
random distance through the parameter space, m. This is the selection process. Next, this
418
new set of samples may be either accepted or rejected. If accepted, the new samples are
419
treated as a new starting point for the sampler and the selection process is repeated. If the
420
samples are rejected, the previous values of the model parameters are used to select a new
421
set of samples. Samples are accepted or rejected based on the probability of their
15
422
occurrence relative to the probability of occurrence of the previous set of samples. First,
423
the probability of occurrence for both the previous set of samples and the current
424
selection of samples is computed using Bayes’ Rule (Eq. 4). We define the probability of
425
the first sample as P1, whereas the probability of the second sample is denoted as P2. In
426
the case that P2 > P1, the new set of samples is always accepted. However, if P2 < P1, the
427
ratio of these two probabilities is computed and compared to a random number drawn
428
along the interval [0, 1]. If the ratio exceeds this random number the sample set is
429
accepted, otherwise it is rejected. Thus, the Metropolis-Hasting sampler explores the
430
parameter space m, is guided towards higher values of P(m|x) in Equation 4, and the
431
frequency of the model parameters chosen by the sampler approximates the posterior
432
density P(m|x). Initially, the values selected by the sampler will depend on the arbitrary
433
initial choices for the model parameters at the beginning of sampling; however, as
434
sampling proceeds, the memory of these initial choices tends to fade and the frequency of
435
m selected by the sampler will reflect P(m|x). Thus, we allow a “burn-in” period of
436
sampling to allow the memory of the arbitrary sampling starting location in the parameter
437
space to be erased, and we only consider the frequency of samples identified by sampler
438
after this burn-in when P(m|x) is computed. In our study, we allow a burn-in period of
439
50,000 samples to allow sample choices to become independent of the initial choices for
440
the model parameters, and collect 100,000 samples to approximate the posterior pdf,
441
P(m|x).
442
443
Modeling Raplee Ridge Monocline:
444
We used Poly3D to calculate bedding-plane deflections due to slip along a blind
445
reverse fault for two scenarios—one in which only the most well-exposed bedding
446
surface (McKim surface) was used to compare observed and modeled fold geometry, and
447
a second in which all mapped bedding surfaces were used to define the fold’s geometry
448
(Fig. 5). The first of these two scenarios is similar to the inversion presented in Mynatt et
449
al. {Figure 7; \Mynatt, 2007 #558}, excepting the fact that in this study, the misfit is
450
normalized by the number of degrees of freedom in the model. This difference does not
451
impact the values of the model parameters that best match the observed elevations;
16
452
however, the uncertainties associated with each model parameter tend to be larger (and
453
likely more realistic) than those presented in Mynatt et al. {Mynatt, 2007 #558}.
454
Elevation estimates along the mapped surfaces are available every square meter
455
using the high-resolution ALSM data. For expediency, we consider only a subset of
456
points separated by a minimum distance of 50 m when inverting for the fold’s geometry.
457
This reduces the number of computations necessary by a factor of ~2000.
458
Metropolis-Hastings inversions described above require several weeks to ~1.5 months of
459
compute time using this low-resolution dataset, and so inversions that use the full-
460
resolution data would be infeasible given today’s computing technology.
The
461
The mapped surfaces are not exposed within the forelimb syncline of the fold.
462
Thus, the observations of the mapped bedding surfaces permit a fold geometry in which
463
layers continue to dip steeply westward at great distances west of the extent of exposure
464
of these surfaces (Fig. 2b). However, surfaces exposed along unmapped higher bedding
465
surfaces within the stratigraphic section show sub-horizontal orientations, indicating that
466
the mapped units share a similar orientation in the subsurface. To account for these
467
observations, we used two different approaches.
468
considered only the observed geometry of the McKim surface, we assumed that the
469
stratigraphic thickness between the exposed units and the McKim bedding plane surface
470
within the fold’s eastern portion was similar to that observed in locations far from the
471
fold-related deformation. This thickness was used to infer the depth at which the McKim
472
surface should be located in the subsurface. These inferred (x,y,z) subsurface locations
473
were used when inverting for the fold’s geometry. In the case where we considered all
474
mapped layers in the inversion, we instead used a rotation constraint within the Halgaito
475
Tongue member, which is well exposed throughout the area surrounding the fold. In this
476
way, rather than assign subsurface locations for which the (x,y,z) locations of each of the
477
bedding planes is located, we instead required areas at the stratigraphic level of the
478
Halgaito Tongue member to experience no rotation (right-hand term in Eq. 1). This has a
479
similar effect of ensuring that the geometry of the fold is finite in areas to its west.
480
481
Results:
17
In the case of the models that
482
We first present the results of our fold inversion when only comparing elevation
483
values measured along the McKim surface to elevations of the deflected surface that were
484
predicted by Poly3D. The best-fit model (Fig. 6) shows that the observed fold geometry
485
is most consistent with a fault whose width is almost three times its down-dip length
486
(best-fitting model parameter values shown as italic numbers in each model parameter
487
panel of Fig. 7). In addition, the steep forelimb dips indicate that the tip-line of the fault
488
is likely only several hundred meters below the current surface in the fold’s center. The
489
strike of the underlying fault is approximately 3° east of north, consistent with the
490
roughly N-S trend of the observed topography. In addition, the inferred fault underneath
491
the fold is inferred to dip steeply (~47°) to the east. The Poisson’s ratio of the sediments
492
overlying the fold inferred by the model (0.28) is close to that typically assumed for
493
crustal rocks . Finally, the regional contraction in the E-W direction (xx) is expect to be
494
~3 times that in the N-S direction (yy), and is ~20 times larger than the regional shear
495
strain (xy). The absolute value of E-W strain (xx) for the best-fitting model was –0.43.
496
The best-fitting model produced surface elevations consistent with those
497
measured (Figure 6a). The quality of the fit is seen on a graph of observed versus
498
predicted values (Figure 6b) at points where elevations were extracted from the ALSM
499
data (locations of points shown in Figure 6a). The inferred subsurface points along the
500
western portion of the fold were assigned a constant value, which appear as a vertical line
501
of points at low elevations in Figure 6b. In contrast, calculated deflections vary smoothly
502
across these points, creating a mismatch between the inferred subsurface elevations and
503
those predicted by Poly3D. The residuals (defined as the modeled elevations minus the
504
observed elevations) show approximately zero mean.
505
systematically underestimated points at the crest of the fold, which skewed residuals
506
negatively (Fig. 6c). The best-fitting model yielded a root-mean-squared error of 16.1 m.
507
This value is significantly greater than the uncertainty in the elevation observations,
508
indicating that that mismatch between elevations predicted by the linear-elastic half-space
509
model and those observed is far larger than the uncertainties in elevation that we
510
measured using the ALSM. Indeed, given the high accuracy of the ALSM elevations (<
511
2-4 m), it is difficult to conceive of an idealized model (of any rheology) whose average
512
misfit might be on the order of, or less than this value. Thus, the high accuracy of our
18
In addition, the model
513
data relative to that expected from our idealized model forces us to conceptualize obs in
514
Eq. 1 as the inaccuracy that is produced by the strict model assumptions of Poly3D. In
515
this context, we adjust obs in Eq. 1 to require that r2 = 1 for the best-fitting set of model
516
parameters. As we see below, this choice for obs increases the variance of the model
517
parameters to a range that takes into account the uncertainty of using a linear elastic
518
model to calculate deflections in a material that might be better characterized by a more
519
complicated, anisotropic, and/or spatially variable rheology.
520
We used the Metropolis-Hastings MCMC method to calculate the posterior
521
probability, P(m|x), of each of the model parameters for which the inversion was
522
performed (Fig. 7). The joint posterior pdf is a nine-dimensional probability distribution,
523
with each of its axes representing the nine model parameters required for the calculation
524
of bedding-plane deflections. Such a distribution is difficult to visualize, and so we
525
present the marginal posterior pdfs by collapsing all dimensions of the joint pdf excepting
526
that of the model parameter of interest onto the dimension of this model parameter (Fig.
527
7). The resulting pdf shows the variability that characterizes each model parameter, but
528
does not capture the covariation of the different model parameters with one another. For
529
each of the model parameters, we report the value for the best-fitting scenario as italic
530
numbers, the mean of the simulated posterior pdfs as bold numbers, and the 95% range in
531
the simulated model parameter pdfs in parentheses in each of the panels. The simulations
532
reveal substantial variation in the model parameters such that a variety of their values
533
may produce similarly good fits to the observed elevations (Fig. 7). For many of the
534
model parameters, the mean values are similar to the best-fitting values, although
535
agreement is substantially less for those distributions that are highly skewed or uncertain
536
(such as , xx/xy and xx/yy). Nonetheless, the model places broad bounds on fault
537
geometries that may plausibly create the observed deflections. In some cases relatively
538
large values of Poisson’s ratio (0.23-0.49) are allowed, which implies that the single-
539
layer models favor a rheology that undergoes minimum volume change during
540
deformation. Considering the range of values produced by the simulation, the fault tip at
541
the center of the fold is expected to be only several km below the surface at the time of
542
deformation, and is currently only several hundred meters below the level of exposure.
19
543
Next, we performed our inversions using all five of the mapped layers that define
544
the geometry of the Raplee Ridge Monocline.
The best-fitting multi-layer model
545
elevations were similar to those observed (Fig. 8-10). The extensive exposure of the
546
McKim, Goodrich, and Shafer surfaces causes the model to most accurately characterize
547
its geometry (Fig. 9). Surfaces lower in the stratigraphy are not fit as well due to the fact
548
that the number of points extracted along these surfaces are fewer, and consequently their
549
weight is less in the inversion. This effect is most pronounced for the Mendenhall and
550
Unnamed surfaces, both of which show larger misfit than their counterparts higher in the
551
stratigraphic section (Figs. 9 and 10).
552
elevations for the multilayer models (Fig. 10), a slight but systematic underestimation of
553
elevations within the McKim and Goodrich surfaces are observed for elevations of ~1600
554
m. These points flank the upper-most crest of the fold along the McKim surface, and
555
their misfit arises because the fold is more cylindrical at its crest than might be expected
556
for the simple elliptical fault geometry assumed in this study. Nonetheless, the best-
557
fitting model geometry produces generally unbiased estimates of the elevations across all
558
surfaces (Fig. 10): model residuals for the McKim, Goodrich, Shafer, Mendenhall, and
559
Unnamed surface have means of -1.0, 1.2, -5.3, 1.0, and 4.6 m, respectively, with a mean
560
residual of 7.6 x 10-15 m when considering all points from all layers. The variation within
561
residuals was similar across the McKim, Goodrich, and Shafer surfaces, with standard
562
deviations equaling 16.9, 15.9 and 16.8 m, respectively. However, the fewer points
563
extracted along the less-well-exposed Mendenhall and Unnamed surfaces resulted in
564
higher standard deviations of 25.0 and 20.7 m, respectively.
When comparing observed versus predicted
565
The best-fitting model parameters deduced when using all of the mapped layers
566
(Fig. 8) shared affinity to those derived using only the McKim surface (Fig. 9 and 10).
567
The dimensions of the fold were similar, although the underlying fault was slightly wider
568
in the along-strike direction when using all mapped layers (Fig. 11, italic numbers in each
569
model parameter panel). In addition, the best-fitting model that uses all of the layers
570
favors a deeper fault tip-line at the center of the fold in comparison to the single layer
571
models. The root-mean-squared error for the best-fitting model was 18.1 m, slightly
572
larger than the single layer model.
20
573
The joint posterior distribution of model parameters estimated using the
574
Metropolis-Hasting sampler permitted sets of fault geometries and loading conditions that
575
were generally consistent with those derived from the single-layer models (Fig. 11).
576
However, inversions that utilized observations from multiple stratigraphic layers
577
generally tended to permit a wider range of model parameters than did single-layer
578
inversions (Figs. 7 and 11). Mean values of the calculated posterior pdfs were similar to
579
those calculated using the best-fitting model. As with the best-fitting modeled fault
580
depth, the range of fault depths allowable using the multilayer inversion tended to be
581
deeper than those predicted using only data from the McKim surface. Unlike single-layer
582
model results, values of Poisson’s Ratio permitted by the multilayer models spanned the
583
range from 0.03-0.48, indicating that this parameter may not be well resolved by such an
584
inversion. Nonetheless, both sets of models predict that the underlying fault dips steeply
585
to the east (Fig. 11).
586
587
Discussion:
588
This study builds on our previous work, in which we used the elevations extracted
589
from the McKim limestone in combination with Poly3D to interpolate the geometry of
590
the this surface and infer the fault geometry and regional strain . The single-layer models
591
created as part of the current study are similar to those reported in our previous work with
592
two important differences. First, our previous work fixed Poisson’s ratio to a value of
593
0.25 when determining the best-fitting and posterior probability densities of the model
594
parameters, while in the current study, we treat this ratio as a free parameter in the
595
inversion. By allowing Poisson’s Ratio to change in our models, the downdip height of
596
the fault decreased, while the depth to its upper edge increased. In addition, the larger
597
best-fitting value for Poisson’s Ratio is associated with a more steeply dipping fault .
598
Second, our previous work did not consider the number of degrees of freedom when
599
estimating P(x|m).
600
distributions to have less variance than those determined in this study. Given the large
601
uncertainties associated with the model itself (discussed below), the higher variance
602
distributions reported in this study likely provide a more realistic estimate of the
603
uncertainties in model parameters than those reported previously .
This had the effect of causing the model parameter posterior
21
604
In addition to expanding our inversion methodology to accommodate
605
observations of bedding rotations and deflections of multiple layers, we explored the
606
impact that spatial resolution and precision of different sources of elevation data might
607
have on our inversions. The standard deviation of the difference between the NED-10 and
608
NED-30 elevations when compared to the ALSM data was found to be ~25% of the RMS
609
error produced by our best-fitting modeled elevations. Thus, given the uncertainties in
610
applying such an idealized model to a complex fold, the NED-10 or NED-30 datasets
611
would have likely been adequate for this analysis. We did find the ALSM data to be
612
invaluable when identifying the extent of the bedding surfaces in the elevation data and
613
the field. In fact, many outcrops that define the fold’s geometry are not visible when
614
using the NED-10 or NED-30 datasets (Fig. 3). Thus, while the precise ALSM mapping
615
may not have increased the precision of elevation measurements used as input for our
616
inversions, it aided in the identification of those areas from which such elevation
617
measurements were extracted.
618
Our methods provide a new, mechanically based way of interpolating the
619
geometry of three-dimensional fault-related folds and imaging the geometry of the
620
underlying faults. As such, our results must be viewed in the context of the simplifying
621
assumptions of the mechanical models upon which our approach is based. In particular,
622
Poly3D idealizes the crust as a homogeneous linear-elastic half-space into which
623
elements that may accommodate slip and opening are embedded. In reality, the rocks
624
currently exposed within the Raplee monocline likely deformed anisotropically,
625
elastically or visco-plastically. In addition, our study infers the geometry of a fault that
626
has not been imaged in the subsurface. In spite of these simplifying assumptions, our
627
model predicts fold shapes consistent with elevations of specific bedding surfaces
628
observed using the ALSM data. To validate such an approach, similar analyses should be
629
performed in areas where subsurface imaging allows comparison of these imaged
630
geometries with those modeled based on surface data.
631
The success of Poly3D in accurately depicting the geometry of Raplee Ridge
632
Monocline indicates either that the approximation of linear elasticity serves as an
633
adequate characterization of the rheology of the folded materials or that different
634
rheological properties may produce similar observed fold geometries by allowing the
22
635
underlying geometry of the slipping fault to change as well. While our Metropolis-
636
Hastings sampler provides estimates of the uncertainties within model parameters that
637
arise from our characterization of the uncertainties in the data, the uncertainties
638
associated with the inappropriateness of the simplified rheology in our model is not
639
captured by our approach. In the future, numerical experiments using finite element
640
models that can consider more complicated rheologies may be used to assess the trade-
641
offs between the loading conditions (represented by the far-field strain conditions and the
642
underlying fault geometry), rheology, and observed surface deflections.
643
from a suite of different rheologies analyzed using the methods presented in this work
644
may be compared with one another to assess the degree of uncertainty that is associated
645
with the choice of the rheology of the folded material. This choice may impact the
646
parameters such as fault geometry that are inferred by our approach, and so by having
647
independent estimates of these values, it might be possible to discern the most
648
appropriate rheological model for these types of folds.
The results
649
650
Conclusions:
651
We present a method that uses the elevations of surface exposures of bedding
652
surfaces of fault-related folds to infer the underlying fault geometry, loading conditions,
653
and Poisson’s Ratio of the folded material, assuming that deformation is due to the
654
release of stress along a frictionless elliptical fault embedded in a homogeneous linear-
655
elastic half-space. This method may be applied to situations in which a single bedding
656
surface is exposed, or multiple stratigraphic surfaces have been deformed and exhumed.
657
In addition to bedding-plane elevations, the orientation of units at various locations in a
658
particular part of the stratigraphic section may be used to further constrain the geometry
659
of the fold and underlying fault. Using a Bayesian Markov-Chain Monte Carlo method,
660
we determined the fault geometry, loading conditions, and elastic parameters that best
661
explain the observed surface deflections, as well as the variation within and covariation
662
between these model parameters. This allowed us to understand the degree to which the
663
data constrain model parameters and the various combinations of model parameters that
664
may produce the observed surface deflections.
23
665
We apply these methods to the Raplee Ridge Monocline, located in southeastern
666
Utah, where high-resolution ALSM data define the geometry of exposed bedding
667
surfaces. While we found that the resolution and precision of the ALSM data are
668
unnecessary for inferring the fault geometry and loading conditions using our approach,
669
these data greatly aid the mapping of the spatial distribution of surface outcrops. Both
670
single-layer and multi-layer inversions agree remarkably well with observations, and
671
image a fault that is broadly constrained to be ~4.5-14 km in downdip height, 13-30 km
672
in along-strike width, and whose tipline was 2.0-9.5 km below the surface at the time of
673
deformation.
674
consistency of our simplified (and likely unrealistic) rheology of the folded material with
675
the observed geometry might suggest that surface deflections may reveal little about the
676
rheology of the fold when viewed in isolation from factors such as fault geometry and
677
loading conditions.
The Poisson’s Ratio was not well resolved by the inversion.
678
24
The
679
References:
680
681
682
683
684
Allmendinger, R. W. 1998. Inverse and forward numerical modeling of trishear faultpropagation folds. Tectonics 17, 640-656.
Bayes, T. 1763. An essay towards solving a problem in the doctrine of chances. Royal
Society of London Philosophical Transactions 53, 370-418.
685
Bullard, T. F. & Lettis, W. R. 1993. Quaternary fold deformation associated with blind
686
thrust faulting, Los Angeles Basin, California. Journal of Geophysical Research
687
98, 8,349-8,369.
688
Bürgmann, R., Kogan, M. G., Steblov, G. M., Hilley, G. E., Levin, V. E. & Apel, E.
689
2005. Interseismic coupling and asperity distribution along the Kamchatka
690
subduction zone. Journal of Geophysical Research, B, Solid Earth and Planets in
691
press.
692
Bürgmann, R., Segall, P., Lisowski, M. & Svarc, J. L. 1997. Postseismic strain following
693
the 1989 Loma Preita earthquake from GPS and leveling measurements. Journal
694
of Geophysical Research, B, Solid Earth and Planets 102, 4933-4955.
695
Cooke, M. L., Pollard, D. D., Erslev, E. A., Fisher, D. M. & Evans, J. P. 1997. Bedding-
696
plane slip in initial stages of fault-related folding. In: Penrose conference on
697
Fault-related folding (edited by Anastasio, D. J.). Journal of Structural Geology
698
19. Oxford-New York : Pergamon, Apr. 1997, Banff, AB, Canada, 567-581.
699
Davis, G. H. 1978. The monocline fold pattern of the Colorado Plateau. In: Laramide
700
folding associated with basement block faulting in the western U. S. (edited by
701
Matthews, V.) 151, Geol. Soc. Amer. Mem., 215-233.
702
Davis, G. H. 1999. Structural geology of the Colorado Plateau Region of Southern Utah,
703
with special emphasis on deformation bands. Geol. Soc. Am. Spec. Paper 342,
704
157 pp.
705
Davis, G. H., Showalter, S. R., Benson, G. S., McCalmont, L. C. & Cropp, F. W. 1981.
706
Guide to the geology of the Salt River Canyon, Arizona. Arizona Geol. Soc.
707
Digest 13, 48-97.
25
708
Gratier, J. P., Guillier, B., Delorme, A. & Odone, F. 1991. Restoration and balanced cross
709
section of a folded and faulted surface by computer program: principle and
710
application. J. Struct. Geol. 13, 111-115.
711
Harris, R. & Segall, P. 1987. Detection of a locked zone at depth on the Parkfield,
712
California segment of the San Andreas fault. Journal of Geophysical Research 92,
713
7945-7962.
714
Hilley, G. E. & Young, J. J. 2008a. Deducing Paleoearthquake Timing and Recurrence
715
from Paleoseismic Data, Part I: Evaluation of new Bayesian Markov-Chain
716
Monte Carlo simulation methods applied to excavations with continuous peat
717
growth. Bulletin of the Seismological Society of America 98, 383-406.
718
Hilley, G. E. & Young, J. J. 2008b. Deducing Paleoearthquake Timing and Recurrence
719
from Paleoseismic Data, Part II: Analysis of Paleosismic Excavation Data and
720
Earthquake Behavior along the Central and Southern San Andreas Fault. Bulletin
721
of the Seismological Society of America 98, 407-439.
722
Huntoon, P. W. 1993. Influence of inherited Precambrian basement structure on the
723
localization and form of Laramide monoclines, Grand Canyon, Arizona. In:
724
Laramide basement deformation in the Rocky Mountain foreland of the western
725
United States, Geol. Soc. of Am. Spec. Paper 280 (edited by Schmidt, C. J.,
726
Chase, R. B. & Erslev, E. A.), 243-256.
727
728
729
730
731
732
Huntoon, P. W. & Sears, J. W. 1975. Bright Angel and Eminence faults, eastern Grand
Canyon, Arizona. Geol. Soc. Am. Bull. 86, 465-472.
Jaeger, J. C. & Cook, N. G. W. 1969. Fundamentals of rock mechanics. Methuen and Co
Ltd, London.
Johnson, K. M. & Johnson, A. M. 2002. Mechanical models of trishear-like folds.
Journal of Structural Geology 24, 277-287.
733
King, G., Stein, R. S. & Rundle, J. B. 1988. The growth of geological structures by
734
repeated earthquakes: 1. Conceptual framework. Journal of Geophysical Research
735
93, 13,307-13,318.
736
Maerten, F., Resor, P., Pollard, D. D. & Maerten, L. 2005. Inverting for slip on three-
737
dimensional fault surfaces using angular dislocations. Bulletin of the
738
Seismological Society of America 95, 1654-1665.
26
739
Metropolis, N., Rosenbluth, A. W., Rosenbluth, M. N., Teller, A. H. & Teller, E. 1953.
740
Equations of state calculations by fast computing machines. Journal of Chemical
741
Physics 21, 1087-1091.
742
Mynatt, I., Hilley, G. E. & Pollard, D. D. 2007. Infering fault characteristics using fold
743
geometry constrained by Airborne Laser Swath Mapping at Raplee Ridge, Utah.
744
Geophy. Res. Lett. 34, L16315.
745
746
747
748
Narr, W. & Suppe, J. 1994. Kinematics of basement-involved compressive structures.
American Journal of Science 294, 802-860.
Pollitz, F. F. 2003. Transient rheology of the uppermost mantle beneath the Mojave
Desert, California. Earth and Planetary Science Letters 215, 89-104.
749
Roering, J. J., Cooke, M. L. & Pollard, D. D. 1997. Why blind thrust faults do not
750
propagate to the Earth's surface: Numerical modeling of coseismic deformation
751
associated with thrust-related anticlines. Journal of Geophysical Research 102,
752
11,901-11,912.
753
754
755
756
Segall, P. & Pollard, D. D. 1980. Mechanics of discontinuous faults. Journal of
Geophysical Research, B, Solid Earth and Planets 85, 4337-4350.
Stein, R. S. 1995. Northridge earthquake; which fault and what next? Nature 373, 388389.
757
Stein, R. S. & Yeats, R. S. 1989. Hidden earthquakes. Scientific American 260, 48-57.
758
Stone, D. S. 1993. Basement-involved thrust-generated folds as seismically imaged in the
759
subsurface of the central Rocky Mountain foreland: Laramide basement
760
deformation in the Rocky Mountain Foreland of the Western United States. Geol.
761
Soc. Am. Spec. Paper 280, 271 pp.
762
Stone, D. S. 2004. Rio thrusting, multi-stage migration, and formation of vertically
763
segregated Paleozoic oil pools at Torchlight Field on the Greybull Platform
764
(Eastern Bighorn Bsin): Implications for exploration. The Mountain Geologist 41,
765
119-138.
766
Thomas, A. L. 1993. Poly3D: A three-dimensional, polygonal element, displacement
767
discontinuity boundary element computer program with applications to fractures,
768
faults, and cavities in the Earth's crust, Stanford University.
27
769
Willemse, E. J. M., Pollard, D. D. & Aydin, A. 1996. Three-dimensional analyses of slip
770
distributions on normal fault arrays with consequences for fault scaling. Journal
771
of Structural Geology 18(2-3), 295-309.
772
773
774
775
Wolkowinsky, A. J. & Granger, D. E. 2004. Early Pleistocene incision of the San Juan
River, Utah, dated with 26Al and 10Be. Geology 32, 749-752.
Ziony, J. I. 1966. Analysis of systematic jointing in part of the Monument Upwarp,
southeastern Utah, University of California, Los Angeles.
776
777
778
779
780
28
781
Figure Captions:
782
783
Figure 1: Shaded relief map (color-coded for elevation), showing the location of Raplee
784
Monocline in southwestern Utah. Upper left inset shows Four Corners area; location of
785
Figure 2 noted in location map.
786
787
Figure 2: The topography of Raplee Ridge, as imaged by the (a) 1-m ALSM-derived
788
data, (b) NED-10 data, and (c), NED-30 data.
789
elevation and map-view scale. (d) Shaded relief map showing the mapped extent of the
790
McKim surface used in single-layer inversions. The location of (x,y,z) points extracted
791
and used to compare accuracy of different datasets are shown as red dots. Note that the
792
mapped surface constrains the shape of the fold at a particular stratigraphic depth.
All elevation maps have the same
793
794
Figure 3: Color-coded, shaded-relief DEM of the center of Raplee Ridge, showing the
795
detailed differences in resolution between the datasets. (a) is the ALSM-derived DEM,
796
(b) is the NED-10 DEM, and (c) is the NED-30 DEM. Color and map-view scales for all
797
panels are identical.
798
799
Figure 4: Comparison of elevations derived from ALSM data with those derived from
800
(a) the NED-30 DEM, and (c) the NED-10 DEM. The residual, or difference between the
801
NED-30 (b) or NED-10 (d) and ALSM data, are shown as histograms, with the mean and
802
standard deviations of these distributions noted in each panel.
803
804
Figure 5: Schematic model setup, showing the definition of each of the parameters used
805
to define the fault geometry and loading conditions in the Poly3D model.
806
parameters shown in figure are defined in text.
Model
807
808
Figure 6: (a) Observed (filled circles) and predicted (shaded background) elevations of
809
the McKim surface that are produced as an initially flat layer is deflected by movement
810
along the best-fitting modeled fault geometry that slips in response to the best-fitting
811
remote loading conditions. (b) Plot of observed versus predicted elevations and (c)
29
812
histogram of residual elevations (observed elevations minus best-fitting modeled
813
elevations) that are produced by the best-fitting modeled fault geometry and loading
814
conditions.
815
816
Figure 7: Posterior probability density functions of each of the model parameters for
817
single layer (McKim surface only) model inversions. In each panel, the solid line denotes
818
the probability density of each model parameter given the elevations observed along the
819
McKim surface, while the dashed lines show the prior probability (assumed uniform in
820
this study) assumed for the model parameters. Numbers in italics represent the value of
821
each parameter for the best-fitting model, while solid bold numbers represent the mean
822
value produced by all simulations. The range shown in parentheses represents the 95%
823
bounds on the model parameter values determined from the posterior probability density
824
function.
825
826
Figure 8: Shaded relief map showing the spatial distribution of the five bedding-plane
827
surfaces exhumed within the fold. From stratigraphically highest to lowest, they are the
828
McKim, Goodrich, Shafer, Mendenhall, and Unnamed surfaces.
829
830
Figure 9: Observed (filled circles) and predicted (shaded background) elevations of the
831
McKim, Goodrich, Shafer, Mendenhall, and Unnamed surfaces produced by movement
832
along the best-fitting modeled fault geometry that slips in response to the best-fitting
833
remote loading conditions. Color and map-view scale are identical between plots. The
834
inferred depth below the surface of each of these layers at the time of deformation is
835
noted in each of the panels.
836
837
Figure 10: Observed and modeled elevations (left panels) and histogram of residual
838
elevations (defined in Fig. 6) for best-fitting fault geometry and loading conditions, when
839
considering multiple layers in the model inversion. Observed vs. predicted elevations
840
and residual histograms shown for the (a) McKim, (b) Goodrich, (c) Shafer, (d)
841
Mendenhall, and (e) Unnamed surfaces.
842
30
843
Figure 11: Posterior probability density functions of each of the model parameters for
844
multiple layer inversions. In each panel, the solid line denotes the probability density of
845
each model parameter given the elevations observed along the McKim surface, while the
846
dashed lines show the prior probability (assumed uniform in this study) assumed for the
847
model parameters. Numbers in italics represent the value of each parameter for the best-
848
fitting model, while solid bold numbers represent the mean value produced by all
849
simulations. The range shown in parentheses represents the 95% bounds on the model
850
parameter values determined from the posterior probability density function.
851
852
853
31
854
Figures:
855
856
857
858
Figure 1
859
860
32
861
862
Figure 2
863
33
864
865
Figure 3
866
867
34
868
869
Figure 4
870
871
35
872
873
874
Figure 5
875
876
36
877
878
879
Figure 6
880
881
37
882
883
Figure 7
884
38
885
886
887
Figure 8
888
889
39
890
891
892
Figure 9
893
894
40
895
896
897
Figure 10
898
899
41
900
901
Figure 11
42
Download