Вы находитесь на странице: 1из 295

Ground Penetrating Radar Imaging of Ancient Clastic

Deposits: A Tool for Three-Dimensional Outcrop


Studies

by

Oluwatosin Caleb Akinpelu

A thesis submitted in conformity with the requirements


for the degree of Doctor of Philosophy
Geology Department
University of Toronto

Copyright by Oluwatosin Caleb Akinpelu 2010

Ground Penetrating Radar Imaging of Ancient Clastic


Deposits: A Tool for Three-Dimensional Outcrop Studies
Oluwatosin Caleb Akinpelu
Doctor of Philosophy
Geology Department
University of Toronto
2010

Abstract
The growing need for better definition of flow units and depositional heterogeneities in petroleum
reservoirs and aquifers has stimulated a renewed interest in outcrop studies as reservoir analogues
in the last two decades. Despite this surge in interest, outcrop studies remain largely twodimensional; a major limitation to direct application of outcrop knowledge to the three
dimensional heterogeneous world of subsurface reservoirs. Behind-outcrop Ground Penetrating
Radar (GPR) imaging provides high-resolution geophysical data, which when combined with two
dimensional architectural outcrop observation, becomes a powerful interpretation tool. Due to the
high resolution, non-destructive and non-invasive nature of the GPR signal, as well as its
reflection-amplitude sensitivity to shaly lithologies, three-dimensional outcrop studies combining
two dimensional architectural element data and behind-outcrop GPR imaging hold significant
promise with the potential to revolutionize outcrop studies the way seismic imaging changed
basin analysis.
Earlier attempts at GPR imaging on ancient clastic deposits were fraught with difficulties
resulting from inappropriate field techniques and subsequent poorly-informed data processing
steps. This project documents advances in GPR field methodology, recommends appropriate data
collection and processing procedures and validates the value of integrating outcrop-based

ii

architectural-element mapping with GPR imaging to obtain three dimensional architectural data
from outcrops.
Case studies from a variety of clastic deposits: Whirlpool Formation (Niagara Escarpment),
Navajo Sandstone (Moab, Utah), Dunvegan Formation (Pink Mountain, British Columbia),
Chinle Formation (Southern Utah) and St. Mary River Formation (Alberta) demonstrate the
usefulness of this approach for better interpretation of outcrop scale ancient depositional
processes and ultimately as a tool for refining existing facies models, as well as a predictive tool
for subsurface reservoir modelling. While this approach is quite promising for detailed threedimensional outcrop studies, it is not an all-purpose panacea; thick overburden, poor antennaground coupling in rough terrains typical of outcrops, low penetration and rapid signal attenuation
in mudstone and diagenetic clay- rich deposits often limit the prospects of this novel technique.

iii

Acknowledgments

I sincerely appreciate the support and guidance of Dr. Andrew Miall throughout the
duration of my graduate studies and especially in seeing this research project to
completion; thank you for giving me an opportunity of a lifetime.
I am particularly grateful to Dr. Nick Eyles for releasing the Ground Penetrating Radar
equipment from his research group throughout the data collection phase of my research;
your guidance in the early years helped set me in the right direction. To Ms. Lynn
Slotkin, I appreciate your assistance from the inception of my graduate program; even
those blunt e-mails were motivational. The painstaking review of the thesis and
suggestions by Dr. Rebecca Ghent and Dr. Uli Wortmann are also well-appreciated.
GPR Field assistance by Thomas Meulendyk helped avoid several field seasons of futile
data collection efforts; I am thankful for helpful tips on GPR data collection and
processing. Field assistance and technical support from Tudorel Ciuculescu remains
matchless; I am highly indebted to you for those long summer days at the outcrops.
Gerald Bryant, I owe a few chapters of my thesis to your assistance, accommodation and
field guidance on all my Utah field trips; I am very thankful for your help.
I am also grateful for funding provided by Ontario Graduate Scholarship, Natural
Sciences and Engineering Research Council, American Society of Petroleum Geologists
(AAPG), International Association of Sedimentologists (IAS) and Society of Exploration
Geophysicists (SEG).

iv

To Toyin and Elizabeth who bore the brunt of my dedicated effort to complete this
research project and for the countless hours of field assistance and proof-reading; I owe
the completion of this research project to you.

Table of Contents

Abstract

ii

Acknowledgement

iv

Table of contents

vi

List of Tables

viii

List of Figures

ix

List of Appendices

xiii

Chapter 1: Research Problem, Objectives and Significance

Chapter 2: Research Methodology:

Pre-survey Assessment

13

GPR Survey Planning

18

GPR Data Acquisition and Recording

28

GPR Data Processing

29

GPR Interpretation Methodology and Radar Stratigraphy

47

Chapter 3: Case Studies:


Whirlpool Formation

65

Navajo sandstone

89

Dunvegan Formation, Pink Mountain (B.C)

114

vi

Shinarump Conglomerate, Hurricane Mesa (Utah)

137

St. Mary River Formation (Monarch, Alberta)

153

Chapter 4: Summary, conclusion and recommendation for future research

170

References

179

Appendix 1: Conductivity, relative permittivity and radar velocity data in sediment

219

Appendix 2: GPR instrumentation

220

Appendix 3: List of published GPR studies on sediments

227

Appendix 4: GPR Data Processing Software and Display

279

Appendix 5: GPR profiles from study locations

280

vii

List of Tables

Table 1

GPR data sheet

28

Table 2

Radar facies chart

61

Table 3

Whirlpool Sandstone lithofacies

75

Table 4

Whirlpool Sandstone Radar Facies

82

Table 5

Navajo Sandstone Radar Facies

105

Table 6

Dunvegan Formation Radar Facies

128

Table A1

Typical radar propagation data in sediment

219

Table A3

List of Published GPR studies on Sediments

227

viii

List of Figures

Figure 1: Relative resolution of different typical geophysical, geological


and remote sensing measurement

Figure 2: Outcrop mapping combined with GPR profile

10

Figure 3: GPR measurement setup

11

Figure 4: GPR antennas size with decreasing antenna frequency

13

Figure 5: Various modes for antenna deployment

22

Figure 6: Single transmitter-receiver common offset survey mode

24

Figure 7: Single transmitter-receiver common midpoint survey mode

27

Figure 8: Artificial structure on downstream accretion element created


by elevation variation along the profile line

31

Figure 9: GPR profile before and after de-wow filtering

32

Figure 10: GPR profile before and after background removal

33

Figure 11: GPR profile before and after deconvolution

35

Figure 12: GPR profile before and after amplitude gain

37

Figure 13: GPR Profile before and after diffraction migration

38

Figure 14: Navajo sandstone outcrop and 100 MHz antenna GPR line

45

Figure 15: Bedding in sediments and associated porosity resulting from


changes in composition, size, shape, orientation and packing
of sediment grains.
Figure 16: Implication of amplitude gain for GPR data interpretation

ix

46
50

Figure 17: Analyses of published GPR studies on sediments

51

Figure 18: GPR Image of a fluvial channel from Dunvegan Formation outcrop

54

Figure 19: Radar facies from Fraser (B.C) and Niobrara River, Nebraska

55

Figure 20: Examples of inclined radar facies

57

Figure 21: Examples of reflection-free radar facies

58

Figure 22: Example of horizontal radar facies

59

Figure 23: Example of hyperbolic reflection

60

Figure 24: Radar facies chart of presumed characteristic reflection


patterns from various sedimentary environments

62

Figure 25: Generalized stratigraphic chart for the Silurian system in New York

69

Figure 26: Early Silurian paleogeographic map of the Niagara Region

70

Figure 27: Location of the Whirlpool sandstone outcrop at Artpark, New York

71

Figure 28: Outcrop photo of the Whirlpool sandstone at Artpark, New York

72

Figure 29: Raw, unprocessed GPR data from the Whirlpool Sandstone

73

Figure 30: Artpark vertical section summarizing the main outcrop observations

74

Figure 31: Location of GPR lines at Artpark State Park, New York

77

Figure 32: Annotated Whirlpool outcrop photo and GPR lines

79

Figure 33: GPR line orthogonal to the Whirlpool Sandstone outcrop at Artpark

81

Figure 34: Three-dimensional reconstruction of the Whirlpool Sandstone


architecture at Artpark integrating outcrop and GPR data

84

Figure 35: Regional paleocurrent pattern for the lower Whirlpool Sandstone

85

Figure 36: Fence diagram display of GPR data from the Whirlpool Sandstone

86

Figure 37: 3D GPR image of the ancient Navajo dune complex

90

Figure 38: Generalized stratigraphy of the Navajo Sandstone

94

Figure 39: Map of Navajo Sandstone location within Utah

96

Figure 40: Map showing field locations of radar profiles at Big Mesa, Utah

98

Figure 41: Outcrop photo of Navajo Sandstone central dune giant cross-bed
compared with smaller cross-bed set at Big Mesa near Moab, Utah.

100

Figure 42: Outcrop photo showing damp eolian interdune deposits overlain
by flat-bedded fresh

101

Figure 43: Outcrop photo and GPR profile at Big Mesa, Utah

103

Figure 44: Radar profiles and perspective views of 3D GPR surveys at Big Mesa 107
Figure 45: Outcrop photo from Moab (Utah) revealing damp phase interdune
deposit from the Carmell Formation

110

Figure 46: Paleogeography during deposition of Dunvegan Formation

118

Figure 47: Lithostratigraphic relationships between the Dunvegan Formation


and the underlying Shaffesbury and overlying Kaskapau formations

119

Figure 48: Map location of Dunvegan Formation outcrop and GPR lines

121

Figure 49: Map showing location of GPR lines at Dunvegan outcrop

123

Figure 50: Dunvegan annotated outcrop photomosaic and GPR line

130

Figure 51: Dunvegan annotated outcrop photo and GPR line

131

Figure 52: Dunvegan annotated outcrop photomosaic and GPR line

133

Figure 53: GPR line and 3D schematic of Dunvegan depositional model, Pink
Mountain

135

Figure 54: Permian and Triassic stratigraphy at Hurricane Mesa, Utah

138

Figure 55: Paleogeographic reconstruction of the Late Triassic within


the southwestern USA.

140

xi

Figure 56: Location of Shinarump Conglomerate outcrop at Hurricane Mesa

142

Figure 57: Location of Shinarump Conglomerate outcrop at Hurricane Mesa


showing GPR line

144

Figure 58: Photomosaic of the Shinarump conglomerate outcrop

148

Figure 59: Uninterpreted and interpreted GPR line at Huricane Mesa

149

Figure 60: Combined Shinarump GPR profile and outcrop Photomosaic

151

Figure 61: St. Mary River Formation Upper Cretaceous Paleogeography

155

Figure 62: Alberta Cretaceous Stratigraphic chart showing St. Mary River

156

Figure 63: Location of St. Mary River formation outcrop

158

Figure 64: Location of GPR lines at St. Mary River formation outcrop

160

Figure 65: Outcrop photo showing St. Mary River Formation channel
sandstone

161

Figure 66: Outcrop photo showing St. Mary River Formation crevasse
splay sandstone

162

Figure 67: Radar image of St.Mary River Formation ribbon channel sandstone

163

Figure 68: Radar profile showing ribbon channels and horizontal reflectors

164

Figure 69: Uninterpreted and interpreted St. Mary River Formation GPR profile

165

Figure 70: Block diagram depicting depositional model of St. Mary River
Formation at the study site near Monarch, Alberta

168

Figure 71: Three-dimensional GPR data display from Navajo Sandstone


interdune deposit at Big Mesa, Utah

175

Figure A1: Mode of operation of Noggin system

221

Figure A2: Shielded PulseEKKO Pro system

222

Figure A3: Mala Geoscience unshielded bistatic antennas

223

xii

Figure A4: Mala Geoscience Rough Terrain Antenna (RTA)

224

Figure A5: Mala Geoscience shielded antennas

225

Figure A6: GSSI monostatic and bistatic 100 MHz antennas

226

Figure A7: Analyses of 150 published GPR studies on sediments

278

Figure A8: GPR profiles from Artpark (New York, USA)

280

Figure A9: GPR profiles from Big Mesa, near Moab (Utah, USA)

281

xiii

List of Appendices

Appendix 1:

Radar propagation data in sediment

219

Appendix 2:

GPR Instrumentation

220

Appendix 3:

List of published GPR studies on sediments

227

Appendix 4

GPR Data Processing Software and Display

279

Appendix 5: GPR profiles from study locations

xiv

280

Chapter 1
Research Problem, Object and Significance

1.1 Introduction
The need for better prediction of reservoir architecture and distribution of lithological
heterogeneity in aquifer characterization, hydrocarbon primary field development and
enhanced recovery projects is driving a burgeoning interest in outcrop-based studies of
ancient clastic deposits both as analogues for subsurface reservoirs and as a tool for
refining existing facies models. This research is driven by the need to account for the
variability and architectural complexity in subsurface clastic reservoirs, especially for the
purpose of petroleum resource evaluation and development, where conventional seismic
resolution is too poor to unambiguously delineate sandbody geometry or map their
internal heterogeneity.
Subsurface interpretation of reservoir architecture currently relies on seismic, well logs
and core data which are rarely closely spaced enough to permit unequivocal
interpretation. Practical limitations of most sources of high-resolution data rarely permit
the close spacing required for unambiguous interpretation (Figure 1). Advances in
seismic data acquisition, data processing and three-dimension data visualization in the
last two decades have significantly improved the resolution of reservoir information
observable between wells at current well spacing. Also, development of procedures for
extracting useful information from attributes of the seismic signal like amplitude, phase
and frequency have in many cases facilitated delineation of stratigraphic traps and other
subtle geological features as well as improvement in determining porosity distribution
1

and fluid content in reservoirs. In spite of these advances, predicting sub-seismic scale
heterogeneities in reservoirs is still a daunting challenge for many primary field
development, enhanced oil recovery and aquifer characterization projects.

Figure 1: Chart showing trade-off between the relative resolution of the information
obtained using different typical geophysical, geological and remote sensing measurement
acquisition approaches and the relative scale of the investigations for which those
acquisition geometries are typically used. (Image taken from Rubin and Hubbard, 2005).
The development of affordable digital imaging technologies as well as advances in
photogrammetry, aerial photography, LiDAR scanning, satellite imagery and positioning
have significantly improved outcrop studies. Increased spatial accuracy from satellite and
laser positioning systems provides access to geostatistical and geospatial analyses that
can inform hypothesis testing during fieldwork. High-resolution geomatic surveys allow
photorealistic outcrop images to be captured from several perspectives and interpreted
using novel visualization and analysis methods. Inaccessible outcrops can now be

captured and brought to the geological laboratory via LiDAR scanning for further
measurements and analysis; bedding orientations and dips can then be measured at the
laboratory, considerably saving time spent doing fieldwork.
Despite these developments, outcrop studies remain largely two-dimensional, as most of
these novel techniques are not penetrative; they only reveal outcrop-face sedimentary and
stratigraphic features that can only be inferred beyond the outcrop-face with much
difficulty. The need for a more reliable tool of reconstructing fluvial architecture from
outcrops led to Mialls (1985) proposition of architectural element analysis following
the pioneering work of Allen (1983). Architectural elements are associations of
genetically related facies, which have some environmental significance. In depositional
systems, they are packets of genetically related strata, which define depositional elements
larger than individual bedform and smaller than channels. In outcrops, they are defined
by grain size, bedform composition, internal sequence, and most importantly, their
external geometry. This mapping technique has not only been applied to ancient fluvial
deposits as proposed by Miall (1985); it has been successfully applied on outcrops of
diverse depositional origin (Kocurek, 1981; Nio and Yang, 1993; Clark and Pickering,
1996). While there have been many outcrop studies with geometry and lithofacies of
channel and channel-belt deposits determined in outcrops described as analogues for
subsurface strata (Tyler and Finley, 1991; Mjos and Walderhaug, 1993; Dreyer, 1993a,
1993b; Dreyer et al., 1993; Robinson and McCabe, 1997; Willis and Gabel; 2003), the
reliability of many of these studies has been questioned due to ambiguity in determining
sandbody geometry, dimensions, stacking pattern and heterogeneity in outcrop sections
(Bridge 2006)

The need to address the two dimensional limitation of outcrop studies motivated the
recent interest surge in Ground Penetrating Radar (GPR) imaging. However, due to the
operational difficulties of GPR imaging on outcrops and the susceptibility of GPR signals
to diffraction resulting in noise-laden images, very few outcrop-based GPR studies
(Appendix 3) have been successfully conducted (Corbeanu et al., 2001; Szerbiak et al.,
2001; Lee et al, 2009). Consequently, much of the current GPR effort has focused on
modern environment and Quaternary sediments (Beres and Haeni ,1991; Jol and Smith,
1991; Gawthorpe et al 1993; Huggenberger et al., 1994; Beres et al., 1995; Bridge et
al.,1995,1998; Van Overmeeren 1998; Bristow et al.,1999,2000b; Baker et al., 2001;
Corbeanu et al., 2001, 2002; Nobes et al.,2001; Hammon et al.,2002, Hornung and
Aigner 2002; Best et al, 2003; Cardenas and Zlotnik, 2003; Heinz and Aigner 2003;
Skelly et al., 2003, Woodward et al., 2003) as radar profiling is more convenient in
modern environments usually of less rugged terrain and modern sediments are less prone
to diagenetic alterations, which often make sedimentological interpretation of GPR data
from ancient clastic deposits a Herculean task. Although there has been an improvement
in our understanding of many ancient deposits via GPR studies of modern sediments, the
varied and unpredictable preservation potential of facies units in their modern setting
limits the usefulness of these studies. It is the preserved ancient record that is ultimately
what reservoir geologists are interested in. Miall (2006) sounded a caveat in this regard
and underscored the continued role of three-dimensional studies of ancient outcrop
analogues in providing a realistic database for predicting sub-surface alluvial
architecture. Outcrops are unarguably still the most important source for learning more
about the geometry, size, and distribution of sedimentary bodies because they permit

direct visual observation of rock types and their spatial arrangement, thereby providing
invaluable insight into reservoir complexities for subsurface modelling (Geehan, 1993).
Key components of significant relevance to reservoir characterization include facies
types, macro-scale lithological heterogeneity, channel dimensions, sandbody geometry
and their stacking pattern in relation to local and regional controls. These outcrop-scale
features control fluid-flow behavior in petroleum reservoirs, and aquifer and outcrop
analogs have been shown to provide valuable insight into reservoir architecture
(Grammer et al, 2004). While outcrop studies typically provide valuable data on vertical
dimensions of architectural elements within a depositional system, rarely do they yield a
three-dimensional picture of potential reservoir distribution especially from the
standpoint of aerial dimensions. Knowledge of the internal architecture is generally
limited because the exposures are either strike-oriented, dip-oriented, or at an orientation
oblique to these directions. These limitations have significantly weakened the usefulness
of outcrop studies as analogs for subsurface clastic reservoirs in the past; successful
application of GPR imaging on outcrop will provide an opportunity to validate current
clastic facies and architectural models that are mostly actualistic and conceptual
constructs.
The purpose of this research is four-fold:

to demonstrate that sedimentological and stratigraphic data can be obtained from


outcrop-based GPR profiles and dispel the widely held notion that GPR imaging
only works on Quaternary sediments,

to demonstrate the effectiveness of GPR imaging in addressing the limitations of


two dimensional outcrop studies of ancient non-marine sandstones in a manner
that has not been successfully explored before,

to develop a field methodology for outcrop-based GPR surveys, and

to provide a reliable framework for stratigraphic interpretation of GPR images as


a guide for future research.

By incorporating ground penetrating radar data with outcrop architectural element


analysis, data that are often difficult to unambiguously obtain, such as local paleoflow
variability, channel belt width, channel dimension, sinuosity, sandbody connectivity,
stacking patterns and lithological heterogeneity can now be readily obtained. Successful
application of GPR imaging on ancient clastic deposits will have far-reaching
significance; being able to obtain data such as true channel dimensions, channel
sinuosity, incised valley dimensions and their internal geometry will not only provide
hard data to validate or dispel many of the current clastic facies models based mostly on
studies of modern environment and Quaternary deposits; it will also test actualisticallyderived empirical equations (often used to predict channel belt width in the subsurface)
that relate maximum channel depth, channel width, and channel-belt width (Collinson,
1978; Lorenz et al., 1985, 1991; Fielding and Crane, 1987; Bridge and Mackey, 1993b).
Two features of GPR imaging make the technology exceptionally suitable for addressing
the limitations of outcrop study: its high resolution that makes imaging of sedimentary
structures, architectural elements and in some cases entire depositional system feasible,
as well as radar signals sensitivity to clay content, which makes it possible to map macro
scale lithological heterogeneities in outcrop.

Although GPR imaging has been used intermittently for outcrop studies for over 25
years, the technology was not widely embraced as GPR instrumentation and data
processing technology was in its infancy and many of the GPR profiles generated were
not thoroughly processed to remove extraneous reflections required to ensure
unequivocal sedimentological and stratigraphic interpretation (Pratt and Miall, 1993;
Stephens, 1994). In the last few years, there have however been significant advances in
GPR instrumentation, data processing and display. Recently developed GPR systems are
lightweight, portable, robust and digital. In addition, the fact that GPR data can be viewed
and test-processed in real time during surveys ensures that both quality and results can be
assessed in the field. These advances as well as the availability of GPR processing
applications with two and three-dimensional visualization and interpretation aids have
enhanced rapid acquisition of continuous, shallow subsurface profiles that are ideally
suited for the investigation of sediments, shallow stratigraphy and sedimentary
architecture from both a 2-D and 3-D perspective.
Outcrop studies will undoubtedly benefit more from full-resolution three-dimensional
GPR surveys which produce 3D volumes that can be sliced and rotated to reveal features
such as map views of channel geometry, dimensions, sinuosity, anabranching and
temporal changes in depositional patterns. However, most GPR surveys are still twodimensional due to operational difficulties of 3D GPR surveying on outcrops. The few
three-dimensional surveys conducted to date are limited to small areas because of the
need for decimeter scale elevation survey, high data volumes and relatively large time
requirements for data collection and processing. Despite this, the effectiveness of GPR
imaging for characterizing reservoir geometry and heterogeneity through studies of

outcrop analogs has been demonstrated with 2-D profiles from earlier studies (Meyers et
al., 1996; Lunt and Bridge, 2004) as well as with fence diagrams constructed from
staggered GPR profiles obtained in different directions behind outcrop face, groundtruthed with outcrop sedimentological and architectural element mapping.

Chapter 2
Research Methodology

2.1 Introduction
The reliability of GPR outcrop based reservoir analogue studies depends on the quality
of outcrop data and GPR survey as well as subsequent processing of the GPR data. Many
of the earlier outcrop studies utilizing GPR (Pratt and Miall, 1993; Stephens, 1994; White
et al, 2004; Zeng et. al. 2004) were plagued with noise making it difficult to correlate
outcrop observations to radar reflectors. This is attributed to susceptibility of
electromagnetic signals to diffractions resulting from fractures as well as diagenetic
cementation in rocks (Smith et al., 2006). Diffraction occurs at discontinuities of
reflectors (such as fractures) and objects whose dimensions are small compared to the
transmitted electromagnetic wavelength. Figures 2 A and B illustrate the prevalence of
diffraction noise in earlier outcrop GPR studies.
Advances in GPR instrumentation and data processing over the last decade have
significantly improved GPR data quality. The field methodology for this research
involves obtaining photomosaics of laterally extensive sandstone exposures
supplemented by GPR profiles behind the mapped outcrop face. The photomosaics serve
a dual purpose: they act as base maps for detailed mapping of bounding surfaces and
architectural elements in the outcrop face, as well as providing ground truth control
required to ascertain the fidelity of the radar reflections.

20 m

Diffraction noise

Diffraction noise

Figure 2: (A) An example of outcrop face mapping of major bounding surface combined
with GPR profile behind the cliff face. (Image taken from Zeng et al, 2004) (B) Groundpenetrating radar traverse measured about 100 m behind a dip-parallel outcrop of the
upper Frewens sandstone body. (Image taken from White et al., 2004); note the difficulty
in correlating GPR reflections with outcrop bounding surfaces.
Architectural-element analysis involves delineating major bounding surfaces as well as
recognition of suites of geometric elements characterized by distinctive facies
assemblages, internal geometry and external form. Over the years, this technique,
integrated with paleocurrent data, has proven to be a useful descriptive framework and a
valuable tool for comprehensive interpretation of ancient clastic deposits (especially

10

fluvial and eolian deposits) and unraveling their three-dimensional architecture (Miall,
1996)
Acquisition and processing of GPR data are the more laborious phases of the research
involving selection of the appropriate GPR system as well as efficient field survey and
data processing procedures. GPR is an electromagnetic method that, in many ways, is
similar to seismic reflection survey. A transmitting antenna radiates an electric pulse into
the ground that, under ideal conditions, behaves kinematically similar to an acoustic
wave. The pulse is transmitted, reflected, and sometimes diffracted by features that
correspond to changes in the dielectric properties of the earth. The waves that are
reflected and diffracted back toward the earths surface may be detected by a receiving
antenna, amplified, digitized, displayed, and stored for further analysis (Figure 3)

Figure 3: GPR measurement setup.


Reflections and diffractions in a GPR section are generated by changes in electrical
properties of the rock across lithofacies boundaries; visible sedimentary structures or

11

sometimes, sequence boundaries. The transmitter and receiver units are usually separate,
so the survey design is flexible. Data can be collected either in continuous or step mode.
In continuous mode, the antennas are dragged over the surface, whereas in step mode, the
antennas are placed on the ground in steps progressively along the profile. GPR surveys
usually proceed by recording a trace at each of a large number of survey points along a
line (or over a grid) with a fixed transmitter-receiver offset. As in reflection-seismic data,
GPR profiles may be plotted directly, or combined and plotted as a volume. The main
differences are that the scale of a GPR survey is about three orders of magnitude smaller
than that of a reflection seismic survey, and the resolution is correspondingly higher,
making the technology suitable for outcrop scale imaging.
The choice of antenna frequency is a tradeoff between the depth of target beds and the
intended resolution of the survey; the lower the frequency of the antenna, the poorer the
resolution but the greater the depth of penetration (Davis and Anan 1989; Jol, 1995).
Currently commercially available frequencies are between 12.5 and 1200 MHz, the time
sample increment is usually about 1 ns, the propagation velocities are one-quarter to onehalf that of light in a vacuum, and the average depth of penetration ranges from 5 to 40m,
depending primarily on the electrical conductivity and water content of the subsurface
materials. Generally, the lower the frequency, the bulkier the GPR equipment (Figure 4) ;
this implies that low frequency antenna (below 100 MHz) ideal for regional-scale
geological imaging may require more than one person to handle during surveys especially
when Common Midpoint (CMP) Surveys are being conducted.

12

Attenuation of GPR signals in rocks increases with water saturation and decreasing grain
size; wet and clayey rocks significantly attenuate radar signals thereby reducing depth of
penetration.

Figure 4: GPR antennas get bulkier with decreasing antenna frequency. (Image taken
from http://www.sensoft.ca/products/pulseekko/pulseekkopro.htmll ; - accessed August
6, 2010).

2.2 Pre-survey Assessment


While GPR has proven to be an invaluable tool for high-resolution imaging in outcropscale studies, it does not always work at every survey location especially on outcrops
with shaly lithology or in those that are highly conductive. Before planning GPR surveys,
it is important to investigate whether GPR would work at the location and on the target
outcrop. Prediction of whether GPR imaging is ideal for the problem at hand is not clear
13

cut; it is easier to rule out situations where georadar is totally unsuitable than to state with
certainty that ground radar imaging will be successful. The following criteria must be
thoroughly evaluated before embarking on a GPR survey program:

2.2.1 Depth of Formation of Interest


How deep is the formation of interest? is perhaps the most important question in planning
GPR surveys. Many geological formations that are potential targets for GPR imaging are
located beneath several metres of overburden (for instance, many outcrops of Whirlpool
Sandstone and St. Mary River Formation) or beneath formations that might significantly
attenuate transmitted radar signals. The lowest frequency antenna currently available for
GPR surveys is 12.5 MHz (Figure 4) and this is limited to less than 50 metres penetration
depth in most clastic rocks, even in dry, clay-free lithologies. In addition, the attenuation
characteristic of the overburden and target medium is critical to penetration depth of
radar signals. A conservative rule-of-thumb is that radar will be ineffective if the actual
target depth is greater than 50% of the maximum range. A rough guide to penetration
depth D is

(2.1)

(Annan, 2004)

where conductivity, is in mS/m. This equation is not universal but is applicable when
attenuation is moderate to high (attenuation > 0.1 dB/m or > 1 mS/m) which is typical
of most lithologies (Annan, 2004)

14

In geological applications, the parameters which have the greatest influence on the
electrical properties, and hence on penetration depth, velocity and reflector characteristics
are: (i) water saturation, (ii) clay content and (iii) pore water salinity. Dry, clay-free
sediments generally have higher velocities of propagation and lower attenuation
coefficients than wet strata. Depth of penetration is therefore highest in dry, porous
sediments (e.g. limestone, where probing depths may theoretically be up to 100 m) and
lowest in water-saturated clay where depth of penetration may be less than 1 m (Cook
1975). In some cases, sand-rich formations can have diagenetic clays not obvious in
outcrop and could result in severe signal attenuation; hence, the need for pre-survey
assessment.

2.2.2 Contrast in electrical properties required to


reflect or scatter a detectable amount of energy
GPR investigates the subsurface by making use of electromagnetic signals, which
propagate into the subsurface. Because reflections of radar signals are generated by
changes in electromagnetic impedance (contrast in electric and magnetic properties)
across the lithologies that the signals travel through in the same way seismic reflections
are generated across lithological contacts with acoustic impedance changes, detectable
reflections are expected in geological formations with contrasting lithologies. When an
electromagnetic wave propagates through the ground and encounters a surface where the
electric and/or magnetic properties of the ground change, part of its energy will be
reflected and part of it will be transmitted; this is controlled by the electrical and
magnetic properties of the rock layers. In most geological applications of GPR, electrical
properties tend to be the dominant factor controlling GPR responses; magnetic variations

15

are usually weak, although occasionally, magnetic properties can affect radar responses.
This is often the case when imaging diagenetically altered formations containing iron
oxides; hence, it is important to be cognizant of such magnetic effects as they can
significantly distort interpretation. Transmission of electromagnetic signals through rocks
is controlled by three parameters: dielectric permittivity, electrical conductivity and the
magnetic permeability. Dielectric permittivity is the ability of a material to store and
release electromagnetic energy in the form of electric charge or its ability to restrict the
flow of free charges or the degree of polarization exhibited by a material under the
influence of an applied electric field. Conductivity is the ability of a material to pass free
electric charges under the influence of an applied field; this in rocks typically occurs via
dielectric conduction (in resistive lithologies which require the atoms to slightly polarize
to produce displacement currents) and electrolytic conduction (dominant in moist or wet
lithologies). Magnetic permeability describes how intrinsic atomic and molecular
magnetic moments respond to a magnetic field. Magnetic permeability in most clastic
rocks is very low and its effect on electromagnetic signal transmission can be ignored.
Both dielectric permittivity and electrical conductivity are strongly dependent on water
content (which is largely dependent on lithology); hence radar reflections patterns are
closely linked with lithological variability in rocks. For outcrops with minimal
lithological heterogeneity, it is sometimes efficient to conduct a preliminary small scale
survey at the outcrop to determine optimal target depth, resolution and appropriateness of
GPR for the location before embarking on detailed survey.

16

2.2.3 Noise sources that could preclude the use of GPR


Recorded GPR signals during surveys can be masked by electromagnetic signals from
nearby sources such as a radio transmitter or an electric power line located near the
survey site; in such cases, external signals may saturate the sensitive receiver electronics.
Radio transmitters are potential sources of interference and powerful radio signals can
overwhelm receiver electronics. Mobile phones are increasingly becoming a ubiquitous
form of interference; proximity to metal objects can also be disastrous for GPR survey.
Reflections can come from objects away to the side (sideswipe) and may be very strong if
metallic reflectors are involved. Surface features can produce strong sideswipe resulting
from substantial radiation of energy along the ground/air interface if ground conductivity
is high. Shielded antennas are very useful in such situations and are available in the 100
MHz frequency range and above. The shields are about the same size as the antenna and
use absorbing material to damp out the undesired signals. At lower frequencies, antenna
size and portability makes shielding impractical (as antenna size increases with
frequency); these considerations probably explain why there are no commerciallyavailable shielded GPR systems for antenna frequencies below 100 MHz. Practical
limitation of portability also explains why most shielded GPR systems consisting of both
transmitter and receiver antenna elements are housed in a single housing; hence, bistatic
antennas useful for common midpoint surveys are rarely shielded.

2.2.4 Outcrop Accessibility


How feasible is hauling bulky GPR and survey equipment to outcrop top? While remote
outcrops can be reached by hiking for kilometers or climbing where necessary during
geological mapping or routine outcrop study, it is rarely possible carrying bulky GPR and
17

surveying equipment to the target outcrop. Difficulty in accessing outcrops is one of the
major factors explaining why there has been limited application of GPR imaging to
solving three dimensional outcrop problems. Although some can be reached by short
hikes, many outcrops ideal for GPR surveys can only be reached by ATVs (All Terrain
Vehicles) or helicopters; hence, accessibility to outcrops should be thoroughly evaluated
before embarking on GPR surveys.

2.3 GPR Survey Planning


Once the feasibility of GPR survey at an outcrop has been ascertained, proper design of
the survey as well as optimizing data acquisition to meet expectations and honor surface
constraints is the next challenge of outcrop-based GPR studies and requires thorough
planning.
In planning for GPR surveys on ancient sedimentary deposits, the following factors
critical to obtaining high-fidelity profiles must be considered:

Choosing the right GPR system there is currently a wide range of


commercially available GPR systems to choose from (see Appendix 2); it is
therefore important to choose a GPR system suitable for the purpose of the study.
Stratigraphic studies requiring time-depth conversion via common mid-point
surveys (data collection by moving the antennas progressively apart at an equal
offset from a central survey point) which gives the flexibility of calibrating radar
profiles in metres cannot be conducted with most shielded systems as they are
mostly designed for common offset (taking data by moving both antennas in the
same direction with the same antenna separation along the survey line) method of
data collection. Choosing a system with the appropriate antenna frequency for the

18

project is also critical; this is a trade off between the depth of the geological
formations being studied and the resolution required in the study. Low frequency
antennas (below 100 MHz) are better-adapted for low resolution deeper target
(20-50 metres deep) surveys while high frequency antennas (above 100 MHz) are
better-suited for high resolution shallow (less than 20 metre deep targets). As a
rule of thumb, the vertical resolution is theoretically one-quarter of the
wavelength = v/f, where v is the velocity of the electromagnetic wave in the
rock (see Appendix 1 for typical velocities in different lithologies) and f is the
center frequency of the GPR antenna. It is also important to note that the
bulkiness of most GPR systems increase with decreasing antenna frequency
(Figure 4); this should be considered when planning GPR surveys as such low
frequency systems may require two or more people to handle.

Surface Elevation Uneven topography is characteristic of most outcrop top.


GPR surveys require flat outcrop tops or centimeter-scale topographic survey over
the GPR transects or across the area covered by the 3D grid if three-dimensional
GPR survey is being planned. This is a major challenge in conducting GPR
surveys on outcrops. Collection of topographic data is an essential component of
GPR survey especially where the outcrop surface is not flat and this could gulp a
significant chunk of survey time. Improper incorporation of topographic data into
the GPR lines may result in inaccurate velocities for static correction, which often
results in introduction of artificial structures and false bedding into the radar
profiles (Figure 8). A variety of instruments can be employed to obtain elevation
data for GPR processing; these include optical levels, total station, laser levels,

19

and GPS. In areas with subdued topography (less than 5 m change in elevation),
laser levels provide a fast and accurate method for collecting topographic data. In
more complex terrains, a total station provides greater flexibility and higher
accuracy; optical levels are also ideal but increase survey time. Where GPS units
are being considered for GPR surveys, Differential GPS units with centimeterscale accuracy are required. GPS surveys significantly reduce GPR survey time
especially in three-dimensional surveys as GPS measurements obviate the need
for a survey grid. In addition, with GPS measurements, real world locations of
GPR profiles can be easily displayed in Google Earth or imported into digital
topographic maps. It is however important to note that GPS receivers have
difficulty working under thick vegetation and require post processing. Some highresolution GPS tools also have compatibility problems with GPR units; if possible
GPS unit should be connected to the GPR equipment and tested before setting out
on field surveys. Most of the GPR surveys for this research were conducted on
flat surfaces requiring minor or no elevation correction.

Mode of data collection and station spacing while GPR data can either be
collected by dragging the antenna over the surface (continuous mode) or by
progressively moving the antenna along the survey line (step mode); the step
mode of data collection is less prone to data degradation and more suitable for
stratigraphic studies although it is more time-consuming. GPR antennas couple to
the ground by the concentration of electric field in the lower medium, and best
coupling is achieved by laying the antenna flat on the ground (Annan et al., 1975);
this is usually better accomplished in step mode. In step mode, data collection

20

distance between survey points must be small enough (usually between 0.25 to
0.5 metres) to be able to resolve steeply dipping and small features as well as to
minimize spatial aliasing of radar reflections. Due to less survey time required,
many GPR surveys are done in continuous mode. Most shielded GPR systems are
designed for continuous data collection mode; therefore, for surveys where
shielded antennas are being considered, care should be taken to ensure optimal
coupling with the ground for efficient transmission of radar signals.

Step size - Spatial resolution places a constraint on the survey design and
selection of antenna centre frequency. Step size (the distance between each data
collection point, also known as station spacing) is extremely important and should
be included in the survey design process. In order to ensure that the ground
response is not spatially aliased in sedimentary studies, a maximum step size of
one metre (less than 1m where antenna frequencies higher than 200MHz are being
employed) should be used to provide detailed horizontal resolution of sedimentary
structures, yet allow profiles to be completed in a timely manner. A typical survey
with 100MHz antenna theoretically should have a step size of 0.25m; however,
larger step sizes can be used where the subsurface stratigraphy is composed of
continuous horizontal layers. If the step size is too large, the data will not
adequately define steeply dipping reflectors or diffraction tails. For 3D GPR
surveys, using a GPR grid spacing larger than a quarter-wavelength in the
presence of diffractions not only decreases horizontal resolution, it also creates
aliased dipping events, which produce migration artifacts that blur the real
reflections. Imaging dipping beds and bounding surfaces is critical to many

21

geological studies; hence, very low step sizes are required for thorough definition
of dipping beds and surfaces. The decision to increase or decrease the step size
should be based on a variety of factors including: size of the sedimentary feature
being investigated, dip angle, and areal extent of the survey. From a practical
viewpoint, increasing the station interval reduces data volume and survey time,
yet from a data interpretation standpoint, adhering to the Nyquist close sampling
interval (which requires that measurements be spaced in all horizontal directions
near a quarter-wavelength of the highest signal and noise frequency content to
avoid spatial aliasing) is very important.

Antenna Orientation - Antennas used for most GPR surveys are dipolar and
radiate with a preferred polarity. The antennas are normally oriented so that the
electric field is polarized parallel to the long axis or strike direction of the target;
there is no optimal orientation for an equidimensional target.

Figure 5: Illustration of the various modes for antenna deployment - E field aligned
along the antenna axis. (Image taken from Annan 2002).

22

Antenna orientation does affect the quality of collected data (Lutz et al., 2003) and it
should be taken into consideration when planning GPR surveys. The various
arrangements of GPR antenna deployment are illustrated in Figure 5; the most
common orientation that provides the widest angular coverage of a subsurface
reflector is the perpendicular broadside to direction of survey approach (PR-BD).
This orientation is also the easiest for surveying in sedimentary environments. In
some instances, it may be advisable to collect two data sets with orthogonal antenna
orientations in order to extract target information based on coupling angle. If the
antenna system is one which attempts to use a circularly- polarized signal, the antenna
orientation becomes irrelevant; however, as most commercial systems employ
linearly polarized antennas, orientation can be important. Maintaining a consistent
antenna configuration and electronic connections throughout a survey is important
because this will allow changes in polarity, due to increases or decreases in
permittivity (impedance) with depth, to be determined. Most shielded GPR units have
fixed antenna orientation, hence no need of deciding optimal antenna orientation.
GPR units with separate antennas however require determination of optimal antenna
orientation.

2.4 GPR Survey Design


Radar surveys for stratigraphic purpose are carried out either in common offset or
common midpoint mode; however, fixed offset, single-fold common offset reflection
profiling using a single transmitter and a single receiver is the most common. Multiple
source and receiver configurations are seldom used for specialized applications. Common
offset surveys usually deploy a single transmitter and receiver, with a fixed offset or

23

spacing between the units at each measurement location. The transmitting and receiving
antennas have a specific polarization character for the electromagnetic field generated
and detected. The antennas are placed in a fixed geometry and measurements made at
regular fixed station intervals, as depicted in Figure 6. Data on regular grids at fixed
spacing are normally needed if advanced data processing and visualization techniques are
to be applied. The parameters defining a common offset survey are GPR center
frequency, the recording time window, the time-sampling interval, the station spacing,
the antenna spacing, the line separation spacing, and the antenna orientation. During
surveying, antennas are either dragged along the ground and horizontal distances
recorded on a time-base, which can be converted to a distance-base through manual
marking or measuring wheel, or they are moved in a stepwise manner at fixed horizontal
intervals (step size). Step-mode operation generates more coherent and higher
amplitude reflections, as antennas are stationary during data acquisition. This ensures
more consistent coupling between the antennas and the ground as well as better trace
stacking (Annan and Cosway, 1992).

Figure 6: Illustration of single transmitter-receiver common offset survey mode. (Image


taken from Annan, 2009).
24

The MAL antenna used for this research (250 MHz MAL Geoscience shielded
antenna-Appendix 2) is adapted for single-fold reflection profiling; this GPR unit has
fixed antenna distance and orientation, which makes it impossible to conduct common
midpoint surveys.
Common midpoint (CMP) soundings are primarily used to estimate the radar signal
velocity versus depth in the ground, by varying the antenna spacing (commonly referred
to as offset) and measuring the change of the two-way travel time. At all distances
between source and receiver with a fixed midpoint, most of the energy that is reflected
by, not too steeply dipping, subsurface reflectors comes from the same points straight
below that given midpoint (Figure 7A). From the differences in arrival time versus offset
very accurate velocity estimates can be obtained from all layers. When a CMP
measurement is carried out, different events can be distinguished in the received signal,
when plotted on a travel time versus offset plot as depicted in Figure 7C. These are the
results of the air wave (1), the ground wave (2), the direct reflection from an interface (3),
and the critically refracted airwave (4). Since the velocity in air is always greater than the
largest velocity in the ground, a critical angle of incidence exists where the transmitted
wave travels in the air along the interface. These angles only exist when the wave travels
from a slower medium to a faster medium, such as a reflection from below the ground
that travels toward the earth surface. In a common-offset survey, the above mentioned
events interfere and cannot be identified separately. On the other hand, in a CMP
measurement they can be recognized. A theoretical result is depicted in Figure 7C. From
the ground wave, the velocity of the waves just beneath the surface (through the soil if
there is a thin soil layer) can be extracted by determining the slope of the event,

25

(2.2)

(Ver der Kruk et al, 2004)

(2.3)

(Ver der Kruk et al, 2004)

The rock velocity can be calculated from

where a and b denote two different offsets (Figure 7B) and t(a) and t(b) are the
corresponding two-way travel times.
While CMP surveys have the advantage of allowing velocities to be calculated from the
variations in reflection time with offset, they are very slow, and therefore rarely used in
geological GPR studies; they are however routine in seismic reflection, where geophones
can be laid out in large numbers.
Two major benefits of this survey mode are that CMP stacking can improve signal-tonoise (Fisher et al, 1992a), and that a full velocity cross section required for time-depth
conversion of the GPR profiles can be derived (Greaves et al., 1996). Multifold GPR
surveys are seldom performed because they are time-consuming, more complex to
analyze, and not cost-effective (most of the cost benefit is obtained with well-designed
single-fold surveys).

26

Figure 7: (A) Diagram illustrating single transmitter (s) -receiver (r) common midpoint
survey mode (B) CMP configuration for GPR (C) Simple GPR CMP traveltime diagram
for estimating velocities. (Images taken from Ver der Kruk et al, 2004).

27

2.5 GPR Data Acquisition and Recording


GPR data have the advantage of being viewed directly as they are being recorded in the
field and many GPR equipment control units have basic editing and data processing
capability. Despite this advantage over seismic data acquisition, radar data acquisition
parameters such as antenna frequency, sampling frequency, number of stacks and time
window can not be changed during data processing. To ensure that data acquisition time
is effectively used, laptop computers with installed data processing applications should be
taken to the field and GPR data edited, checked and test-processed in the field before the
completion of field surveys as it is often difficult to fix problems in GPR data that are due
to poor acquisition parameters without having to return to the field.
Detailed record of GPR lines, outcrop photos and their locations should be accurately
documented in the field as the effectiveness of data processing and geological
interpretation hinge on proper documentation of GPR line locations and other outcrop
observations. GPR data sheet such as Table 1 should be completed as each GPR line is
being recorded.
Table 1: Example data sheet that can be used for collecting data from daily GPR surveys

28

2.6 GPR Data Processing


GPR signals are usually processed like seismic data although the two are slightly
different. GPR data are most often treated as scalar although as electromagnetic fields,
they are vector quantities. Hence, GPR signals may behave differently due to frequencydependent absorption and phase changes at reflections.
In the early days of GPR data processing, much of the data processing was done with
seismic data processing software. However, GPR-specific processing software is now
commercially available from packages offered by manufacturers of GPR or as add-ons
for seismic processing software. For this study, REFLEX and GPR Slice processing and
display software were used for data processing and perspective display.
Like seismic reflection data, GPR data require processing aimed at sharpening the signal
waveform by improving the signal to noise ratio of the radar profiles (Reynolds 1997).
The amount of processing depends on the quality of field data obtained, and this can
range from basic processing steps to the more complex application of data processing
algorithms. Data obtained with shielded antennas are often less cluttered with extraneous
signals and usually require less rigorous data processing. In most instances, basic
processing steps can be applied in real-time during data acquisition; however, these might
not be enough to remove extraneous noise from the acquired data. A detailed discussion
of GPR processing steps is important, as the final output of GPR surveys is dependent on
processing steps; radar facies must therefore be compared only between data processed
with similar processing steps and criteria.
Definition of radar facies is based on shape of reflections; dip of reflections; relationship
between reflections, reflection continuity and reflection amplitude (Neal 1994); these

29

criteria can be significantly influenced by the processing steps. Continuous reflection can
appear discontinuous by applying inappropriate gains during data processing and
reflection dips can result from inappropriate processing algorithms; hence, the
importance of proper signal processing steps and outcrop data as ground truth to verify
the fidelity of radar data. This also underscores the importance of documenting data
acquisition and processing steps in GPR publications to allow comparison between
studies.
Commonly applied processing steps on GPR data include:

2. 6.1 Topographic Correction


The collection of topographic data is an essential component of GPR survey as radar data
collected in the field does not take into account topographic variation along a survey line.
This can be corrected by moving traces up and down by an appropriate two-way-time
relative to a common datum, based on knowledge of the velocity, and, therefore, depth
profile of the uppermost part of the radar profile. In order to do this, survey line
topography must be adequately characterized. Topographic surveys are typically
performed using a variety of instruments including optical levels, total station, laser
levels, and differential GPS as these have the required vertical resolution and allow
relatively rapid data collection. Spatial sampling in such surveys should ensure that all
significant breaks in slope are accounted for. While lower frequency antenna can tolerate
minor elevation changes, high frequency antennas (above 200 MHz) are very sensitive to
topographic variation (elevation changes over 1 metre) along the survey line and often
create artificial structures in the radar profiles if not corrected (Figure 8). Survey

30

equipment was not available for this study; hence outcrops with flat surfaces were chosen
for this study to avoid the need for elevation correction of the recorded GPR profiles.

Figure 8: Artificial structure on downstream accretion element created by elevation


variation along the profile line. (Profile recorded on Castlegate sandstone outcrop, Tusher
Canyon, Utah).

2.6.2 Dewow
Wow noise is peculiar to ground penetrating radar and it is as a result of the close
proximity of receiver to transmitter. Dewow is one of GPR processing basic temporal
filtering steps aimed at removing very low frequency components from the data (Figure
9). Very low frequency components of GPR data are associated with either inductive
phenomena or possible instrumentation dynamic range limitations. This process has
historically been done using analog filters in hardware but with the advent of true digital
data acquisition, it has also become a data processing step (Gerlitz et al, 1993). Dewow
filter acts on each trace independently by calculating a mean value of each trace and the
running mean subtracted from the central point.

31

Figure 9 (A) Display of a single data trace (left) and data section (right) with the low
frequency wow component masking the real data. (B) Display of a single data trace (left)
and data section (right) where the Wow seen in (A) above has been removed with the
dewow high pass filter. (Image taken from Annan, 2004).

2. 6.3 Backgound Removal


One of the most common operations specifically applied to GPR data is the use of
background removal. Background removal is a form of spatial filtering; most often, this
takes the form of a high pass filter or a filter that takes the mean of all traces in a section
and substracts it from each trace.

32

Figure 10: (A) GPR profile before background removal (B) GPR profile after
background removal. (Image taken from http://www.malags.com/Downloads/ProductBrochures.aspx; accessed, August 6, 2010).
In situations where antenna ringing, transmitter reverberation and time synchronous
system artifacts appear, it is very effective in allowing subtle weaker signals, which are
lost to become visible in a processed section. Background filter can eliminate temporally
consistent noise from the whole profile and therefore possibly make real signals
previously covered by this noise visible and suppress horizontally coherent energy

33

(Figure 10 A and B). It is quite effective in relatively lossy materials (conductive


lithologies such as wet or shaly rocks).

2. 6.4 Deconvolution
Deconvolution is aimed at removing effects of a previous filtering operation (Yilmaz,
1987, 2001; Kearey and Brooks, 1991). In both seismics and radar signals, deconvolution
attempts to remove filtering effects resulting from propagation of a source wavelet
through a layered earth, and the recording system response. The intended effect of the
deconvolution process is to shorten pulse length and, therefore, improve vertical
resolution (Figure 11).
It is often tempting to apply seismic processing steps like deconvolution to GPR because
of their kinematic similarity; it is however important to note that deconvolution of GPR
data is not straightforward and rarely yields impressive results (Fowler and Still, 1977;
Payan and Kunt, 1982; LaFleche et al., 1991; Maijala, 1992; Todoeschuck et al., 1992;
Fisher et al., 1996; Arcone et al., 1998). The primary reason for this is that the radar pulse
is often as short and compressed as can be achieved for the given bandwidth and signalto-noise conditions. Another important factor is that some of the more standard
deconvolution procedures have underlying assumptions required for wavelet estimation
such as minimum phase and stationarity, which oftentimes are not appropriate for GPR
data (Annan, 1999). The rapid attenuation of GPR signal amplitude means that
deconvolution artifacts may mask weaker deeper events if time gain is not applied first
and the non-linear nature of time gain may substantially alter wavelet character if gain is
applied before deconvolution. Hence, deconvolution can be difficult to apply and might
yield marginal enhancement in resolution.

34

Figure 11: (A) GPR profile before deconvolution (B) GPR profile after deconvolution.
(Image taken from http://www.malags.com/Downloads/Product-Brochures.aspx accessed on August 6, 2010).
Few instances where deconvolution has proven beneficial occurred when extraneous
reverberation or system reverberation have been involved. In such instances,
deconvolution provides substantial pulse compression benefits (Turner, 1994; Xia et al.,
2003)

35

2. 6.5 Time Gain


Radar signals are prone to rapid signal attenuation as they propagate into the ground.
Signals from greater depths are often very weak, such that simultaneous display of this
information with signals from a shallower depth requires preconditioning for visual
analysis and display. When the amplitude of display is optimal for shallow depth signals,
events from greater depths may be invisible or indiscernible. Gains are required because
the amplitude of a reflected signal decreases with time and depth due to attenuation,
geometrical spreading, partial reflection and scattering (Davis and Annan 1989; Reynolds
1997). "Time gain" applied to radar data attempts to equalize amplitudes by applying
some sort of time-dependent gain function, which compensates for the rapid fall off in
radar signals from deeper reflectors. A variety of gain functions can be applied to radar
signals (e.g. constant, linear and exponential gains); it is however important to understand
that the choice of gain adopted might alter the amplitude fidelity of the signals in the
data.
Geological radar surveys are especially susceptible to signal attenuation in mud-rich
environments; hence, for stratigraphic horizon continuity, displaying all the information
irrespective of amplitude fidelity might be important. In this case, manual gain or a
continuously adaptive gain such as AGC (automatic gain control) is often used. With
AGC gain, each data trace is processed such that the average signal is computed over a
time window and then the data point at the centre of the window is amplified (or
attenuated) by the ratio of the desired output value to the average signal amplitude.
Systematic gains, such as Spherical and Exponential Compensation (SEC) gain, attempt
to emulate the variation of signal amplitude as it propagates in the ground. Whichever

36

gain function is chosen, it should be selected based on the specific goal of the survey and
data processing requirement; the objective should be to modify the data while retaining
fidelity without introducing artifacts.
A

Queenston Shale
Whirlpool channel sand

Figure 12: (A) Un-gained Whirlpool Sandstone GPR Profile. (B) Manually-gained
profile at the Whirlpool Sandstone outcrop at Artpark, USA.
It is important to document the type of gain function applied to radar data while
discussing radar facies as radar reflector character is dependent on the signal processing
steps applied to the data (relative amplitudes and/or phase relationships are changed); for
instance, reflectors described as discontinuous low amplitude facies may show up as
continuous reflectors where stronger gains are applied (Figure 12).

37

2.6.6 Migration
Migration relocates reflections to their true spatial position based on the velocity
spectrum of radar signals through the rock layers to produce a real structure map of
subsurface features. Migration algorithms do this by removing diffractions, distortions,
dip displacements and out-of-line reflections resulting from the fact that radar antenna
radiate and receive electromagnetic energy in a complex 3- D cone.
The goal of migration is to make the reflection profile look like the geological structure
in the plane of the survey. It attempts to correctly position subsurface reflection events
(Hatton et al., 1986). However, due to various uncertainties, a significantly improved but
still imperfect image is usually achieved (Figure 13). Such improvements are quite
beneficial in sedimentological studies, where the nature and form of stratigraphic units
and primary sedimentary structure is of utmost importance.
A

Figure 13: (A) Unmigrated GPR Profile (B) GPR profile migrated by diffraction
hyperbola collapse. (Image taken from Anan, 2004).

38

Migration is generally used for improving section resolution and developing more
spatially realistic images of the subsurface and is, perhaps, the most controversial of the
GPR processing techniques. Migration techniques, like deconvolution, were originally
developed for the seismic industry where they are considered as vital for even basic
interpretations. Unfortunately, migration tends to be less successful with GPR, and
although it can be used in relatively homogeneous environments, it is not so good with
complex, heterogeneous environments (lithologies) with variable radar velocity typical of
clastic many rocks. Despite this, migration is still a vital georadar processing step and
classical techniques have been applied successfully to a range of different applications.
Examples include reverse time migration (Sun and Young, 1995; Meats, 1996), FK
migration (Fisher et al., 1994; Pettinelli et al., 1994; Pipan et al., 1996; Yu et al., 1996;
Hayakawa and Kawanaka, 1998) and Kirchhoff migration (in Moran et al., 1998).
Specific GPR-based methods have been developed to overcome some of the limitations
in the seismic data migration algorithms. Examples include matched filter migration
(Leuschen and Plumb, 2000); Kirchhoff migration modified for radiation patterns and
interface reflection polarisation (Moran et al., 1998; Van Gestel and Stoffa, 2000);
eccentricity migration for pipe hyperbola collapsing (Christian and Klaus-Peter, 1994),
3D-based vector and topographic migration (Lehman and Green, 2000; Heincke et al.,
2006; Streich et al., 2007) and frequency domain migration for lossy soils (Di and Wang,
2004; Sena et al., 2006; Oden et al., 2007). These new methods are yet to be incorporated
into mainstream GPR processing packages, although most do contain some form of
relatively sophisticated (if classical) migration algorithm. The most common are
diffraction stack migration, FK migration (or Stolt migration), Kirchhoff migration and

39

wave equation or finite-difference modelling migration. These can be applied to 2D GPR


profiles or across 3D volumes of data. The specific theories and quantitative details of
how each method works are discussed in Yilmaz (2001). Success of GPR data migration
is dependent on accurate estimation of the velocity profile through the different layers of
the rocks which is often obtained by Common Mid-Point (CMP) surveys. Unfortunately,
increased surveying time/equipment cost required for CMP surveys often limit the
number of CMP surveys conducted, hence, sparse velocity information is usually
available. Other less reliable methods such as average radar velocity associated with the
lithologies at the outcrop as well as hyperbola matching are often used but these
determine only the average velocity to isolated horizons or spatially restricted zones.
While migration is a routine seismic processing step, GPR data obtained for stratigraphic
studies are rarely migrated; perhaps this is because migration requires a thorough
knowledge of the velocity structure in the imaged formation (which is often not available
in most GPR single fold common offset surveys) and susceptibility to errors in the
velocity structure which introduces artifacts into migrated GPR section. Data obtained
with shielded antennas are usually less prone to diffractions than data obtained with
unshielded antennas; shielded antennas might therefore be more advisable in locations or
outcrops susceptible to significant signal scattering; they are however not designed for
CMP surveys needed to obtain velocity profile required for data migration.

2.6.7 Advanced Data Processing


Advanced data processing techniques address the types of processing, which require data
processing expertise to be applied and which often results in data which could be
significantly different from the raw information fed into the processing unit. Application

40

of such processing techniques well-known in seismic processing operations such as trace


attribute analysis, FK filtering, selective muting, normal move out correction, dip
filtering, and velocity semblance analysis is still in its infancy in GPR data processing;
although there has been an increasing application of the more GPR-specific operations
such as multiple frequency antenna mixing and polarization mixing (Tillard and Dubois,
1992). With careful data collection and effective ground coupling of radar antennas,
however, most GPR profiles hardly require advanced post-processing to enable basic
interpretation of sedimentological features.

2. 6.8 Time - Depth Conversion


GPR data are usually displayed as distance-two way time (in nanoseconds) profiles;
conversion to distance-depth profiles is more easily done with data obtained in CMP
(Common-Midpoint) mode. Reliable subsurface-radar-wave velocity information is
important for converting two-way-travel time (TWT) to an accurate estimation of depth.
The standard approach to velocity estimation in many studies is the CMP survey as it is
entirely non-invasive and can be supplemented by subsequent ground truthing (Annan
and Davis, 1976; Beres and Haeni, 1991; Tillard and Dubois, 1995; Greaves et al., 1996;
Van Overmeeren et al., 1997). Because most shielded antennas are not designed for CMP
surveys and significant data acquisition time associated with CMP surveys, most GPR
profiles are obtained in common offset mode. In such cases, velocity can be roughly
estimated by measuring TWT to a bounding surface or bedding of known depth where
major bounding surfaces in outcrop can be directly correlated to radar reflectors. The
travel time for a pulse to travel from the transmitter to a reflector and up to the receiver,
the two-way travel time, t, as a function of the velocity at which the electromagnetic

41

pulse is propagated through the host medium, Vm, and the depth to the bounding surface
d, are related by the Equation
t= 2d/Vm

(2.4)

Velocity for depth conversion of the radar profiles recorded in this study was estimated
using this relationship. Although rarely, velocity required for time-depth conversion is
also estimated by direct laboratory measurements of dielectric permittivity on field
samples; measuring travel time between two wells using borehole radar; and
transillumination surveys between two parallel exposures (Annan and Davis, 1976; Topp
et al., 1980; Fisher et al., 1992a; Greaves et al., 1996; Reynolds, 1997; Binley et al.,
2001; Hammon et al., 2002; Tronicke et al., 2002a).

2. 7 Visualization and Display


Ground-penetrating radar has been used for geological imaging since the 1980s and has
since developed into a valuable tool for stratigraphic studies. Most studies utilize
common-offset, 2-D radar reflection profiles to display stratigraphic information obtained
in modern environments and outcrop. Where there is significant lateral variability in
internal structure, pseudo-3D or true 3D surveys may be desirable. Pseudo-3D surveys
involve collecting data on regular or irregular survey grids, usually in two mutually
perpendicular directions, and often displaying results as fence diagrams (for example,
Bristow, 1995; Aigner et al., 1996; Bristow et al., 1996, 1999, 2000b; Roberts et al.,
1996; Asprion and Aigner, 1997, 1999; Leclerc and Hickin, 1997;; Pedley et al., 2000;
Neal and Roberts, 2001; Russell et al., 2001; Holden et al., 2002; Skelly et al., 2003). In
the 1990s, time-slicing, a mechanism for producing horizontal plan maps routine at that

42

time in three dimensional seismic imaging revolutionized the way GPR results are
displayed and interpreted (Goodman et al., 1995), and recent GPR three-dimensional
imaging produced exciting results for visualization and enhanced interpretation (Conyers
et al., 1997; Jol et al, 2003; Leckebusch, 2003).
While full resolution 3D GPR surveys are generally more informative than single
profiles, it requires significantly increased data acquisition time. Hence, they are seldom
used in geological studies. Data processing for three dimensional GPR surveys is more
time consuming partly because of greater data volumes and complexity, but also due to
lack of efficient commercially available software for three dimensional GPR data
processing; these explain why there are few 3D GPR outcrop studies in the GPR
literature. Not all studies require 3D surveys; staggered GPR lines parallel and orthogonal
to the outcrop face displayed as a fence diagram often yield sufficient architectural details
for stratigraphic interpretation.

2. 8 Causes of GPR Reflections


Although GPR images appear similar to seismic reflections generated across surfaces due
to changes in acoustic impedance in the subsurface, GPR reflections are generated by
changes in electromagnetic impedance due to variation in the electrical and magnetic
properties of the lithologies that an electric pulse transmitted from the antenna into the
ground passes through as it makes its way through lithologies with different electrical and
magnetic properties. While the theoretical basis of radar signal propagation and reflection
have been exhaustively discussed by many authors (Daniels, 2004; Baker et al., 2007;
Rubin and Hubbard, 2005), the actual causes of GPR reflections in sediments and rocks
and their implications for interpretation of GPR data from modern and ancient sediments

43

have been loosely defined. Few studies that have investigated the causes of GPR
reflection in sediments (Van Dam and Schlager, 2000; Van Dam, 2001) discovered that
propagation of electromagnetic signals is primarily dependent on the electrical
conductivity and magnetic permeability of the medium that the signals pass through.
Electrical conductivity is the ability of the material to transmit electric charges under the
influence of an applied field. In rocks and sediment, this often depends on the dielectric
permittivity associated with variations in water content and ionic concentrations. Water
in sediment pore space normally contains ions, and the electrical conductivity associated
with ion mobility is the dominant factor in determining bulk-material electrical
conductivity. Since water is invariably present in the pore space of sediments and rocks,
it has a dominant effect on electrical properties. Magnetic permeability is the degree of
magnetization of a material that responds linearly to an applied magnetic field; this is
significant in rocks or sediments rich in ferromagnetic minerals. However, in most
sedimentary rocks, the amount of ferromagnetic minerals is minimal and hardly
contributes significantly to generation of radar reflections. Hence, in most GPR surveys
on sediments and rocks, the dominant control on radar transmission and reflection is
electrical conductivity governed by water content and the sediment static conductivity.
The variation in sediment water content often relates to both changes in sediment
porosity and permeability, which in turn is dependent on changes in grain size and fabric
often associated with lamination, cross-bedding and bounding surfaces. Rock property
research in the petroleum industry in the seventies and eighties geared at understanding
controls on permeability distribution in clastic reservoirs suggested that fluid flow in
cross-bedded sandstone is controlled by porosity and permeability contrasts of the cross-

44

bed laminae and the associated bottom-set layer (Pryor, 1973; Beard and Weyl, 1973,
Jordan and Pryor, 1987). Emmet et al (1971) observed that porosity and permeability
parallel to the cross-bed laminae are higher than perpendicular to the laminae. Such
porosity and permeability changes have direct control on water content, which determines
electrical conductivity changes that generate radar reflections (Figure 14).
A

5m

High porosity cross-bed

Low porosity
bounding surface

Figure 14: (A) Navajo Sandstone outcrop at Zion National Park, Utah (B) Behind
outcrop GPR profile using 100 MHz antenna across the Navajo Sandstone outcrop in (A).
(Image taken from Jol. et al., 2003).

45

Sedimentary bedding is a product of changes in sediment composition and changes in the


size, shape, orientation and packing of grains (Collinson and Thompson, 1989) which
results in corresponding changes in porosity (Figure 15).

Higher porosity and permeability

Lower porosity and


permeabilitty

Figure 15: Bedding in sediments and sedimentary rocks and associated porosity resulting
from changes in composition, size, shape, orientation and packing of sediment grains.
(Image modified from Neal, 1994).
Van Dam et al., (2002a) reported that goethite iron-oxide precipitates, occurring either in
bands or irregular layers, were responsible for significant reflections in the aeolian sands
they studied. This was due to the higher water retention capacity of goethite with respect
to the host quartz sand, which resulted in higher dielectric permittivity.
While most radar reflections are associated with electrical conductivity associated with
porosity changes across primary bedding, bounding surfaces and lithological breaks,
changes in magnetic permeability (although rarely) may generate radar reflections.

46

Textural differences between laminae of differing origins may also control cementation
patterns and, hence, post depositional diagenesis, which may influence radar reflection
patterns in consolidated sediments.

2. 9 GPR Interpretation Methodology and Radar Stratigraphy


Because seismic reflection and GPR data are analogous in terms of wave propagation
kinematics (Ursin, 1983; Carcione and Cavallini, 1995) as well as reflection and
refraction responses to subsurface discontinuities (McCann et al., 1988; Fisher et al.,
1992a) many of the broad assumptions that underpin processing and interpretation of
seismic reflection data (Sangree and Widmier, 1979; Yilmaz, 1987, 2001) are often
routinely applied to GPR data. However, it must be borne in mind that radar images are
different from seismic images in the following respect:
1. Signal penetration and reflection amplitude are controlled by water saturation,
clay content and magnetic property of the sediment.
2. Radar signals are more prone to scattering making diffraction noise more
prominent in radar profiles.
3. The investigation depth is much less.
4. Resolution is in sub-metre scale.
5. Radar surveys require direct contact between the pulse-transmitting antenna and
the medium which the signals are transmitted through (rocks or water/moist
sediment for surveys conducted in modern fluvial environments); this limits the
size of most GPR surveys and often results in much smaller coverage than seismic
surveys.

47

From the interpretation standpoint, the basic assumption in both techniques is that, at the
resolution of the survey and after appropriate data processing, reflection profiles will
contain accurate information regarding the external geometry and internal structure of a
sediment body. This implies that the form and orientation of bedding and sedimentary
structures in the plane of the survey will be adequately represented by recorded
reflections, and non-geological reflections can be readily identified and removed by data
processing, or by simply discounting them from the interpretation. While this assumption
holds in many instances, accurate interpretation of GPR data is dependent on the nature
and appropriateness of data processing steps undertaken, the interpretation techniques
employed, and the overall understanding and experience of the interpreter with respect to
GPR; hence, care must be taken to ensure that GPR profiles are not treated as geological
cross sections but interpreted in the context of the local geology and ground-truthed with
outcrop data.
Based on the similarities between GPR and seismic reflection, the concept of seismic
stratigraphy was adapted for interpretation of GPR profiles soon after the realization that
GPR could provide useful data for stratigraphic and sedimentological studies (Baker,
1991; Beres and Haeni, 1991); Jol and Smith (1991) coined the term radar stratigraphy
for this new interpretation technique. Gawthorpe et al (1993) defined radar facies based
on the terminology used to describe seismic facies (Mitchum et al., 1977; Brown and
Fisher 1980) as three-dimensional packages that represent particular combinations of
physical properties such as lithology, stratification style and fluid content, and may be
used to interpret depositional processes and environments. They also defined radar
sequence boundaries as systematic reflector terminations indicating non-depositional and

48

erosional hiatuses despite acknowledging that while seismic sequences can be mapped
on at least a basin scale and reflect units that form over hundreds of thousands to millions
of years, radar sequences may be limited to particular depositional environments (e.g. a
point bar in a fluvial channel belt) and may form over a much shorter time scale of tens to
thousands of years. Although, this interpretation methodology has been widely
embraced over the years, it is important to underscore the fact that GPR resolution is an
order of magnitude higher than conventional seismic resolution and coverage of GPR
surveys rarely has the capability to image stratigraphic sequences and sequence
boundaries; it is more appropriate for imaging of meso-scale depositional features or
outcrop scale architectural elements. Although decimeter-scale resolution has been
reported in high resolution seismic surveys, the technology is rarely used for outcrop
studies; hence, ground penetrating radar remains the choice technology for high
resolution outcrop imaging.
To avoid interpretive connotations in describing radar reflections, it is recommended that
unambiguous descriptive criteria such as geometry; dip of reflections, relationship
between reflections be used to describe radar reflections rather than terms with
interpretive connotations like, sequence boundaries, systems tract, etc used in seismic
stratigraphy. Interpretations based solely on processing-dependent attributes such as
reflection continuity and amplitude should also be avoided as these are dependent on the
data processing algorithms or gain magnitude applied to the data and are hardly
comparable between different GPR studies. The potential for misinterpretation of radar
data based on reflection continuity is illustrated below (Figure 16 A and B)

49

Reflection packages a, b and c (Figure 16A) could be interpreted as different architectural


elements based on reflection continuity but the reflectors are seen as continuous in the
better gained data (Figure 16B).

A
0

Depth (m)

B
0

Depth (m)
6

Figure 16: (A) Poorly-gained radar profile giving the impression of facies with low
continuity (B) Same data in (A) with gains applied to display all reflectors. GPR data
acquired at Whirlpool Sandstone.
Also, while in most cases, radar reflections parallel bedding and bounding surfaces, care
must be taken to ensure that reflections generated by the water table and diagenetic
horizons, diffractions generated by isolated reflector points, out-of-line reflections,
reflections generated by faults and joints, surface reflections, and ambient and systematic
noise are not interpreted as primary sedimentary bedding or bounding surfaces.

50

2.10 Radar Facies


Analysis of 150 published studies in the last thirty years on stratigraphic imaging with
ground penetrating radar reveals an obvious bias towards studies of modern and recent
deposits (Figure 17 and Table A3, Appendix 3). Much of the GPR effort on rocks has
been concentrated on carbonates as the low electrical conductivity of carbonate rocks
make them highly conducive to georadar imaging.

Number of published GPR studies

Number of published GPR studies

Figure 17: (A) Analyses of 150 published GPR studies on sediments showing bias for
Quaternary sediments (B) Analysis of ancient deposit studies in (A) showing
comparatively more GPR studies on carbonate rocks.
51

Observations from analyses of published studies also show that highest radar signal
penetration is observed in carbonate rocks due to their high resistivity. This study
however focuses on the application of GPR imaging technology to the studies of ancient
clastic depositional units.
There is currently no formal methodology for documenting radar profiles and radar
facies; this in addition to different processing and display formats adopted by various
authors make comparison of radar facies rather challenging; but an attempt is made
below to summarize observations and describe radar facies commonly observed in a
variety of depositional environments. Radar facies are defined by: external reflection,
geometry, dip of reflections and relationship between reflections. Due to the paucity of
GPR studies on ancient sediments, most of the radar facies discussed in this section are
from modern sedimentary environments and Quaternary deposits.
Since the inception of its use as a tool for stratigraphic and sedimentological
interpretation, much of the emphasis of GPR studies has been on characterization and
interpretation of radar facies, with radar surfaces and radar packages typically not being
defined. This overemphasis on radar facies analysis, rather than true radar stratigraphy,
led to the common misconception that any radar reflection pattern constituted radar
facies. As a result, use of radar facies to describe reflections not related to primary
sedimentary structure and stratigraphy became rife in the GPR literature. Terms such as
water-table radar facies and hyperbolic or diffraction radar facies can be found
throughout the GPR literature. In addition, radar surfaces have also been described as
radar facies. The term palaeosol radar facies has been used on several occasions, despite
the fact that the facies is defined by a single reflection (it is a radar surface at the

52

resolution of the radar profile). These examples are in violation of the correct use of the
term radar facies and, therefore, the true principles of radar stratigraphy. Further
confusion to the interpretation of radar profiles has also been caused by use of
interpretive facies names, such as channel-fill facies and overbank facies. If the
context is not made clear, use of interpretive facies names has the danger of suggesting
that particular depositional environments or sub-environments are characterized only by
certain radar facies which is not true based on the findings from published studies and
GPR outcrop data obtained in this study. Radar facies should therefore be described
without interpretive connotations and an uninterpreted version of the radar facies
provided to provide readers an unprejudiced data and basis for interpretation.
The value of radar facies interpretation is demonstrated in all the case studies for this
research especially at the Shinarump Conglomerate (Figure 21 A), Navajo Sandstone
(Figure 21 B) and Whirlpool Sandstone (Figure 22A). Common radar reflection patterns
(facies) typically observed from a variety of sedimentary environments are discussed
below:

2.10.1 Concave upward/trough-shaped reflectors


Concave-upward ground radar reflectors are observed as valley fill (Ilya, 2006), cut and
fill (Heinz et al, 2003) channel (Figure 18), chute, scour, bar top hollows (Best et al,
2006) and trough cross-beds (Bristow et al, 1999). Documented concave upward GPR
reflectors range from sub-metre length to a few kilometers, although published kilometrescale GPR studies are very few (Mller and Anthony, 2003; Hickin et al., 2007). Submetre scale concave-upward reflectors are usually trough cross beds or minor scours.

53

Channels are often several metres to hundreds of metres in width and may have simple to
complex fill.

CH

CH

Figure 18: GPR Image of fluvial channels from Dunvegan Formation outcrop (Pink
Mountain, British Columbia).
Valleys are typically hundreds of metres to kilometre-scale in width, they contain
channels and channel belts and are rarely defined in GPR profiles.
While the various concave-upward radar facies are sometimes geometrically-identical, it
might be difficult to correctly interpret them; they can however be discriminated by their
scale, internal geometry and facies associations. This underlines the importance of
documenting the horizontal and vertical scale of GPR profiles. The geometry of
reflectors often depends on the orientation of the GPR line in relation to the channel axis
(for bedload deposits). GPR profiles along the axis of a channel deposit usually reveal
different reflector geometry compared to profiles perpendicular to the channel axis; hence
it is important to document the orientation of GPR profile from which the radar facies are
obtained.

54

Figure 19: (A) Radar facies from Fraser and Squamish River showing concave upward
reflectors. (Image taken from Wooldridge and Hickins, 2005) (B) Concave upward radar
facies from Niobrara River, Nebraska. (Image taken from Bristow, 1999) Red arrow
highlights major channel.

55

2.10.2 Inclined Reflectors


Inclined radar reflectors are observed in a variety of environments; they range from
planar cross-beds, lateral accretion and downstream accretion in fluvial deposit to
Inclined Heterolithic Stratification (IHS) in tidal bars. Inclined radar facies are also
observed in eolian deposits as giant cross-beds and also in deltaic clinoforms (Figure 20).
They are often vital paleocurrent indicators; hence the orientation of the GPR profiles
where they are observed as well as the orientation with respect to channel trend or
regional paleocurrent direction should be carefully documented. They can be
discriminated by the vertical and horizontal scale of the reflectors as well as the
association with contiguous facies, ground-truthed with outcrop observations or sediment
trenches. Examples from modern and ancient sediments are given below:

2.10.3 Reflection-free facies


Reflection-free configuration often indicates one of the following:

massive homogenous lithological units with no contrasting dielectric properties

the presence of highly conductive dissolved minerals in groundwater

magnetic sediment

or the presence of sediments containing high clay content that rapidly attenuates
transmitted electromagnetic signal.

56

Figure 20: (A) Inclined Radar facies from Fraser and Squamish River (Image taken
from Wooldridge and Hickins, 2005) (B) Downstream accretion macroform from
Castlegate sandstone, Utah (C) GPR profile at the aeolian Navajo Sandstone from Zion
National Park, Utah revealing a set of cross stratification. (Image taken from Jol et al.,
2003).
The susceptibility of radar signals to rapid attenuation in argillaceous sediments is vital
for interpreting lithology from radar profiles. Chanel cutbank, overbank and floodplain

57

deposits can be clearly interpreted in fluvial deposits and interdune deposit discriminated
from dry phase eolian dunes based on reflection geometry and amplitude (Figure 21)

Reflection-free facies
Shinarump Conglomerate
Moenkopi Shale

Muddy interdune deposits

Low

Reflection-free facies
High

Amplitude Scale

Figure 21: (A) Example of reflection-free radar facies- Shinarump Sandstone (Hurricane
Mesa, Utah) (B) Radar facies from Navajo Sandstone (Moab, Utah)

2.10.4 Horizontal Reflectors


These facies is characterized by continuous, sub-horizontal to horizontal reflectors.
They are observed from diverse depositional environments: fluvial overbank levees
(Figure 22 A), vertically aggraded channel deposit (Figure 22 B), sheetflood deposit,

58

floodplain deposits and lacustrine sediments. They are often recognized by their
reflection amplitudes and facies association; for instance, muddy horizontal overbank
deposits are usually high to medium amplitude reflectors with low amplitude at the
base as observed at the Whirlpool Sandstone Figure 22A) while sand-rich intra channel
horizontal reflectors are usually characterized by high amplitude sometimes underlain
by a thin layer of low amplitude reflectors indicating shaly channel base breccia or mud
chips (Figure 22 A and B). Mud beds typically have low amplitude thin horizontal
reflectors with indistinguishable reflectors at the base.

A
Overbank deposit
Queenston Shale
Channel margin

Sandy Bedform

Figure 22: (A) Horizontal radar facies overbank deposit from Whirlpool Sandstone,
Artpark, New York (B) Horizontal reflectors from Shinarump sheet conglomerate, Utah
interpreted as upper plane bed flood deposits.

2.10.5 Hyberbolic Reflections


Hyperbolic radar reflections were once thought to be diagnostic of glacial push moraine
deposits (Van Overmeeren, 1998); observations from several published GPR studies
59

(Arcone, 1998; Snchal et al., 2000; Ekes and Friele 2003; Van Dam, et al., 2003)
however reveal that these reflections are due to diffractions of radar signals and thus, are
not diagnostic of any depositional environment. Due to the acute susceptibility of radar
signals to scattering; they are common in un-migrated radar profiles especially those
obtained with unshielded antenna. Diffraction noise is easily recognized in GPR sections
as hyperbolas (Figure 23).
Unfortunately, migration requires a thorough knowledge of the velocity structure in the
rock or sediment being imaged which is not often easily determinable; hence, uninformed
signal processing steps aimed at removing diffraction hyperbolas might introduce
artefacts into the radar profile (section 2.6.6).

Figure 23: Example of hyperbolic reflection (HP) from an un-migrated GPR profile.
(Image taken from Woodward et al., 2003). Red arrow points at hyperbolic reflections.
In the early days of GPR imaging on sediments, observation of distinct reflection patterns
from a variety of modern sedimentary environments led to the assumption that certain
reflection configurations are diagnostic of specific sedimentary environments (Table 2).

60

Table 2: Radar facies elements of different sedimentary depositional environments.


(Table obtained from Van Overmeeren, 1998)

Based on this surmise, Van Overmeeren (1998) suggested a compilation of radar facies
atlas from a variety of depositional environments as guide for facies interpretation and
identification of depositional environment from sediments (Figure 24). Although there
have since been many published studies involving GPR imaging of sediments, many of
the early ideologies still persist.

61

Figure 24: Radar facies chart of characteristic reflection patterns from various
sedimentary environments. (Image taken from Van Overmeeren, 1998).
Analysis of radar facies observed in this study and from other published GPR studies
evaluated confirms that no reflection configuration is diagnostic of a specific
depositional environment. For instance, concave upward radar reflections could be
associated with confluence scours, main channel fill, small cross-bar channels adjacent
to unit bars, scours upstream of obstacles such as logs or ice blocks or bases of trough
62

cross strata. Hyperbolic reflectors are observed in most un-migrated or poorly-migrated


radar profiles irrespective of the sedimentary environment (modern or ancient) where the
radar profiles are obtained. Also, while the link between sedimentary architecture and
radar reflection patterns implies that different radar facies represent spatially varying
geologic settings in the subsurface, Jol and Bristow (2003) cautioned that many different
geologic scenarios can produce similar reflection patterns in a GPR image. Similarly, the
dependence of GPR resolution on frequency means that a single depositional
environment could produce many different textures depending on antenna frequency
used for the survey (Huggenberger and Pugin, 1994; Jol, 1995; Jol et al., 2003). While
certain reflection patterns are more common in certain environments (e.g large scale
high amplitude dipping reflectors characteristic of eolian dunes) similar reflection
patterns can be observed in different depositional environments.
Hence, without additional information or data to provide a geologic context for the field
site, the interpretation of geologic facies directly from a GPR image or radar facies
interpretation should be avoided; rather, interpretation of radar reflections should be
made in the context of their three dimensional geometry and an understanding of adjacent
strata ground-truthed by outcrop observations. The scale and vertical exaggeration of the
radar profile should also be considered when making sedimentological interpretation
from radar reflections (Bridge and Lunt, 2006).
In summary, many studies have interpreted sedimentary facies and facies associations
and have determined depositional environment and formative processes from radar
reflection profiles ground truthed with sedimentological data from trenches, cores or
outcrop (e.g. Jol and Smith, 1991; Smith and Jol, 1992; Gawthorpe et al., 1993;

63

Huggenberger, 1993; Bristow, 1995; Beres et al., 1999; Neal and Roberts, 2000, 2001;
Corbeanu et al., 2001; Heinz, 2001; Neal et al., 2001, Russell et al., 2001; Szerbiak et al.,
2001; Hornung and Aigner, 2002; ONeal and McGeary, 2002; Heinz and Aigner, 2003;
Skelly et al., 2003). While these studies confirm the usefulness of GPR imaging as a
powerful technique for imaging clastic sedimentary rocks, the idea that radar facies
elements used in interpretation of radar reflection patterns are characteristic of certain
sedimentary depositional environments should be applied with caution while interpreting
outcrop data as many radar facies are not diagnostic of specific ancient depositional
environments. Unlike seismic attributes, extraction of radar reflection attributes (e.g
coherence, azimuth and dip) as proxy data for sedimentary architecture is still in its
infancy; such attributes can be easily distorted by the signal algorithms employed in data
processing as well as misinterpreted due to moisture content of the imaged lithologies
and should be applied with much care. For most stratigraphic studies, it seems preferable
to adopt processing steps that highlight bedding continuity and reveal all reflectors
irrespective of amplitude fidelity as this often provides a more reliable basis of
interpretation. Reflection amplitude can also yield vital clues about lithology and guide
architectural interpretation (Figure 22); in this case, amplitude distorting algorithms
should be avoided and amplitude comparison between data subjected to different
processing algorithms should be done with caution or avoided.

64

Chapter 3
Case Study 1 - Whirlpool Formation

3.1 Introduction
There have been significant advances in our understanding of flow dynamics and
associated fluvial deposits in the last decade and much of the result is credited to the use
of ground penetrating radar on sediments ground-truthed with sedimentological data from
cores and trenches. Most of these studies focused on modern fluvial deposits as earlier
attempts to apply GPR imaging technology to ancient fluvial deposits were largely
unsuccessful due to poor data acquisition practice, fractures and diagenetic signal
overprint (Bristow and Jol, 2003).
Despite several years of GPR imaging of modern rivers and Quaternary fluvial sediments
with impressive results, there have been very few applications of GPR imaging
technology to ancient fluvial deposits due to operational difficulty of conducting GPR
surveys on outcrops. This study however, shows that with careful survey planning and
efficient data collection and processing, GPR imaging technology can be successfully
applied to the study of ancient fluvial deposits in the quest for refining current fluvial
architectural models and as analog data for subsurface reservoirs.
The lower unit of the Whirlpool Sandstone at Artpark (New York, USA) provides an
opportunity for GPR imaging integrated with outcrop architectural element analysis in an
attempt to evaluate the effectiveness of GPR imaging on outcrops as well as to refine
existing interpretation via three-dimensional architectural study of the Whirlpool
Sandstone. The Whirlpool Sandstone outcrop at Artpark was chosen as a case study
65

because of its distinct architecture (apparently clean, trough-cross-bedded fine to


medium-grained texture), flat top, reasonably extensive exposure, absence of thick
overburden (which prevents radar signals from reaching the target formation) and
moderate stratigraphic thickness (this makes it feasible to image the entire formation).
Although lithological contrast is often suggested as a pre-requisite for successful GPR
survey, the Whirlpool Sandstone provides a test of the effectiveness of GPR imaging on
sandstone with considerable homogeneity and minimal obvious lithological contrast.

3.1.1 Study Objectives


The objectives of this study are:

to demonstrate that sedimentological and architectural interpretation can be made


from GPR profiles by correlating outcrop observation with GPR images.

to document radar facies from the Whirlpool Sandstone; currently there are few
documented radar facies from ancient fluvial deposits. Radar facies from the
Whirlpool Sandstone will provide an opportunity to compare radar reflection
patterns with lithofacies and architectural elements observed in similar outcrops.

to re-evaluate current interpretation of the Whirlpool Sandstone at the study


location in the light of integrated outcrop and GPR data

extend interpretation of the Whirlpool Sandstone outcrop at Artpark based on


two-dimensional architectural element analysis into three dimensions by
integrating outcrop with GPR data.

to test the reliability of three-dimensional GPR imaging over vertical profiles and
architectural element analysis (from earlier studies of the Whirlpool Sandstone at
Artpark) in unravelling depositional processes from outcrops
66

3.1. 2 Whirlpool Sandstone - Geological Overview


Overlying the Queenston Shale is the Whirlpool Sandstone (Figure 25); an Early Silurian
prograding siliciclastic wedge in the Appalachian Basin derived from tectonic source
areas to the southeast (Figure 26). It unconformably overlies the red Queenston Shale
with a sharp, near-planar contact in outcrop described by Dennison and Head (1975) as
the Cherokee unconformity. The Queenston Formation, which represents an extensive
deposit displaying environments ranging from coastal plains to supratidal and shallow
marine settings, is dominated by mud and in some places calcareous silt and sand,
bioclastic-rich terrigenous units and silty bioclastic limestones. The upper continental
muddy part of the Queenston Formation in Ontario and New York as well as the
overlying Whirlpool Sandstone (Martini and Salas, 1983; Middleton et al., 1983) has
been attributed to the westernmost filling phase of the Appalachian Basin during the
relaxation stage of the Taconic orogeny (Quinlan and Beaumont 1984; Beaumont et al.,
1988; Ettensohn 1991). Eustatic drop in sea level, associated with Upper Ordovician
Saharan glaciations, was also postulated to have enhanced a continental setting and led to
the development of the unconformity at the top of the Queenston Formation (Dennison
1976; Ziegler et al., 1977; Brenchley and Newall 1984) and subsequent deposition of the
Whirlpool Sandstone. More recent studies in the light of foreland basin models however
pointed out (Middleton 1983; Cheel and Middleton 1993) that the development of the
unconformity at the top of the Ordovician rocks and subsequent spreading of a braided
plain across southwestern Ontario, which led to the deposition of the lower part of the
Whirlpool quartzose sandstone, are best explained tectonically. Whirlpool Sandstone
deposition was attributed to lithospheric flexure resulting from tectonic loading of

67

adjacent Taconic highlands (Figure 26). A period of tectonic quiescence in the westward
thrusting of the Appalachians would have allowed for isostatic uplift and erosion
accompanied by rapid spreading of a clastic unit over the foreland basin. Easing off of
thrusting just prior to this would explain reduced subsidence and rarer marine incursions
into the basin in the early stage of Whirlpool deposition. The thin Whirlpool Sandstone
extends very far out into the foreland basin because accommodation space was exceeded
by sediment supply derived from the Taconic orogen. Thickness of the lower part of the
Whirlpool Sandstone is variable because of irregular topography on the unconformity,
which may have been related locally to differential incision and, more broadly, to highs
and lows caused by tectonic activity.
The Whirlpool Sandstone has been extensively studied resulting in a gamut of
interpretation of its architecture as well as its depositional environment (Gilbert 1899;
Fairchild, 1901; Wilson 1903; Grabau, 1913; Williams, 1919; Johnson, 1934; Alling,
1936; Holstein, 1936; Lockwood, 1942; Bolton, 1949; 1953, 1957; Geitz, 1952; Fisher,
1954; Koepke, 1960; Martini, 1966; Sanford 1969; Pitrowski 1981; Seyler, 1981,
Metzeger, 1982; Calow, 1983; Martini and Salas, 1983; Laughrey, 1984; Martini and
Kwong, 1985; Pees, 1986; Coogan, 1990). Brett et al., (1990) interpreted the lower, nonmarine part of the Whirlpool as a major eustatic lowstand deposit, while Ryder et al.,
(1996) attributed Whirlpool deposition to fluvial backfilling of incised valleys during sealevel rise. Cores from the Whirlpool show evidence for transgressive reworking of the
upper part of the sandstone, and it is likely that the lower part was deposited in shallow
incised valleys either during late lowstand or during early transgression. These two parts

68

of the Whirlpool are separated by a distinct, regionally correlative boundary, which may
represent a transgressive surface (Castle, 1998).
Vertical Scale
10 m

Whirlpool
Sandstone

Figure 25: Generalized stratigraphic chart for the Silurian system in New York.
(Modified from Telford 1978).

A near-shore marine environment has also been proposed for the Whirlpool Sandstone,
but studies in the last decade tilt evidence in favour of a near-shore, wave-influenced
marine upper unit and a lower unit formed in a braided fluvial depositional environment
(Rutka et al., 1991) probably deposited in incised valleys developed on the Cherokee
unconformity (Castle and Byrnes, 1998).

69

Figure 26: Early Silurian paleogeographic map of the Niagara Region. (Image taken
from Tesmer, 1981); study area highlighted in red circle.
This study focuses on the lower unit of the Whirlpool Sandstone exposed at Artpark,
New York and applies the concepts of architectural element analysis aided with GPR
imaging in the interpretation of the anatomy as well as the fluvial style responsible for its
deposition.

3.1. 3 Study Location and Direction


The Artpark exposure of the Whirlpool Sandstone is located on the U.S side of the
Niagara River near Lewiston (New York, USA) by the Niagara River (Figure 27).

70

80 Km

500 m

Figure 27: (A) Location of the Whirlpool Sandstone outcrop in New York (B) Detailed
outcrop location - outcrop highlighted in red.
From Robert Moses Parkway, turn west to Centre Street and south on 4th Street. Earl
Brydges Park is at the south end of 4th Street (Figure 27B). The outcrop is at the
southwest end of the park by the Niagara River. Location coordinates of the outcrop are

71

43 9'44.01"N, 79 2'41.18"W. The entire outcrop is about 200 metres wide and it is
reasonably well exposed (Figure 28). Accessbility is limited and outcrop photos had to be
taken from the Canadian side of the Niagara River.

NE

SW

A
Queenston Shale

20 m

A
5m
5m

Figure 28: (A) Outcrop photo of Whirlpool Sandstone at Artpark (B) AB section of the
outcrop in (A).

3.1. 4 Study Methodology


Outcrop interpretation is based on identification of architectural elements defined by
grain size, bedform composition, internal sequence, and external geometry (Miall, 1985)
in addition to geophysical data from Ground Penetrating Radar profiles obtained behind
the outcrop face both parallel and orthogonal to outcrop orientation. GPR surveys were
conducted with MAL Geoscience Ramac 250 MHz shielded antenna designed for submetre scale resolution and moderate penetration depth. This antenna is ideal for surveys
in urban areas where surface electromagnetic signals (from power lines and radio towers)

72

interfere with reflected radar signals. The antenna resolution is also ideal for resolving
channel-scale features and intra-channel architectural elements.
Four GPR profiles were recorded in common offset mode; two parallel to the outcrop and
the other two orthogonal to the outcrop. Each line was re-recorded in reverse direction to
detect and remove transient noise that does not repeat at the same location in the
radargram.
Data acquisition parameters used are:
Antenna separation: 0.36 m
Sampling Frequency: 2415 MHz
Trace Interval: 0.099 m
Velocity: 0.12 m/ns (calculated from depth to Whirlpool - Queenston contact horizon on
the radar profile).

Figure 29: Raw, unprocessed GPR data from the Whirlpool Sandstone, Artpark (USA)
In raw unprocessed GPR data, beddings and bounding surfaces are hardly visible (Figure
29); acquired radar profiles were processed with REFLEX GPR processing software and
displayed in three dimensions with GPR-Slice display and interpretation software.
Shielded antennas used for data acquisition minimized the occurrence of spurious

73

reflections in the radar profile; hence, few data processing steps like de-wow, signal gain,
and diffraction stack migration were applied.

3.1. 5 Outcrop Description


The Whirlpool Sandstone is thin, sheet-like in character and rarely exceeds 9 metres in
thickness. It has been studied in many locations along the Niagara Escarpment but the
fairly continuous exposure of about one hundred and fifty metres at Artpark, provides a
suitable dimension required for architectural element analysis. The Whirlpool Sandstone
was interpreted as a two-unit division consisting of lower unit (generally 1-7 m thick)
deposited in a braided fluvial depositional environment overlain by a marine unit
(generally 1-2 m thick) formed in nearshore wave-influenced environment.

Figure 30: Artpark vertical section summarizing the main outcrop observations (Image
taken from Rutka, 1986).
74

The upper unit of the sandstone outcrop studied at Artpark has however been lost to post
depositional erosion leaving predominantly the lower unit of braided fluvial origin.
Rutka (1986) earlier studied the Whirlpool Sandstone at Artpark and compiled her
observations at various outcrop exposures along the Niagara Escarpment at Artpark into a
vertical profile. The vertical section reveals three upward fining successions dominated
by trough cross-beds. Planar cross-beds and ripple cross-laminations occur only as minor
lithofacies at the top of each trough cross-bedded unit reflecting waning flow (Figure 30).
Table 3: Lithofacies in the Whirlpool Sandstone at Artpark (New York)

Facies Code

Facies

Sedimentary Structures

Interpretation

St

Medium to

Solitary or grouped

Sinuous crested and

fine sand

trough cross beds.

linguoid (3-D) dunes

Medium to

Ripple cross

Ripples

fine sand

lamination

(lower flow regime)

Medium to

solitary or grouped

Transverse and

fine sand

planar cross beds.

linguoid bedforms

fine sand

Horizontal Lamination

Plane bed flow

Sr

Sp

Sh

Parting lineation

75

This study extends previous interpretations at the outcrop to three dimensions via
architectural element analysis and GPR imaging. Architectural elements are defined and
annotated on an outcrop photomosaic of the Whirlpool Sandstone at Artpark. Lithofacies
types in the Whirlpool Sandstone exposure at Artpark are listed in Table 3. They are
based on Miall (1978c) lithofacies codes for fluvial environments.
The Artpark exposure of the Whirlpool Sandstone is about 150 metres long oriented
northwest-southeast and has a thickness of 4 -7 metres. The sandstone is thicker (about 7
metres) towards the southeast end while much of the upper part of the sandstone beds in
the northwestern end of the profile have been eroded (Figure 28 A).
Interpretation of the rank and bounding surfaces is based on their shape, lateral extent and
associated lithofacies. Three main bounding surfaces numbered 1, 2 and 3 subdivide the
sandstone exposure into three major units (Figure 32A) described as lower, middle and
upper sand sheets.
Bounding surface 1 is the channel base separating the Lower Silurian Whirlpool
Sandstone from the underlying Ordovician Queenston Shale; it is the lower bounding
surface of the basal sand sheet with a slight upward curvature and shaly fill of rip up
clasts from the underlying Queenston Shale. This surface is a well-known marker
observed at many other locations where the Whirlpool Sandstone is exposed. It is
interpreted as a 6th order unconformity surface separating the Ordovician Queenston
Shale from the Silurian Whirlpool Sandstone. This surface is observable on outcrop
photomosaic and GPR profile (Figure 32A and B)
Surface 2 is the lower bounding surface of the middle sand sheet. This unit comprises
about 2 metres of thick trough-cross-bedded sands of mostly sandy bedforms. The upper

76

3 metres of the sandstone succession rests on Surface 3 and reveals scour (oriented eastwest and filled with shaly trough cross-bedded sandstone) into the underlying sandsheet,

3.1.6 GPR Profiles


Unlike most Whirlpool Sandstone outcrops along the Niagara escarpment, the absence of
thick overburden overlying the Whirlpool Sandstone and a flat surface (Elevation
changes across recorded profiles are less than 1 metre) make GPR survey feasible at the
Artpark outcrop.

Niagara River

Robert Moses

104

Earl Brydges

200 m

Outcrop exposure (see


photo in Figure 32 A)
D
G C
A
F
E
GPR Lines
H
B
Figure 31: Location of GPR lines at Artpark State Park.

77

Radar facies observed are sedimentary structures and architectural element-scale features
described and interpreted based on shape of reflectors, reflection dip, relationship
between reflectors, reflection continuity and amplitude.
Line AB (Figures 28A and 31) is the transect where a GPR profile parallel to outcrop
face was acquired; it was recorded one metre behind the outcrop. The GPR profile
reveals major bounding surfaces labeled 1, 2 and 3 also observed on the outcrop photo
(Figure 32A and B). Northwest dipping inclined reflectors at the east end of the GPR
profile are visible in outcrop; they are interpreted as sandy planar cross-beds (facies Sp in
Figures 32A and B). Metre scale curved reflectors seen on the radargram and outcrop are
interpreted as trough cross-beds (facies St). The larger scale concave upward reflectors
with low amplitude fill (indicative of high mud content) observed towards the north end
of the GPR profile and annotated outcrop photo is labeled SC on Figures 32A and B. It is
interpreted as late stage scour incised into the middle sand sheet (Figures 32A and B).
Ground Penetrating Radar profiles orthogonal to the outcrop clearly reveal that the
Whirlpool Sandstone-Queenston Shale contact is erosional Figure 33 (interpreted from
reflection attenuation pattern and horizontal to sub-horizontal low reflection amplitude
away from the channel margin).

78

A
SC
St

B
St

St

SB
3m

St
Sp

Sp

SB

Queenstone Shale

Sp
Sp

5m
1

B
Low

SC

St
Sp

Sp

St

SB

1
Queenstone Shale

Depth (m)

SB

Sp

High

6
Amplitude Scale

Figure 32: (A) Annotated Whirlpool Sandstone photo at Artpark (B) Annotated GPR line AB (SB Sandy Bedform; SC Scour; St troughcross-bedded sandstone, Sp - planar cross-bedded sandstone).

79

The main channel dimension, not obvious in outcrop, is evident in radar section recorded
orthogonal to the outcrop (Figure 33); the channel is about 6 m deep and between 60 and 100 m
wide (assuming the width of the channel is at least twice the width of the portion of the channel
imaged.

3.1.7 Radar Facies


Radar facies observed are lithofacies and architectural element-scale features described and
interpreted in the context of the depositional environment of the Whirlpool Sandstone. Individual
sedimentary structures are not described as radar facies or correlated to the outcrop photomosaic
as the resolution of the photo and GPR image do not permit unequivocal correlation of most
sedimentary structures to the radargram; where they can be clearly resolved, they are however,
highlighted on the radar profiles. Orientation of the GPR lines is documented to reflect the fact
that reflection patterns and geometry observed in radar profile depends on the orientation of the
profile in relation to the imaged channel fill.
Four major radar facies (architectural element scale features) were recognized and correlated to
their lithological expression in outcrop in addition to the major bounding surfaces. These facies
are described below and summarized in table 4.
Concave-upward reflector with low amplitude fill
This is highlighted in Figure 32B as SC; this facies is characterized by low amplitude concaveupward basal reflector with a low amplitude fill. It is interpreted as intra-channel scour about 15
metres wide and 2 metres deep filled with muddy sand (interpreted from low reflection
amplitude)

80

Concave-upward reflector with high amplitude fill


This is highlighted in Figure 33 as CH; this facies is characterized by high amplitude fill
adjoining low amplitude planar to sub-horizontal reflectors at the margin (OF). This is
interpreted as the main Whirlpool Sandstone channel.
High amplitude planar reflectors
This is highlighted below in Figure 33 as OF (Channel overbank deposit); this facies is
characterized by high amplitude horizontal reflectors underlain by low amplitude planar to subhorizontal reflectors. This is interpreted as channel margin overbank deposit.

CH
Queenston Shale

Depth (m)

OF

Figure 33: GPR line EF orthogonal to the Whirlpool Sandstone outcrop at Artpark (CH
Channel; OF Overbank deposit).

Low amplitude horizontal reflectors


This facies is characterized by thinly-bedded low amplitude horizontal reflectors with a basal
indistinguishable reflection (reflection-free) zone indicating attenuation of radar signals typical
of clay-rich sediment. This radar facies is interpreted as the basal the Queenston Shale.

81

Table 4: Radar facies identified in the Whirlpool Sandstone GPR profile at Artpark. (Radar facies are annotated on profiles - Figures 32 and 33).

Radar Facies

Outcrop Photo
Profile perpendicular to outcrop photo not available.

Architectural
Elements

Geometry and Radarfacies


description

Main Channel.

Concave upward base and channel margin


outline resting on the Queenston Shale.
Channel fill is characterized by high

CH

amplitude reflectors with low amplitude


basal section suggesting significant shale
content.

OF

Profile perpendicular to outcrop photo not available.

OF

SC
SC

Overbank sand sheets

Horizontal reflectors with lower amplitude

underlain by muddy

near contact with underlying Queenston

overbank fines.

Shale.

Intra-channel Scour

Profile parallel to the channel axis

(2-3 m deep) at the

Concave upward reflectors within the main

top of the Whirlpool

channel with lower amplitude near the base.

Sandstone. Scour
outline suggests an

5m

2m

Profile perpendicular to outcrop photo not available.

East-West trend.

Queenston Shale.

Low amplitude horizontal reflection


underlain by weak indistinguishable
reflection.

Queenston Shale

82

3.1.8 Outcrop-GPR Integration


While outcrop observation at Artpark reveal a sheet geometry and apparently homogeneous
lithology, integration of outcrop mapping at Artpark with GPR data suggests that the entire
outcrop lies in a northwest-southeast trending channel geometry (Figures 34 and 36) with a
sandy mostly trough-cross-beded fill. Depth and width obtained from radar profiles show the
small scale of the Whirlpool River and reflection amplitude suggests a shaly fill. Variability in
reflection amplitude (Figure 32B) shows that even in braided fluvial deposits, there could be
significant lithological heterogeneity (lower amplitude towards the north end of the outcrop) not
obvious in outrop section. The implication of this study is that architectural data that are often
difficult to obtain from outcrops such as reliable channel dimension, local paleoflow data,
sandbody stacking patterns and lithological heterogeneity can be obtained via GPR imaging.

3.1.9 Discussion
Definition of the various types and scales of bounding surfaces, the constituent lithofacies
assemblage, lack of fossils or ichnofossils, predominantly sheet-like geometry, lack of lateral
accretion macroform as well as radar reflection patterns suggest deposition in a distal low
sinuosity sheetflood sand-bed river characteristic of distal braidplain. Preservation of transitional
to upper flow-regime beds, presence of plane-laminated sand, trough cross-beds, low angle
cross-bedded sand, erosional facies relationships as well as vertical aggradation within the
shallow channels and radar facies validate this interpretation. Paleocurrent measurements by
Rutka (1986) at the Whirlpool Sandstone outcrops along the Niagara escarpment shows variable
local paleoflow possibly due to intra-channel braiding and scours (Figure 35)

83

Overbank deposit

Crossbar channel

Braid bar
Braid bar

Depth (m)

5m

N
Figure 34: Three-dimensional reconstruction of the Whirlpool Sandstone architecture at Artpark integrating outcrop and GPR data.

84

Outcrop data and radar profile parallel to the outcrop face suggests predominance of
element SB (Sandy Bedforms) with late stage muddy cross channel deposit. GPR profiles
perpendicular to the axis of the main channel show channel margin and overbank areas
with elements OF (Overbank deposit) predominant; they were probably deposited as a
product of waning overbank flow.

Figure 35: Regional paleocurrent pattern for the lower Whirlpool Sandstone based on
trough cross-bed measurements. (Image taken from Rutka, 1986) red circle highlights
outcrop location at Artpark.

85

While the lack of vegetation and abundant sediment supply from the Taconic orogen
might have influenced the fluvial style, the cohesive Queenston Shale possibly exerted an
opposing bank stabilizing control on the Whirlpool channels resulting in the formation of
channelized deposition observed in the GPR profiles.

ba

nk

os

it

Depth (m
)

r
ve

ep

Figure 36: Fence diagram display of GPR data from the Whirlpool Sandstone, Artpark
(New York, USA) See location of GPR lines in Figure 31.
The origin of the Whirlpool Sandstone has been explained by both lithospheric flexure
associated with the Taconic Orogeny (Rutka et al, 1991) and eustatic drop in sea level
resulting in the development of incised valleys into the Queenston Shale (Castle et al,
1998). Brett et al., (1990) interpreted the lower, nonmarine part of the Whirlpool as a
major eustatic lowstand deposit, while Ryder et al., (1996) attributed Whirlpool

86

deposition to fluvial backfilling of incised valleys during sea level rise. Similar braided
fluvial deposits such as Lower McMurray Formation and Dina Formation (Riediger
1999; Crerar and Arnott, 2007) overlain by transgressive deposits have been observed as
the basal fill of incised valleys cut into the sub-Cretaceous unconformity in Alberta and
Saskatchewan respectively. Regional study of the Whirlpool Sandstone integrating
outcrop and GPR data will be vital for validating either model.
The Queenston-Whirlpool contact has often been described as distinctively flat because
most of the Whirlpool Sandstone exposures are near parallel to Whirlpool channel axes,
giving an apparent impression of a planar contact. In outcrops, the apparent planar
contact with the underlying Queenston Shale suggests a non-erosional contact described
by Rutka et al (1991) as a disconformity and the Whirlpool Sandstone as the product of
weakly and possibly unconfined fluvial channels. This implies lack of deep channels cut
into the Queenston Shale, interpreted as an indication that most of the Whirlpool streams
energy was used primarily for bedload transportation rather than for erosion of channels
into the underlying shale. Radar profiles perpendicular to the outcrop, however, reveal
about 7 metres of incision into the Queenston Shale by the lower Whirlpool channels
(Figure 33 and 34), suggesting that the Whirlpool Sandstone was not deposited in
unconfined channels.
Lack of vegetation is sometimes invoked as an important control on pre-Devonian fluvial
architecture (Schumn 1968; Long 1978; Cotter; 1978; Davies and Gibling, 2010) as
vegetation affects channel morphology and shifting behavior through increased bank
stability. Experimental study by Gran and Paola (2001) demonstrates that vegetation
plays a critical role in bank stability, constrains channel migration resulting in the

87

formation of deep and narrow channels. Pre-Devonian fluvial systems such as those that
deposited the Whirlpool Sandstone might have had reduced bank stability which
promotes braided sheetlike deposition. While the lack of vegetation in the Lower Silurian
has been invoked as a possible explanation of predominance of weakly-channelized
bedload streams with the resultant fluvial architecture and composition similar to low
energy distal braidplain sand sheets (Rutka et al., 1991), evidence of cross channel scour,
main channel margin and upper plane bed channel margin deposit suggest a channelized
higher energy sandy braided fluvial system.

3.1.10 Conclusion
GPR imaging at Artpark shows that even in a geological formation with minimal
lithological contrast, GPR imaging can be quite revealing. This study also shows that
reflection configuration, geometry, and radar reflection amplitude can be vital
interpretive tools in the study of ancient fluvial environments, since amplitude is
controlled largely by lithology. While full three-dimensional GPR survey was not
conducted at the outcrop, it is obvious from this study that a few staggered GPR lines
acquired properly with the ideal equipment can yield vital architectural information.
This study illustrates the value of three dimensional outcrop studies integrating GPR with
outcrop data as it reveals a more complex architecture than portrayed by earlier outcrop
study at the site (Figure 30). This confirms the usefulness of GPR as an invaluable tool
for three dimensional outcrop studies, as it reveals architectural details not otherwise
visible in outcrop and helps refine previous interpretations based on two-dimensional
outcrop data.

88

Case Study 2 Navajo Sandstone

3.2.1 Introduction
Aeolian deposits were long thought of as relatively simple and homogeneous reservoirs,
yet production histories from well-known eolian reservoirs such as the Permian
Rotliegendes Formation in the North Sea (Glennie, 1990) and Jurassic eolian deposits in
the Western USA (Lindquist, 1988) prove that this is far from reality. Research findings
in the last two decades suggest that eolian reservoirs are highly heterogeneous and are
often characterized by highly layered architecture with strong directional permeability.
Detailed and basin studies of modern and ancient eolian deposits over the last few
decades reveal that despite their internal architectural complexities, anisotropy in eolian
reservoirs is quite predictable, as this is often controlled by the initial depositional history
(Glennie, 1990 ; Kocureck and Havholm, 1993)
There has been considerable research effort in the last few decades geared at unraveling
controls on eolian depositional architecture, but much of the effort has been focused on
modern and recent eolian deposits (Leatherman, 1987; Bristow et al., 1996, 2000a, b;
Harari 1996; Jol et al., 1996a; Van Heteren et al., 1996, 1998; Smith et al., 1999; Neal
and Roberts 2000; Mller and Anthony, 2003). Because many ancient eolian deposits do
not have applicable modern analogs (Fryberger, 1990a), unraveling architectural
complexities in eolian reservoirs hinges on detailed studies of ancient eolian deposits via
subsurface three dimensional seismic imaging and three-dimensional outcrop studies.
Jol and Bristow (2003) assessed the viability of three dimensional eolian outcrop studies
utilizing Ground Penetrating Radar imaging at a Navajo Sandstone outcrop in Utah. They
89

evaluated the effectiveness of various GPR antenna frequencies in delineating the internal
stratification of the Navajo Sandstone and conducted a test three dimensional GPR survey
(Figure 37) at Zion National Park, Utah.

Figure 37: 3D GPR image of the ancient Navajo dune complex obtained with 200-MHz
antenna .The image shows three sets of dipping reflections interpreted as crossstratification with two horizontal reflections that are interpreted as erosional truncation
surfaces. (Image taken from Jol et al., 2003).
They concluded that the technique holds immense potential as a source of input data for
computerized simulations of sedimentary successions and realistic geostatistical
modelling of eolian deposits with better constraint on paleocurrent trends, distribution of
nonrandom heterogeneities and scale of architectural units. Penetration depth of greater
than 40 m was obtained with the 50-MHz antenna and 16 m penetration depth with the
200-MHz antenna.

90

Unlike the test study at Zion National Park, considered to be within the centre of the
Navajo erg, this study was conducted on the interdune deposits at a Navajo outcrop near
Moab, Utah (Figure 39A) ; the study location is strategically chosen, because
stratigraphic relationships and facies associated with interdune processes, and their
internal architecture often yield vital clues about allogenic controls (especially climate
and relative sea level) and autogenic controls (groundwater levels and vegetation) on
sedimentation and subsequent preservation. This study also shows the value of GPR
imaging in revealing architectural details that are not visible in outcrop due to limited
exposure. In addition, it demonstrates the capability of GPR imaging on ancient interdune
deposits as a tool for high-resolution three dimensional outcrop studies in the quest for
deciphering autogenic and allogenic controls on architecture of ancient eolian deposits.
The sophistication of current eolian facies models reflects the tortuous journey from 1D
facies model to 4D facies models that combine the spatial geometrical and architectural
complexity of dune and interdune deposits on a variety of scales with the temporal
dimension. Despite these advances, there are still many unsolved puzzles in eolian
sedimentary research, especially those relating preserved sedimentary architecture to the
processes responsible for its generation. Demonstrating an unequivocal link between
allogenic forcing mechanisms such as climate change and the generation of stratal bodies,
differentiating between the products of intrinsic (autogenic) process such as bedform
migration and external (allogenic) processes such as climate change and tectonic basin
evolution, and demonstrating the extent to which these two sets of processes are
independent of each other remain a challenge. While much effort is currently being
expended in the study of modern ergs (Bristow et al., 1996, 2000 a, b, c; Harari 1996; Jol

91

et al., 1998; Bailey et al., 2001; Clemmensen et al., 2001, Botha et al., 2003; Havholm et
al., 2003; Hugenholtz et al, 2007), the answer to most of these unanswered questions lie
hidden in ancient eolian deposits, and GPR imaging technology holds significant promise
in unraveling these mysteries.

3.2.2 Study Objectives


The objectives of this study are:

to demonstrate that sedimentological and architectural interpretation can be made


from GPR reflection amplitude, reflection configuration and stratal pattern.

to document interdune radar facies from the Navajo Sandstone as a guide for
future research on eolian interdune deposits; currently there are few documented
radar facies from ancient eolian deposits, mostly from dry phase dune strata.
Radar facies from the Navajo Sandstone will provide an opportunity to compare
interdune radar reflection patterns with lithofacies and architectural elements
observed in other eolian outcrops.

to evaluate the architecture of interdune deposit at the study site and the
relationship to local and external stratigraphic controls.

to evaluate the value of three-dimensional versus two-dimensional GPR imaging


for outcrop studies.

3.2.3 Navajo Sandstone - Geological Overview


The Jurassic Navajo Sandstone is the largest preserved eolian accumulation on earth. The
volume of sand deposited by the Navajo Sand Sea is estimated at between 40000 and
60000 km3 (Marzolf, 1988) covering about 265,000 km2 in the US Western interior

92

(Blakey et al 1988) and exposed extensively on the Colorado Plateau (USA). The Navajo
Sandstone has been extensively studied (Freeman and Visher, 1975; Middleton and
Blakey, 1983; Kocureck and Hunter, 1986; Herries, 1993) as an excellent analogue for
subsurface eolian reservoirs, yet it continues to pose puzzles for eolian researchers.
The Navajo Formation is the uppermost part of the Glen Canyon Group and overlies the
Triassic Kayenta Formation; it is overlain by the Middle Jurassic Temple Cap and
Carmel Formations (Figure 38). The Navajo Sandstone consists of fine grained, wellsorted frosted quartz sands. Crossbeds are typically long, sweeping sigmoidal shaped
structures, over 15 metres in length, with longer asymptotic bottom sets. Within the
dunes, cross-bed sets can reach over 30 metres in thickness with foresets dipping 20 0350. Body fossils and ichnofossils are rare in the Navajo Sandstone; terrestrial biota such
as ostracods, freshwater crustaceans, dinosaur tracks and skeletal remains of mammallike reptiles have, however, been reported from interdune areas (Blaird, 1980; Lockley
and Conrad, 1989).
While the Navajo Sandstone is best known for the occurrence of gigantic, sweeping sets
of cross-bedding and their internal lamination, significant lithological heterogeneity
occurring as lacustrine limestones, cherts, fluvial deposits, conglomerate lenses and
deformed cross-bedding are present as non-aeolian facies, mostly in the eastern fringe of
the ancient erg. Most of these non-eolian facies are associated with interdune processes,
lacustrine and fluvial incursions into the ancient eolian deposit; their internal architecture
and stratigraphic relationships often unravel regional controls on eolian sedimentation
and preserved architecture. Interdune areas are troughs between dunes ranging from
interdune flats to interdune depressions. They are typically classified as dry, damp and

93

wet eolian systems with the greatest variety of interdune structure occurring in wet and
damp eolian systems commonly expressed as wind ripples, adhesion laminae, adhesion
ripples and contorted bedding.
,

Chart Legend

Figure 38: Generalized stratigraphy at the study site showing the Navajo Sandstone.
(Chart from Blakey, 2008).

94

They consist of thin sandstone laminae capped by a calcareous unit possibly formed in
ephemeral ponds. Although, it has long been recognized that interdune deposits pose
problems in petroleum reservoirs as permeability barriers to fluid flow, the wealth of
information within the relatively small proportion of facies within interdune regions has
rarely been given adequate consideration, although this facies often holds vital clues to
paleoenvironmental interpretation and provides predictive insight into depositional
controls (Ahlbrandt and Fryberger, 1980; Hummel and Kocurek, 1984). This study
evaluates the usefulness and limitations of GPR imaging in discriminating centimeterscale interdune facies as observed in the radar profiles as well as the potential for a basinwide three-dimensional study of eolian interdune deposits using this technology.
Compared to cross-bedded sand dunes, interdune deposits have received little attention.
Where reported in the Navajo Sandstone, interdune deposits occur as beds 2-6 metres in
thickness sandwiched between the eolian cross-bedded dune sandstone.

3.2.4 Study Location and Direction


This study was conducted at Big Mesa, a Navajo Sandstone exposure about 21 km
Northwest of Moab (Figure 39A and B), Utah (USA). The outcrop is located off
Highway 163 via Highway 313 through Mineral Bottom Road, about 900 m from the
intersection of Highway 313 and Mineral Bottom Road (Figure 40). GPS location of the
outcrop is 3838'6.84"N, 10948'20.68"W.

95

N
313
163
313

163
3 Km

279

Figure 39: (A) Map showing study location within Utah (B) Map showing outcrop
location relative to Moab, Utah.

96

3.2.5 Methodology
GPR surveys were conducted with MAL Geoscience Ramac 250 MHz shielded antenna
designed for centimetre scale resolution and moderate (5-15 m) penetration depth. This
antenna is ideal for surveys in areas where extraneous surface electromagnetic signals
(from power lines and radio towers) interfere with reflected radar signals. The antenna
frequency provides the required resolution for sub-metre scale sedimentary features and
interdune architectural elements observed at the study site.
Acquisition parameters used are:
Antenna separation: 0.36 m
Sampling Frequency: 2760 MHz
No of stacks: 16
Trace Interval: 0.099 m
3D survey line spacing: 45 x 45 lines at 1 m spacing
Velocity: 0.118 m/ns (calculated from diffraction hyperbolae)
Acquired radar profiles were processed with REFLEX GPR processing software and
displayed in three dimensions with GPR-Slice display and interpretation software.
Data processing steps include de-wow, signal gain, background filtering and diffraction
stack migration. Penetration depth was significantly lower than observed at Zion National
Park (maximum penetration of 6 metres), Utah where earlier GPR surveys on the Navajo
dune have been conducted (Jol. et. al, 2003) suggesting higher clay content.

97

3.2.6 GPR Lines


Ninety-five GPR profiles were recorded in common offset mode: single lines 136 - 140
and ninety lines (forty five each in othorgonal directions) laid out in order to obtain a 3D
GPR volume of the outcrop at the survey site. Topographic survey was not conducted at
this site, hence GPR lines were recorded along transects with reasonably flat topography
(elevation variation of 1 metre or less). The GPR profiles were recorded about 20 50
metres behind the outcrop; location and scale of the GPR lines are documented in Figure
40.

139 140
138

3D
137
136

313

oad
R
m
o
Bott
l
a
r
e
n
Mi
N
150 m

Outcrop (see Figure 41B and


43A for outcrop photo)

Figure 40: Map showing field locations of radar profiles and 3D survey.

98

Radar reflection and signal attenuation patterns reveal three radar facies corresponding to
outcrop observation. They are annotated on the radar profiles, described below and
summarized in Table 5.

3.2.7 Outcrop Data


Outcrop exposure at the study location reveals a wetting upward interdune succession
from a basal dry eolian cross-bed phase through damp stage typified by adhesion ripple
and contorted beds to a fully wet phase preserved as an interdune pond carbonate (Figure
42).
Three facies associations were identified at the outcrop, each indicative of the dune
phases preserved at Big Mesa. They are described below:

Facies Association 1
This unit comprises well-developed foreset laminae which dip downwind at angle of 15
25. Near the base of the slip face, the foreset laminae often flatten out, giving rise to a
concave-upwards profile.
Individual foreset laminae, which are usually 510cm thick, are grouped together into
cross-bed sets which are separated by sub-horizontal bounding surfaces. Crossbed sets
are much smaller (mostly 3 5 metres high) than their gigantic central dune counterparts,
which commonly reach heights of over 20 metres (compare Figure 41A and B).

99

Crossbed Set

Interdune carbonate

Crossbed Set

1.5 m
5m

Figure 41: (A) Outcrop photo showing central dune giant cross-bed sets of the Navajo
Sandstone, Utah (USA) note geologist highlighted in white oval for scale (B) Photo
showing smaller cross-bed set of Facies Association 1 at Big Mesa near Moab,Utah
(photo taken on Mineral Bottom Road looking north east - see outcrop location in Figure
40).

Facies Association 2
This unit comprises a reddish-brown, diagenetically altered, fine-grained sandstone unit
with contorted bedding (Figure 42) indicating an episode of increased moisture within the
interdune environment. Contorted beds typically occupy the stoss slope, slip face and
apron of the underlying undeformed cross-bed set.

100

Interdune freshwater
carbonate

2m

Contorted damp interdune deposit

Figure 42: Outcrop photo showing damp eolian interdune deposits of Facies Association
2 overlain by flat-bedded fresh water carbonates of Facies Association 3 photo taken at
Mineral Bottom Road looking north east. Image location is from the south end of the
outcrop outline shown in Figure 40. Scale rod is 1.5m in length.

Facies Association 3
This unit, about 1.5 m thick, comprises 10 20 cm of bedded cherty limestone with
horizontal to warty bedding planes (Figure 42). This facies is indicative of fully wet
interdune conditions in which limestone was precipitated in fresh-water interdune ponds.
101

3.2.8 GPR Facies


Radar facies at Big Mesa are defined based on reflection configuration and amplitude.
They are interpreted in the context of outcrop observations and knowledge of local
geology. Four radar facies were observed, they are described below:

High-amplitude horizontal reflectors


This facies is characterized by high amplitude horizontal reflectors each about 20 cm
thick comprising a unit of about 1 metre in thickness (Figure 43B). It is the topmost set of
reflectors capping the underlying dipping reflectors. This facies is interpreted as
interdune cherty limestone observed in outcrop section recording the wet interdune phase
of the Navajo Sandstone at the study location

Low-amplitude reflectors
Low amplitude reflectors are observed underlying the high amplitude planar reflectors.
They are found on the flanks of the underlying mound shaped reflector (Figure 43B) with
internal dipping surfaces. They are interpreted as clay-rich damp phase interdune deposits

High amplitude mound-shaped reflectors


High amplitude mound-shaped reflectors are observed underlying the high amplitude
planar reflectors (Figure 43B); they are interpreted as dry-phase eolian dunes. Their scale
as observed in outcrop suggests comparatively small cross-bed sets typical of dry-phase
interdune deposition (Figure 43 B)

102

A
North

Interdune carbonate

South

Dry phase interdune cross-beds

Damp phase interdune deposit

Wet phase
interdune
carbonate

Damp phase interdune deposit

Depth (m)

Dry phase dune

Depth (m)

Wavy interdune beds

Figure 43: (A) Outcrop photo annotated with major architectural elements (B) First
outcrop-parallel radar line of the 3D survey annotated with major radar facies (C) Radar
line 139 showing wavy interdune beds (D) Radar line 140 showing wavy interdune beds
see location of GPR lines in Figure 40.
103

Depth (m)

Wavy interdune beds

High amplitude wavy reflectors


20-30 cm high amplitude wavy reflectors with about 5 metres wavelength (figures 43C
and D). This facies was not observed in outcrop probably due to limited exposure.

104

Table 5: Radar facies at Big Mesa Navajo outcrop, Utah. (Red oval highlights major radar facies)
Radar Facies

Radar Facies Configuration

Facies Interpretation

Planar reflectors each between 20 to 30 cm thick Wet phase thinly-laminated cherty freshwater limestone pond
deposits

High amplitude wavy reflectors.

Damp phase interdune wavy beds

High amplitude mound-shaped reflectors with

Mound-shaped dry phase eolian dune with internal cross-beds

internal dipping reflectors.

Low amplitude reflections with

Wet phase clay-rich interdune deposit occupying flanks of the

indistinguishable bedding pattern

mound shaped dry phase deposit

105

3.2.9 3D Survey
The test 3-D GPR survey at Big Mesa, Utah evaluated the usefulness of 3-D surveys over
staggered 2D lines; three dimensional GPR imaging shows the benefit of GPR imaging
especially via time slicing and and incorporation of a plan-view image with vertical
reflection profiles (Beres at al, 1995). Three dimensional surveys have been particularly
informative in seismic imaging where changes in channel planform and large scale intrachannel features have been observed in GPR time slices (Posamentier et al, 2007); this
has however not been widely embraced in GPR imaging due to the seb-metre spacing
required between GPR lines for three dimensional surveys. The advantage of threedimensional GPR survey lies in the possibility of viewing depositional patterns via fence
diagrams of closely spaced profiles and map-view amplitude mapping. Analysis of
amplitude variation in time slices is one of the most important uses of time slices
obtained from 3D surveys as reflection amplitude is closely linked with lithology. 3D
GPR survey at study site reveals the northwest orientation of higher amplitude dry phase
interdune deposits flanked by clay-rich low amplitude interdune sands (Figure 44B). This
is of predictive value for subsurface modelling of interdune reservoir sands as
understanding the spatial trends of reservoir type facies relative to the clay-rich damp and
wet phase facies is a critical prerequisite for interpreting facies distribution. Similar
mound-shaped dunes can be expected on either side of the clay-rich interdune deposits
with identical orientation to the imaged dune, as this is typical of eolian dunes recording
the prevailing wind direction at deposition.

106

Dry phase interdune deposit


Wet phase limestone beds

Damp phase interdune deposit

Depth (m)

10

Dry phase interdune deposit

Damp phase interdune deposit

10

Figure 44: (A) Perspective view of GPR cross sections. (B) 3-D GPR survey showing
map view of major facies; profile locations shown in Figure 40.
107

3.2.10 Discussion
Interdune units commonly form flat-lying regions between eolian dunes and exhibit a
range of geometries varying from hollows and corridors to extensive sheets. Strata of
interdune origin reflect the depositional surface at the time of accumulation and can be
classified into three broad facies associations: dry, damp and wet interdune deposits.
Dry interdune deposits are characterized by well-sorted and mature sediments forming
sets of cross-bedded sandstones a few metres thick, indicative of interdune sedimentation
on a reasonably dry surface. Damp interdune deposits include adhesion structures
characterized by wavy laminated sets formed by the adhesion of wind blown sand grains
to a damp surface, creating a bedding surface covered in low relief warts. A surface is
maintained in damp condition by capillary rise from a water table in the shallow subsurface. These sands commonly contain relatively high amounts of clay and anhydrite
cement. They are characterized by low amplitude radar reflections indicating a clay-rich
lithology rapidly attenuating radar signals. The adhesion of grains in motion to a damp
surface results in the generation of a range of structures including adhesion plane beds,
adhesion ripples (Kocurek and Fielder, 1982), and adhesion warts (Olsen et al., 1989),
which are characterized by low-relief ridges and mound that grow by adhesion to their
upwind edge and thereby undergo upwind migration. Adhesion structures are preserved
both on bedding surfaces and in section where strata form crinkly and wavy laminae. The
generation of adhesion strata requires the accumulation surface to be damp, and such
strata often occur in low-lying damp interdune and dune-flank settings (Hummel and
Kocurek, 1984).

108

Damp interdune deposits display a range of structures that imply the periodic presence of
standing water: wavy laminae, contorted bedding whereas wet interdune facies such as
interdune pond carbonates suggest subaqueous deposition.
Outcrop observations and radar facies from Big Mesa (Moab, Utah) reveal dry phase
eolian cross-beds overlain by clay-rich damp interdune deposits capped by flat-bedded
fresh water carbonates. These imply an upward evolution from a dry depositional phase
evidenced by the well-formed eolian cross-beds (Figure 43A) through a damp phase
recorded as interdune adhesion wavy beds typical of wind ripples in damp interdune
environments (Reading 1996) to wet conditions where fresh water pond carbonates
accumulate as a result of continued rise of the water table.
At the early stage of damp interdune deposition, clay-rich deposits are preferentially
preserved in troughs flanking the dry phase dunes (Figure 43A and B). Damp and wet
interdune deposits usually contain high clay content; in the North Sea Rotliegende
Sandstone, wet interdune deposits are distinguished from dry interdune deposits by their
higher gamma-ray response, reflecting higher clay content (Glennie, 1983; Lahann et al.,
1993). Damp and wet interdune deposits are known to pose permeability challenges in
subsurface eolian reservoirs (Lahann et al., 1993) and ability to predict their spatial
distribution is critical to characterization of eolian reservoirs. As these deposits are
typically caused by wet episodes, thorough observations and documentation of evidence
of wet conditions within eolian deposits may yield vital clues to paleoclimatic changes
during erg deposition. GPR amplitude maps play a vital role in this respect as they reveal
the orientation of dry phase dunes and the clay-rich interdune facies occupying the dune

109

flanks (Figure 44B). The uppermost unit of horizontal reflectors (Figure 43 A and B),
about 1.5 m thick capping the underlying mound-shaped reflector shows that conditions
were increasingly wetter through the deposition of the Navajo interdune deposits
preserved at Big Mesa (Moab, USA); the interdune area was eventually completely
flooded and interdune ponds developed where the limestone beds were precipitated.
Absence of large-scale eolian cross-beds (compare Figure 41A with Figure 41B),and
presence of adhesion wavy beds as well as horizontally-laminated limestone beds suggest
a clear contrast in depositional processes within the Navajo Sandstone at the study site.
Observations at Big Mesa (Moab, Utah) are indicative of eolian dune migration that
occurs synchronously with accumulation in damp, water-table-controlled interdunes
(Kocurek 1993).

25 m

Figure 45: Outcrop photo from Moab (Utah) revealing damp phase interdune deposit
from the Carmell Formation (Image taken from Chan et al., 2007).

110

Similar damp phase structures have been observed at other locations within the western
U.S Jurassic erg deposits and have been described as Harmonic crenulations by Chan et
al., (2007). These structures are seen in outcrops on a bed to multi-bed scale and are
commonly bounded above and below by undeformed units (Figure 45).They typically
display an undulating morphology, occur on tens of centimeter to metre scale and tend to
be laterally extensive.

3.2.11 Implications for Reservoir Development


Eolian sandstones typically form prolific oil and gas reservoirs due to their excellent
sorting; Leman Sandstone of the Permian Rotliegend Group is a very prolific example
(Weber, 1987; Ellis, 1993). Sedimentary structures and associated textural variations
exert an important influence on the porosity and permeability characteristics of eolian
hydrocarbon reservoirs and aquifers (Lupe et al, 1979; Ahlbrandt and Fryberger 1981;
Weber 1987) and much of the heterogeneity in the reservoirs is attributed to damp or wet
interdune depositional processes. Discriminating dry phase interdune deposits from damp
and wet phase interdune facies is very critical to well placement within the dry phase
facies in developing reservoirs formed in interdune environments. Directional and
horizontal wells drilled through the long axes of the cross-bedded dune selectively
targeting dry phase interdune facies constitute perhaps the most efficient means of
developing such reservoirs.

111

3.2.12 Conclusion
This study underscores the potential of GPR imaging as a tool for three-dimensional
stratigraphic studies of ancient eolian deposits. While Jol et al (2003) earlier proved the
viability of GPR imaging technology for detailed stratigraphic studies of large-scale
eolian dunes; this study shows that interdune deposits can be imaged at decimeter scale
resolution although at comparatively shallower penetration depth than on the dry erg
dunes.
Interdune deposits are of interest to eolian sedimentary research for two main reasons.
First, their internal structures have been extensively used although with varying degrees
of success, to distinguish between eolian and subaqueous depositional processes and to
identify the bedform types represented by ancient eolian deposits (e.g. Glennie 1972;
Kocurek and Dott Jr 1981; Rubin and Hunter 1983, 1985; Steele 1983). Second, internal
structures in fossil interdune deposit provide clues about paleoclimatic conditions at the
time of deposition of the dunes.
While radar signals could only target depth of less than 7 metres at the study site, higher
penetration depth could be achieved with lower frequency antennas although with a
consequent loss in resolution. A shielded 100 MHz antenna might provide a practical
compromise to achieving deeper penetration in such a situation without significant loss in
resolution. Within the damp interdune beds occupying the flanks and troughs of the dry
phase dune, sedimentary structures and beddings are hardly visible in outcrop. This study,
however, shows that in such highly altered outcrops with significant obliteration of
primary sedimentary structures such as those studied here; GPR imaging could reveal

112

features that would be otherwise invisible. The revealing findings of Jol et al (2003) and
the interdune architectural details imaged in this test study shows that three-dimensional
GPR imaging of ancient dune and interdune deposit, hold significant promise in
answering many questions about controls on eolian dune and interdune architecture still
unanswered in eolian research.

113

Case Study 3 Dunvegan Formation, Pink Mountain (B.C)

3.3.1 Introduction
Predicting the internal architecture of subsurface clastic deposits, especially deep water
turbidites (Stow and Johansson, 2000, Pyrcz et al, 2005; Deptuck et al, 2008), incised
valley fills (Lericolais, 2001; Chaumillon et al, 2008) and those of fluvial origin (Larue
and Hovadik, 2006; Nichols and Fisher, 2007) has been an area of active research over
the last two decades. Because it is not possible to observe the processes of development
of fluvial architecture over the large spatial and temporal scale involved in fluvial
evolution and preservation of associated sediment, studies of alluvial architecture have
largely been model-based. While qualitative and quantitative models of fluvial
architecture have been insightful in predicting subsurface alluvial architecture (Fielding
and Crane, 1987; Shanley and McCabe 1993; Bridge and Mackey, 1993b; Deutsch and
Tran, 2002), the validity of many of these models is yet to be thoroughly tested on three
dimensional seismic and outcrop data. Three dimensional seismic data has contributed
immensely to our understanding of macro scale architecture of clastic deposits (Brown,
2003; Cartwright and Huuse, 2005) but inability to unambiguously resolve architectural
elements and meso-scale depositional features has been the major limitation of three
dimensional seismic imaging. This limitation of 3D seismics makes outcrop studies the
most applicable methodology of understanding architectural complexities in clastic
reservoirs. Unfortunately, outcrop studies are still largely two-dimensional and this

114

makes it difficult to apply outcrop observations as analogue data for subsurface threedimensional reservoir modelling.
This study evaluates the use of Ground Penetrating Radar imaging on outcrop as a tool
for extending outcrop studies into three dimensions. GPR profiles behind roadside
outcrops near Pink Mountain, British Columbia were recorded and integrated with
outcrop interpretation.

3.3.2 Study Objectives


Most of the earlier Dunvegan studies focused on basin-scale mapping of Dunvegan
deltaic allomembers (Plint, 1996, 1997; Plint and Nummedal, 2000; Plint, 2002); very
few studies examined the valleys incised into the top of the allomembers with an
emphasis on the sedimentology and internal architecture of the valley fills due to
limitations of sparse cores and well spacing (Plint and Wadsworth, 2003 ; Lumsdon-West
et al., 2005); this study however examines the architectural details of the outcrop exposed
near Pink Mountain, British Columbia and evaluates the use of ground penetrating radar
as a tool for future three-dimensional Dunvegan studies.
Specific objectives of this study are:

to demonstrate the importance of decimetre-scale elevation survey for GPR


surveys.

to evaluate the added value of GPR imaging as an aid for future three dimensional
outcrop studies.

115

to document radar facies from the Pink Mountain outcrop as a guide for future
studies; the nature of radar reflections is best evaluated via actual recorded data.
Documented radar facies from ancient fluvio-estuarine deposits are very scanty;
radar facies from the Pink Mountain outcrops will provide an opportunity to
compare radar reflection patterns with lithofacies and architectural elements
observed in similar ancient deposits.

to extend interpretation of the Dunvegan Formation outcrop at the study site based
on an earlier two-dimensional architectural element analysis into 3 dimensions by
integrating outcrop with GPR data.

3.3.3 Dunvegan Formation Introduction


The middle Cenomanian (lowermost Upper Cretaceous) Dunvegan Formation has been
studied extensively in outcrops and subsurface in Northwestern Alberta and Northeastern
British Columbia (Bhattacharya and Walker, 1991b; Bhattacharya, 1994; Gingras et al,
1998; McCarthy et al., 1999; Batten, 2000; Lumsdon-West, 2000; McCarthy 2002). The
Dunvegan Formation is a deltaic complex about 300 m thick which prograded 400 km
parallel to the Cordillera, over a period of about 2 million years. It ranges from 90 - 270
m in thickness and crops out extensively on thrust sheets in the Alberta and British
Columbia Foothills, as well as in the Peace River Valley, where undeformed strata are
exposed in the vicinity of Fort St. John, (British Columbia). Hydrocarbon-bearing
intervals within the Dunvegan Formation are confined to Alberta.

116

3.3.4 Geological Setting


Regional tectonic and eustatic forces played significant roles in Dunvegan deposition.
The Dunvegan Formation was deposited on the western margin of the Cretaceous
Western Interior Seaway of North America which was at the time a retro-arc foreland
basin that experienced over 3 kilometres of subsidence during the Late Jurassic and
Cretaceous (Beaumont, 1981). North America was progressively flooded from the south
by warm saline waters of the ancestral Gulf of Mexico-Tethys and from the north by
cooler, less saline water of the Boreal Ocean during the Cretaceous. These marine
embayments merged just before the AlbianCenomanian boundary (96 Ma) to form the
Western Interior Seaway (Figure 46), which continued to widen until mid-Turonian time
(Williams and Stelck, 1975; Hay et al., 1993; Kauffman and Caldwell, 1993).
The seaway was bounded to the west, by the rising Cordillera, which yielded large
volumes of clastic detritus. To the east, the low-lying cratonic interior of North America
yielded little sediment. The bulk of the sediment forming the Dunvegan Formation was
sourced from northeastem British Columbia and southern Yukon, and was transported to
the southeast by rivers flowing parallel to the active margin of the basin (Stott, 1982).
Additional sediment was introduced by rivers flowing to the east and northeast,
perpendicular to the adjacent Cordillera (Kaulbach, 1995; Plint, 1996). The source of
sediment was related to tectonism in the ancestral Rocky Mountains (Stott, 1982). The
Interior Seaway was not only influenced by regional tectonic processes, eustatic sea level
rise, broadly ascribed to accelerated plate-spreading rates (Kauffman and Caldwell,
1993), is also considered to have been a factor, with Turonian sea level estimated to have

117

peaked at 180 m (Sahagian and Jones, 1993) to 300 m (McDonough and Cross, 1991)
above the present level. A relative sea-level rise at the AlbianCenomanian boundary led
to widespread deposition of marine shales, marking the onset of the Greenhorn Cycle
(Kauffman and Caldwell, 1993).

Figure 46: Paleogeography of the Western Interior Seaway during the middle
Cenomanian including the distribution of the Dunvegan delta complex in Alberta and
British Columbia. (Image modified from Williams and Steck, 1975).

118

3.3.5 Stratigraphy
The Dunvegan Formation is a river-dominated deltaic complex comprising a succession
of alluvial and shallow marine sandstones, siltstones and shales that separate underlying
marine shales of the Shaftesbury Formation from overlying marine shales and minor
sandstones of the Kaskapau Formation (Figure 47). The formation was divided into ten
allomembers (denoted by letters A to J, A being the youngest), broadly comparable to
transgressive-regressive sequences, bounded updip by erosional valleys and interfluve
surfaces and downdip by regionally traceable marine ravinement and flooding surfaces
(Plint, 2000).

Figure 47: Summary of the lithostratigraphic relationships between the Dunvegan


Formation and the underlying Shaftesbury and overlying Kaskapau Formations (Image
taken from Plint and Wadsworth, 2003). Outcrop section in the stratigraphy highlighted
with red circle.

119

This study focuses on the fluvio-estuarine deposit presumed to have formed in an incised
valley cut into Dunvegan Allomember E (Batten, 2000) although direct evidence of an
incised valley was not observed at the study site.
Dunvegan valleys are 0.52 km wide and between 20 to 30 m deep; they are traceable for
up to 20 km in outcrop, and for over 320 km in the subsurface (Plint, 2002). Plint and
Wadsworth (2003) mapped Allomembers J to E using a dense database of subsurface
logs; allomembers D to A were however lumped together as they could not be mapped
individually with available logs and outcrop data.
The valleys are typically filled with stacked fluvial channel deposits separated by fifthorder channel-base erosion surfaces. Individual channel sandstones filling the valleys are
less than 8 m thick, a few tens to about 100 m wide and rarely mappable in the subsurface
due to wide well spacing (McCarthy et al., 1999; Lumsdon-West, 2000; Lumsdon-West
and Plint, 2004).
Dunvegan valleys tend to be channel sandstone-dominated and are usually between 3 and
10 m thick. Some exposures of the sandstone units show horizontal stratification, whereas
others reveal fourth order (based on architectural elements definition in Miall, 1985)
accretion surfaces with dips of up to 15 degrees. Accretion surfaces are sometimes
marked by erosion surfaces that separate cross-bedded sandstone layers and sometimes
are defined by alternating decimeter to metre-scale sandstone and mudstone layers,
defining inclined heterolithic stratification (IHS; Thomas et al., 1987).

120

N
Fort Nelson

97

Chetwynd
97

Dawson Creek

97

Mackenzie

5 Km

Grand Prairie

Figure 48: Map location of Dunvegan Formation outcrop (outcrop location denoted by red circle).
121

Muddy layers often grade down the accretion surface into more massive sandstone;
where IHS is present, it is usually confined to the uppermost storey of the valley fills
(Aitken and Flint, 1996).

3.3.6 Study Location


Roadside exposures of the Dunvegan Formation 26 km southwest of Pink Mountain,
British Columbia (Figure 48) provide an ideal setting for combining outcrop architectural
element study with behind-outcrop georadar imaging. The outcrop is easily accessible,
without overburden and has sufficient lithological contrast considered critical to
generating radar reflections. The outcrop is about 350 m long and about 8 metres high.
Location coordinate of the outcrop location is N56 59 27.0, W12210.2.

3.3.7 Methodology
GPR surveys were conducted with MAL Geoscience Ramac 250 MHz shielded antenna
designed for moderate (5-10 m) penetration depth and sub-metre scale resolution ideal for
resolving intra channel architectural elements and lithofacies. The shielded antenna is
particularly ideal at this Dunvegan exposure as nearby radio towers could have caused
significant signal interference if unshielded antennas were used.
Three GPR profiles were recorded in common offset mode; two lines recorded parallel to
the outcrop and one line in orthogonal direction to the outcrop face (Figure 49). Each line
was re-recorded in reverse direction to detect occurrence of spurious signals not
generated by geological boundaries.

122

N
Townsend Creek

Campground

D
B
C
97

Figure 49: Map showing location of GPR lines at Dunvegan outcrop.

200 m

Acquisition parameters used are:


Antenna separation: 0.36 m
Sampling Frequency: 2760 MHz
No of stacks: 16
Trace Interval: 0.099 m
Velocity: 0.17 m/ns (calculated from depth to outcrop bounding surface observable on the
radar profile). Acquired radar profiles were processed with REFLEX GPR processing
software. With a shielded antenna, few data processing steps were required; data
processing steps are dewow, signal gain and diffraction stack migration.

123

Three radar lines AB, BC and DE were recorded behind the outcrop; lines AB and BC
parallel to the outcrop face were recorded 2 metres behind the outcrop. Line DE is
perpendicular to the outcrop face with D, 35 metres behind the outcrop (Figure 49).
Surface elevation along lines BC and DE is reasonably flat (less than one metre relief);
relief along line AB is significant (more than 5 metres) and the effects on radar reflection
patterns are documented to highlight the importance of elevation surveys to GPR outcrop
studies.
Radar facies are defined based on reflection configuration and amplitude; identified radar
facies and their outcrop equivalents are described below and summarized in Table 6.

High Amplitude Inclined Reflectors


This facies is characterized by high amplitude dipping reflectors each about 20 - 30 cm
thick comprising a unit of about 1 metre in thickness. These reflectors downlap into a
layer of low amplitude horizontal reflectors representing horizontally bedded sandstone
(Figure 50C). This facies is interpreted as lateral-accretion macroforms in the topmost
sand-rich and inclined heterolithic stratification in the more tidally dominated intervals
typical of the middle sandstone unit at the study site.

Low Amplitude Horizontal Reflectors


This facies underlies the high amplitude curved reflectors with high amplitude fill at the
top of the succession interpreted as channel fill sandstone. They are interpreted as mud
beds (facies sh) charactized by low amplitude horizontal reflectors each 10 20 cm thick
(Figure 50C) with indistinguishable reflectors at the base. This reflection pattern was also

124

observed at the Whirlpool Sandstone (Figure 33) and St. Mary River Formation (Figure
67) outcrops and it is indicative of mud beds.

Concave-Upward Reflector with Medium to High Amplitude Fill


This facies is observed mostly incised into the underlying moderate to high amplitude
horizontal reflectors; it is observed in radar lines recorded parallel and perpendicular to
the outcrop (Figure 50C and Figure 53A). It is characterized by a basal concave upward
reflector with internal reflectors interpreted as channels fills; the latter occur at various
scales, the small (less than 10 metres wide) interpreted as chute channel and the wider
ones as main (possibly meandering) channels.

3.3.8 Outcrop Description


Major architectural elements observed in outcrop and radar profiles (Figures 50 and 51)
are described below and annotated on outcrop photomosaic.

Lateral Accretion (LA)


This facies is observed within the topmost unit containing clean channel sandstone. It is
characterized by 20 30 cm thick dipping sandstone beds (dips of 10 22 degrees)
bounded by second order (coset-bounding) surfaces with gradational top and erosive
base; the lower terminations of these surfaces downlap onto the channel floor and often
terminate downdip into massive channel sandstone (Figure 50A). Although the lateral
accretion macroform are inclined bedsets like the inclined heterolithic stratification, they
are predominantly sandstone without the muddy interbeds.

125

Lateral accretion macroforms are typically indicative of meandering channels and their
recognition can be an important first step in paleohydraulic analysis. The height or
thickness of the element has been used as approximate bankfull depth of the channel and
the width is often used as input data for reconstructing channel width (Allen, 1965).

Inclined Heterolithic Stratification (IHS)


This architectural element is a variant of the lateral accretion macroform and is defined
by alternating decimetre- to metre-scale sandstone and mudstone layers with dips of
between 10 and 20 degrees described as inclined heterolithic stratification (Thomas et al.,
1987). They constitute erosive-base, channel-filling units 2 to 3 m thick and few hundred
metres wide that are laterally and vertically interstratified with the cross-bedded
sandstone facies. Muddy layers commonly grade down the accretion surface into more
massive sandstone; this element constitutes a major proportion of the middle unit
observed at the study site.
The predominance of inclined heterolithic stratification in the middle unit (labeled 2 in
Figure 50B) suggests deposition on tidally dominated point bars. A similar style of
stratification has been observed in some ancient point-bar deposits interpreted to have
been tidally influenced (e.g. Thomas et al., 1987; Ainsworth et al., 1994). The relatively
large thickness (decimetresmetres) of the layers comprising the IHS observed in the
Dunvegan valley fills seems to be incompatible with semi-diurnal tidal rhythms and,
instead, is more likely to record a complex response to up and down channel migration of
a turbidity maximum zone in response to episodic (possibly seasonal) changes in the
balance between fluvial discharge and tidal current strength (Allen, 1991).

126

Mud Plug (MP)


This facies is observed only within the heterolithic unit at the studied outcrop; it is
annotated on the outcrop photomosaic as MP (Figure 50B). The only occurrence of this
facies is about 15 m wide and 4 m in thickness; it occurs on the down-dip end of the
adjoining inclined heterolithic stratification unit. Such mud plugs often record the cut-off
of the active meander and an establishment of an oxbow lake.

Channel (CH)
This is observed at the upper unit of the Dunvegan outcrop; the channels range in
dimension from the typical Allomember E channel sandstone (about 50-100 metres wide
and 5 metres thick) to small chute channel deposit (3 metres wide and 2 metres thick).
The channel sandstones are predominantly clean with minimal mud content.

127

Table 6: Radar facies and architectural elements from Dunvegan outcrop (Pink Mountain, BC); radar facies and their outcrop equivalents are highlighted with red arrows.

Radar Facies

Radar Facies

Outcrop Photo
5m

High amplitude
inclined reflectors
dipping between 10 and
20 degrees
downlapping to flatbedded reflectors

Moderately dipping
(200-300) reflectors
concave upward
towards the base

High amplitude
Concave upward
reflector cut into the
underlying horizontal
reflectors

Lithofacies Architectural
Element

Facies Description

Fm, Sp, Sh

Inclined Heterolithic
Stratification (IHS)

Rhythmically-bedded
sandstone and mudstone beds
(each between 0.3-0.5 m thick)
;
beds are moderately dipping
(10-20 degrees)

Sp, Sh, Sl

Lateral accretion
(LA)

Sandstone beds dipping


towards the channel floor with
inclinations from 20 30
degrees; beds do not have the
mud interbeds of IHS.

St, Sp, Sl

Channel cut into


underlying flatbedded sandstone,
most have width:
depth ratio of about
30 (CH)

Sandstone channels; filled with


lateral accretion and sandy
bedforms

Ss, Sp

Mud Plug (MP)

Remnant of abandoned
channel filled with mud

8m

No photo ; profile line orthogonal to the outcrop

Low amplitude
indistinguishable
reflectors bounded by
adjoining IHS unit

2m

128

3.3.9 Integration of GPR and Outcrop


Descriptions of major architectural elements observed in outcrop and radar facies are
summarized in Table 6. Although the major architecturral elements were imaged with
reasonable clarity depsite the surface relief at the outcrop, the relative positions of some
of the architectural elements are distorted.
Radar Profiles AB and BC reveal low-amplitude basal horizontal reflectors overlain by
inclined reflectors with low amplitude toesets (Figure 50C). Incising into this unit at the
midsection of the outcrop is high amplitude concave upward reflector with inclined
reflectors at the margin underlain by the basal low amplitude horizontal reflectors (Figure
50C). Outcrop observations also reveal a three stage depositional history (labeled 1, 2 and
3 in Figure 50B): unit 1, the basal unit dominated by horizontally-bedded sandstone of
low sinuosity possibly fluvial braidplain origin overlain by unit 2, a tidally influenced
meandering IHS-dominated (Inclined Heterolithic Stratification) unit and mudstone. This
fine-grained meandering fluvial style is characteristic of low energy estuarine
environments with significant tidal influence preserved as sand-mud couplets within the
point bars (Thomas et al., 1987); such IHS-dominated fluvial style is common in the
McMurray Formation, host of the gigantic oil sands resource in Alberta also interpreted
as estuarine point bar deposit.
Overlying this unit are unit 3 sand-dominated high sinuosity channel sandstones with
well-developed lateral accretion macroforms. Radar line DE recorded orthogonal to the
outcrop reveals evidence of channel migration (Figure 53A) and preserved lateral
accretion deposits typical of meandering fluvial environments.

129

IHS
CH

LA
IHS

CH

FF

LA

CH

MP

IHS

5m
20 m

CH

IHS

C
LA

IHS

CH

FF

MP IHS

LA
No reflection signals attenuated

No reflection signals attenuated

Figure 50: (A) Uninterpreted outcrop photomosaic (B) Annotated outcrop photomosaic (C) GPR lines recorded parallel to the outcrop (see GPR profile locations in Figure 49).
IHS Inclined Heterolithic Stratification
LA Lateral Accretion
FF - Floodplain Deposit

CH Channel
MP Mud Plug

Ground surface
3rd Order bouding surface
5th Order bounding surface

130

B
A

A
IHS

CH

SB

CH

10 m

FF

CH

CH

IHS
SB

FF

Figure 51: (A) Outcrop photo AB section of Figure 50 (B) Annotated GPR profile (Note the apparent postion of elements CH and IHS at the same depth due to surface relief not corrected for in radar profile).
IHS Inclined Heterolithic Stratification
LA Lateral Accretion
FF - Floodplain Deposit

131

CH Channel
SB Sandy Bedform

2m

Low mud content in the topmost unit is indicative of an active fluvial system and minimal
or no tidal influence as observed in the underlying IHS-dominated unit.
Regional stratigraphy of the Dunvegan Formation suggests that the study area lies in a
proximal delta plain environment and was influenced by accommodation changes
controlled by regional tectonic events (Batten 2000). Observation at the Pink Mountain
outcrop demonstrates an evolution in fluvial style from low sinuosity sheet sandstone
through mud-rich tidally-influenced meandering to migrating sand-bed meandering
fluvial system. This may represent a fluvial response to tectonically driven
accommodation change (possibly modified by eustatic sea level changes).
Change in regional paleoslope is commonly accommodated by river systems by changes
in channel patterns (Schumm, 1993). Wescott (1993) noted that valley slope and channel
sinuosity are directly related such that an increase in regional slope due to tectonic
activity, as was the case during Dunvegan deposition, is typically compensated for by an
increase in river sinuosity. Vertical juxtaposition of fluvial styles at the studied Dunvegan
outcrop could therefore be explained by allogenic tectonic subsidence controlling
accommodation, sediment dispersal and deposition patterns during Dunvegan Formation
deposition. During the deposition of Dunvegan Allomember A, rates of tectonic
subsidence were accelerating, tilting the basin towards the west (Batten 2000); this
increasing subsidence and tilting might have resulted in the observed change in fluvial
style from sand sheets through mud-rich meandering to migrating sand bed meandering
fluvial style.

132

B
LA

SB

CH

LA

IHS

LA
FF

IHS

C
SB

LA

CH

LA

FF

LA

Radar signals attenuated


LA

Figure 52: (A) Enlarged outcrop photomosaic of the sand-bed channels (BC section of Figure 50): third stage of fluvial evolution (labeled 3 in Figure 50B) (B) GPR line behind the outcrop showing major radar facies.

133

3.3.10 Implications for Petroleum Field Development


Paleovalleys may have complex architecture reflecting the depositional controls through
the successive stages of the valley infill. Architectural changes from unchannelized
fluvial depositional style through a heterogeneous tidally dominated unit to channelized
meandering fluvial style could indicate regional tectonic or eustatic control or a
combination of both. Based on outcrop observations and GPR reflection configuration
and attenuation pattern (indicative of mud content), the best reservoir sands are the
stacked high sinuosity sandy meandering channel deposits at the top of the succession
(Figure 50A). Well log correlation typically used for building reservoir models at
conventional well spacing cannot reliably delineate reservoirs of this complexity. Also,
the scale of the channels (width of 50 70 metres; depth of 3-5 metres) imply that their
resolution via seismic imaging requires ultra-high quality three dimensional seismic data
which is usually unavailable. Closely spaced dipmeter logs, in addition to conventional
log suites, could help in mapping the internal architecture of such paleovalleys; the
cleaner sandstone is better targeted with horizontal wells drilled through the long axis of
the meander belt of the sinuous channels guided by analogue data from Dunvegan
outcrop studies.

134

E
Topsoil

LA

SB

LA

CH

CH

FF

Radar signals attenuated

No reflection signals attenuated

IHS

IHS
FF
FF

CH

20 m

CH

Figure 53: (A) Annotated GPR profile of line DE (B) Schematic showimg depositional model at Pink Mountain (SB Sandy Bedform ; IHS Inclined Heterolithic Stratification ; CH Channel ; FF
Floodplain Deposit).
135

3.3.10 Conclusion
This study shows that combining GPR imaging with outcrop mapping is a potent tool in
the study of ancient paleovalley deposits. While surface constraints and unavailability of
survey equipment required for topographic correction did not allow for three-dimensional
survey and resulted in depth errors in profile AB recorded along line AB with significant
relief (Figure 51), this study confirms that a few GPR lines properly oriented relative to
the outcrop can provide architectural details not observable in outcrop and significantly
improve interpretation of outcrop observations. Evidence of migrating high sinuosity
channels as well as channel dimensions are observable on the GPR line recorded
orthogonal to the outcrop.
Although signal penetration depth is an obvious limitation of GPR imaging at most
outcrops, especially those with prevalence of muddy units, antenna frequency in the 100 200 MHz might provide a practical compromise between resolution and penetration
depth. This study also shows the promise of GPR imaging for the field of
paleohydraulics; geometrical data of ancient clastic sandbodies rarely obtainable from
two dimensional outcrop data could become available via three dimensional GPR images
of ancient clastic deposits. Having been successfully applied at the Pink Mountain
outcrop (although with significant limitation due to lack of topographic data required for
elevation correction), GPR imaging can by applied at other Dunvegan outcrops for more
detailed architectural studies.

136

Case Study 4 Shinarump Conglomerate (Utah)

3.4.1 Introduction
This study integrates GPR and outcrop data from an outcrop exposure of the Shinarump
member of the Chinle Formation and evaluates the usefulness of architectural element
analysis combined with Ground Penetrating Radar imaging as an aid for three dimensional
outcrop interpretation. An outcrop of the Shinarump conglomerate was studied, supported by
a GPR line recorded orthogonal to the outcrop.
The Shinarump Conglomerate lies at the base of the Chinle Formation, a fluvio-lacustrine
unit of Late Triassic age (Figure 54) widely exposed in the Colorado Plateau of the
southwestern USA (Lucas, 1993). Deposited in paleovalleys carved in the underlying
Moenkopi Formation, the Shinarump directly overlies the reddish brown soft mudstone
deposits of the Middle Triassic Moenkopi Formation (Figure 58) across a regional
unconformity representing about 17 m.y. (Stewart et al., 1972; Doelling, 1985).
The Shinarump conglomerate has been a long-standing puzzle (Blakey and Gubitosa, 1983;
Dubiel, 1987, 1989, Dubiel et al., 1991); its enigmatic widespread deposition covering over
250,000 km2 across the Colorado Plateau has attracted various explanations. The earliest
attempt to explain the areally extensive deposition of the Shinarump Conglomerates was by
Stokes (1950); he suggested erosion at the top of the Moenkopi unconformity and a
contemporaneous fill of the incised channels with the overlying comglomeratic sandstone
formed via the process of pedimentation.

137

Lithospheric tilting of the U.S Cordillera was also suggested as a possible explanation of
Shinarump Conglomerates widespread deposition (Heller et al., 2003). This study however
focuses on the architectural details of the Shinarump conglomerate at the Hurricane Mesa
outcrop and relates the outcrop observations to the various sedimentary controls postulated
for Shinarump deposition.

Figure 54: Triassic stratigraphy at Hurricane Mesa, Utah (chart taken from Blakey,
2008).

138

3.4.2 Regional Setting and Stratigraphy


The Upper Triassic Shinarump member of the Chinle Formation on the Colorado Plateau
was deposited in a continental back-arc basin located about 5-15 north of the
paleoequator near the west coast of Pangaea (Dickinson, 1981; Pindell and Dewey, 1982;
Ziegler et al.,1983; Parrish and Peterson, 1988; Molina-Garza et al 1991; Bazard and
Butler,1991). A magmatic-volcanic arc on part of the western edge of the Triassic
continent probably provided volcanic ash and clastic sediment to the Chinle depositional
basin (Stewart et al., 1972; Blakey and Gubitosa, 1983; Busby-Spera, 1988), although a
clearly defined source area remains a puzzle (Stewart et al., 1986). Paleogeographic
reconstructions of the Late Triassic Western Interior suggest that the Ancestral Rocky
Mountains in Colorado, the Mogollon Highlands to the south, and a more distant
volcanic-arc complex (Cordilleran Arc) to the west (Figure 55) served as the sediment
source areas for these northwest-trending fluvial systems. Kraus and Middleton (1987)
suggest that the dominant control on Chinle deposition is tectonism. Climate might also
have influenced Chinle deposition, but the relationship between climatic changes and the
chronology of Chinle depositional style is still not very clear. Climatic changes that
occurred during Chinle deposition show a general shift from a sub-humid, seasonal
climate during Shinarump deposition to a more arid climate towards the late stage of
Chinle deposition (Stewart et al., 1972; Blakey and Gubitosa 1983, 1984; Dubiel 1987,
1989; Blodgett 1988; Dubiel et al., 1991; Therrien and Fastovsky 2000; Tanner 2003b;
Prochnow et al., 2006b). The widespread, cross-stratified sandstone and conglomeratic
units of the Shinarump Formation represent the deposit of a complex braided fluvial
system deposited under conditions of rising base level in the lowest parts of the

139

paleovalleys cut into the Moenkopi Formation (Blakey and Gubitosa, 1984; Dubiel, 1983,
1987).

Figure 55: Paleogeographic reconstruction of the Late Triassic within the southwestern
U.S.) The study location is denoted by red circle. (Image taken from Cleveland et al.,
2007).
Blakey and Gubitosa (1984) described several Chinle sandstone bodies in detail in an
effort to understand the basin dynamics controlling the architecture of coarse-grained
deposits in Upper Triassic strata. They described two types of Shinarump deposits:
paleovalley fills and sheet sandstones. Paleovalley fills are thick, multistory sandstone
units filling erosional valleys that developed at the base of Chinle succession. These units
represent aggradational deposition by braided streams in a region with a high gradient
(Blakey and Gubitosa, 1984). Sheet sandstones are relatively thin, commonly only a few

140

meters thick, and are regionally extensive. Blakey and Gubitosa (1984) interpreted these
sandstone bodies to represent braided stream complexes within a broad plain.
(Dubiel, 1987, 1989; Dubiel et. al., 1991) revisited the Shinarump Member in northern
New Mexico and southern Utah and examined its sedimentology in detail. He interpreted
the coarse-grained conglomeratic sandstones at the base of the Shinarump as bedload
deposition in braided streams, and the fine- to coarse-grained sandstones in the upper
Shinarump as suspended and mixed-load deposition in a more sinuous fluvial system.
This study focuses on the basal sheet sandstone of the Shinarump Conglomerate and
integrates outcrop observation with behind-ouctrop GPR imaging.

3.4.3 Study Objectives


The objectives of this study are:

to document radar facies from conglomeratic Shinarump member of the Chinle


Formation; currently there are very few documented radar facies from ancient
conglomeratic braided fluvial deposits. Radar facies from the Shinarump
Sandstone will provide an opportunity to compare radar reflection patterns with
lithofacies and architectural elements observed in other conglomeratic fluvial
sandstone.

to extend interpretation of the Shinarump Sandstone outcrop at Hurricane Mesa


based on two dimensional architectural element analysis into the third dimension
by integrating outcrop with GPR data

141

3.4.4 Study Location


The Shinarump Conglomerate outcrop examined for this study is located at Hurricane
Mesa, southwest Utah. The outcrop is located 10 Km NE of Hurricane, Utah (Figure 56).

Utah

89

15

18

59

Mesa Road

Hurricane
89

59
Kanab

St. George

5 Km
Figure 56: Location of Shinarump Conglomerate outcrop at Hurricane Mesa, Utah. (Outcrop location
denoted by red circle).
142

Mesa road is precipitous towards the outcrop and requires extreme driving care. GPS
coordinate of the Shinarump outcrop studied is 3714'20.45"N; 11312'24.90"W.

3.4.5 Methodology
This study is based on architectural element analysis of the Shinarump outcrop exposed at
Hurricane Mesa (Utah, USA) and a GPR line (broken line EF; E is 10 metres behind the
outcrop) recorded behind outcrop (Figure 57). GPR survey was conducted with MAL
Geoscience Ramac 250 MHz shielded antenna designed for moderate (5-10 m)
penetration depth and sub-metre scale resolution ideal for resolving intra channel and
channel-scale architectural elements. The major limitation of this study is inaccessibility
to outcrops on nearby private land and inability to acquire GPR data along lines of
varying elevations due to unavailability of differential GPS or survey equipment; hence,
only one GPR profile was recorded in common offset mode (in orthogonal direction to
the outcrop face) along a flat surface behind the outcrop.
Acquisition parameters are:
Antenna separation: 0.36 m
Sampling Frequency: 2760 MHz
No of stacks: 16
Trace Interval: 0.05 m
Velocity: 0. 10 m/ns (calculated from depth to known reflector)
Acquired radar profile was processed with REFLEX GPR processing software. Data
processing steps include de-wow, signal gain and diffraction stack migration.

143

Mesa Road

E
Hurricane Mesa
Testing Facility

F
Mesa Road

100 m
Figure 57: Location of Shinarump Conglomerate outcrop (red circle) at Hurricane Mesa,
Utah (USA) showing EF GPR line.

Outcrop Description
Outcrop observation at the Hurricane Mesa Shinarump conglomerate reveals two distinct
architectural features: a lower succession of horizontally bedded sandstone and minor
planar cross-bedded units with thin mudstone interbeds and an upper channelized unit
consisting of pebbly to coarse-grained planar cross-bedded sandstone. Individual outcrop
units (Figure 58) are described below:

144

Unit 1
Medium-grained planar bedded sandstone with coarse-to-pebble size sands at the base
and medium scale planar cross-beds towards the top of the unit. Dominant lithofacies is
plane bed horizontal stratification with planar tabular cross stratification toward the top of
the unit. Some petrified logs are present. This unit rests on the basal Chinle
Unconformity.
Unit 2
This unit is dominated by fine-grained planar cross-bedded sandstone with ripplelaminated sands towards the top. A few pebbly sandstones are also found in this unit.
Unit 3
This unit comprises rippled silty and fine grained sand interbeds about 1 to 3 metres
thick. Individual beds have sharp, planar upper and lower contact. The beds are laterally
extensive with sharp planar basal contact and a concave upward erosive top. This unit is
truncated at the top by overlying channels.
Unit 4
This unit comprises coarse to pebble-size channel sandstone with medium scale planar
cross-beds. Each bed is about 1 to 3 metres thick. Lenses of coarse-grained sands are
interspersed in this unit as well as preserved petrified wood.
Unit 5
This unit consists of medium-grained trough cross-bedded channel sandstone with planar
beds at the top of the unit. Lenses of pebbly sandstone are also found in this unit.

145

Unit 6
This unit consists of medium to coarse-grained trough cross-bedded channel sandstone
overlying pebble size sandstone unit. This unit is predominantly coarse sand with
minimal mud content.

Three architectural elements based on Miall (1985) fluvial architectural element


classification scheme were identified at the Shinarump sandstone outcrop; recognition of
architectural elements is based on facies associations, tracing of bounding surfaces and
bedform interpretation in the context of local paleoflow. Identified architectural elements
are described below:

Element LS (Laminated Sand Sheets)


Element LS comprises sheets of laminated sand with rippled and planar cross-bedded
sandstone. This element is found in many ancient fluvial successions and is interpreted as
product of flash floods deposited under upper flow regime plane bed conditions. The
sandstone beds are fine to medium grained and about 1-3 m thick; mudstone beds are
thinner and light grayish in color. Element LS is dominated by rippled sandstone towards
the top indicating waning flow after flood event.

Element CH (Channel)
Element CH is the preserved Shinarump Conglomerate sandstone deposited in channels
incised into the underlying element LS. The channel sandstone observed in outcrop
(Element CH in Figure 58) is 4-6 metres thick comprising mostly sandy bedforms and

146

thin pulses of gravelly sandstone. Channel width obtained from GPR profiles suggest
narrow width (40 -50 metres).

Element SB (Sandy Bedforms)


This lithofacies assemblage dominates the middle to upper part of the Shinarump
conglomerate; it comprises planar-tabular cross-stratified sandstone units each 1-3 metres
thick. Architectural element is predominantly Sandy Bedforms with a varying assemblage
of planar and trough cross-beds. This element represents intra-channel bedforms that
accumulated by predominantly vertical aggradation.

147

A 62

B
Figure 36: Section of GPR line EF from Hurricane Mesa, Utah highlighting major radar facies

20 m

2m

Moenkopi Shale
A
B
Unit 6

SB

St

SB

Unit 5

SB

SB

Unit 4

Unit 3

CH

LS
LS

Mean Paleoflow direction - 248

Unit 2

Sp

Shinarump-Moenkopi Unconformity

Unit 1
3rd order surface

5th order surface

Figure 58: Photomosaic of the Shinarump conglomerate outcrop at Hurricane Mesa (Utah, USA) photo taken looking northeast.

148

20 m
6th order surface

Remnants of petrified wood are also preserved within the sand sheets observed at the mid
section of the outcrop.

Radar Facies
The radar line reveals four radar facies (Figure 59): concave upward reflectors (Element
CH) with a fill of high amplitude horizontal reflectors (Element SB) underlain by low
amplitude planar reflectors (Element LS), beneath which radar signals are rapidly
attenuated (Moenkopi Shale). Horizontal intra-channel reflectors are interpreted as
horizontally-bedded coarse-grained Shinarump sheet sandstone.
A

B
Sandy Bedform

Shinarump-Moenkopi

Basal sandstonemudstone interbeds


Channel Base

Figure 59: (A) Uninterpreted black and white GPR section of GPR line EF from Hurricane
Utah (B) Color section of GPR line in (A) annotated with main depositional features.
149

The concave upward surfaces are Shinarump channel bases incised into interbedded
horizontal laminated sandstone-mudstone interbeds blanketing the Shinarump-Moenkopi
unconformity (Figure 58).

3.4.6 Discussion
Outcrop observation and GPR data at the Shinarump Conglomerate outcrop suggest
deposition in a sandy braided fluvial environment. The Shinarump channels were incised
into laminated sheet sandstones interpreted as flood stage deposits.
The lithofacies assemblages and architectural-element characteristics including lack of
lateral accretion, lack of preserved muddy facies, presence of coarse to pebble-grained
sand sheets, deposition by vertical aggradation, observed in this study (both in outcrop
and radar profile) suggests that the Shinarump Conglomerate was deposited in low
sinuosity proximal conglomeratic sand bed river. At the outcrop, element SB
predominates, with lithofacies Sp (sandy planar cross-beds) and St (sandy trough crossbeds) present in varying assemblage and comprising planar sand sheets separated by
minor internal erosional scour surfaces. Observed channel fill is about 3-4 metres thick
with small width (about 60 m wide); channel margins are hardly identifiable in outcrop;
they are seen in the GPR profile recorded orthogonal to the outcrop. The Shinarump
outcrops reveal stacked multistory sandsheets with little preserved overbank fines typical
of the Platte fluvial style.
The exact climatic conditions under which the Shinarump was deposited have been the
subject of much discussion, and no conclusion appears to be generally accepted. Regional
climate during the deposition of the Chinle Formation is interpreted to have been
subtropical monsoonal (Dubiel, 1987; Dubiel et al., 1991), although paleosol data at other
150

3m

Figure 60: (A) Combined GPR profile and outcrop Photomosaic at the Shinarump conglomerate outcrop (Hurricane Mesa, Utah); paleocurrent data from Hurlbert, 1995 (B) Depositional model built from (A)

151

Chinle Formation outcrops imply semiarid conditions during some intervals (Prochnow et
al., 2006). An arid or semi-arid climate however seems improbable at the study site;
abundance of petrified wood found preserved in the lower and mid section of the outcrop
suggests a climate suitable for the growth of large trees.

3.4.7 Conclusion
This study underscores the value of GPR imaging in outcrop studies. Outcrop observation
and GPR data at the Shinarump Conglomerate outcrop suggest a braided fluvial system
with small channels filled with vertically aggraded sheet sandstone. Shinarump channel
dimensions (about 50 metres wide; 3 metres deep) and architecture are clearly defined by
integrating outcrop observations with GPR imaging. Although only one GPR line was
recorded at Hurricane Mesa outcrop due to inavailability of survey equipment for
topographic correction of GPR data and restricted access, GPR data obtained shows that
future GPR imaging at the study site holds immense potential in unraveling detailed
architecture of the Shinarump Conglomerate.

152

Case Study 5 St. Mary River Formation (Alberta)

3.5.1 Introduction
This study evaluates the usefulness of Ground Penetrating Radar as an aid for three
dimensional outcrop study on the outcrop exposures of St. Mary River Formation at
Oldman River, 7 km west of Monarch (southern Alberta). St Mary River Formation is
one of the few ancient clastic deposits interpreted as an anastomosing fluvial deposit
(Nadon, 1994). Current facies model for anastomosing rivers are based mostly on
actualistic studies of modern rivers propagated by Smith (1983, 1986), Smith and Smith
(1980) and augmented by Trnqvist et al., (1993) and Weerts and Bierkins (1993). These
workers and recently North et al., (2007) have distilled the variability in their
observations of modern anastomosing rivers and constructed facies models to provide a
means of interpreting anastomosed fluvial deposits in the stratigraphic record.
Anastomosing rivers have been less studied than their meandering and braided
counterparts in modern environment and more so in ancient deposits; this explains the
gap in our understanding of the architecture of ancient anastomosing fluvial deposits.
North et al (2007) warn that since current geomorphological knowledge of anabranching
systems is incomplete, it logically follows that knowledge of the sedimentological record
created by anastomosing rivers is even more incomplete. Hence, sedimentological
interpretation of ancient anastomosing rivers is much less certain and considerably more
varied than implied by published facies models and caution should therefore be exercised
in applying those models.

153

To this end, outcrop studies of ancient deposits of interpreted anastomosing origin aided
by ground penetrating radar imaging holds significant promise in expanding our
understanding of depositional processes in ancient anastomosing rivers and their
preserved products.
This study documents outcrop observations of facies associations as well as stacking
patterns of architectural elements and evaluates the potential for detailed three
dimensional outcrop studies using ground penetrating radar imaging technology at St.
Mary River Formation outcrops in southern Alberta and similar ancient anastomosing
fluvial deposits. Although St. Mary River Formation is exposed extensively in southern
Alberta, many of the outcrops are covered by thick relict Pleistocene glacial deposit
making it difficult to run GPR profiles on the outcrops.

3.5.2 Regional Stratigraphy


The sediment of the St. Mary River Formation was deposited as an eastward-thinning
clastic wedge that recorded the progradation of continental conditions linked to the Late
Cretaceous southeast retreat of the Bearpaw Sea within the Alberta foreland basin.
Sediment source was principally from the west and depositional environments of the
aggrading section were predominantly fluvial (Figure 61). The St. Mary River Formation
is a non-marine unit about 230 m thick observable as extensive exposures in the river
valleys of the Plains and in the thrusted strata of the Foothills of the Rocky Mountains in
southwestern Alberta (Nadon, 1994). Palynomorph data indicate that the St. Mary River
Formation is uppermost Campanian to upper Maastrichtian in age (Nadon, 1994). The St.
Mary River Formation overlies marine shoreface sandstones and brackish water shales of
the Blood Reserve Formation (Figure 62) and underlies the fluvial redbeds of the Willow

154

Creek Formation (Nadon, 1991). The formation consists of medium-grained lenticular


channel sandstone surrounded by thin sheets of sandstone, siltstone and mudstone.

British
Columbia

Saskatchewan

Alberta

Manitoba

200 Km

Sediment transport direction

Mountains
Uplands

Lacustrine (schematic)
Coal swamp (schematic)

Transgression-regression direction
Fluvial coastal plain
Barrier islands
Seaway

Figure 61: St. Mary River Formation Upper Cretaceous Paleogeography (Image taken
from Dawson et al, 1994).

155

Petrographic evidence suggests that the main sediment sources for the St. Mary River
Formation were Upper Paleozoic sediments and older Mesozoic rocks of the then rising
Ancestral Rocky Mountains to the west (Rahmani and Schmidt, 1982).

Lithology Code
Clastics

Shale

Carbonates

Figure 62: Alberta Cretaceous Stratigraphic chart showing St. Mary River Formation.
(Chart taken from Alberta Energy Resource Conservation Board Table of Formations,
2009).

156

3.5.3 Study Objectives


The objectives of this study are:

to document radar facies from sandy ribbon-shaped channels of St. Mary River
Formation; currently there is no documented radar facies from ancient
anastomosing fluvial deposits. Radar facies from the St. Mary River Formation
will provide an opportunity to compare radar reflection patterns with lithofacies
and architectural elements observed in other ribbon fluvial sandstone.

to create a depositional model of the St. Mary River formation at the study site by
integrating outcrop and GPR data

evaluation of added value of GPR imaging as an aid for three dimensional outcrop
studies.

Study Location
The location for this study is by the Canadian Pacific railway, about 7 Km northwest of
Monarch in Southern Alberta (Figure 63A). The location can be reached via Range Road
244 off Highway 519 (Figure 63B). GPS location of the study site is N49 49 08.1 W113
12 09.1.

3.5.4 Methodology
GPR survey was conducted with MAL Geoscience Ramac 250 MHz shielded antenna
designed for moderate (5-15 m) penetration depth and decimetre-metre scale resolution
ideal for resolving intra channel and channel-scale architectural elements. Two GPR
profiles were recorded (Figure 64) in common offset mode (1 parallel and 1 in orthogonal
direction to the outcrop). Acquisition parameters are:

157

Claresholm

Alberta

520

Barons

23

Edmonton
519
Granum
Calgary
Monarch
5 Km
200 Km

3A

519

B
Range Road 243

Range Road 244

Township Road 104

Canadian pacific rail

500 m

Figure 63: (A) Location of study outcrop in Alberta (B) Detailed outcrop location. (Outcrop location
highlighted in red).
158

Antenna separation: 0.36 m


Sampling Frequency: 2760 MHz
No of stacks: 16
Trace Interval: 0.05 m
Velocity: 0. 16 m/ns (calculated from depth to outcrop bounding surface observable on
the radar profile)
Line AB has significant relief (about 3 metres); relief along line CD is minimal (less than
1 metre). Elevation correction was not performed on the acquired data due to lack of
survey equipment.
Acquired radar profiles were processed with REFLEX GPR processing software. Data
processing steps include de-wow, signal gain, background filtering and diffraction stack
migration.

Facies Description
Facies Assemblage 1: Channel Sandstone
Facies assemblage 1 consists of northeast-southwest trending ribbon-shaped channel
sandstones. These channels are either simple or have at the most two stories; individual
channel sandstones are about 3-5 m thick and 60-90 m wide. The basal contacts vary
from erosional to sharp and flat or slightly undulatory at the margins. Above the erosional
base is a lag, a few decimeters to half a metre thick, composed of shale/ siltstone
intraclasts. The overlying sandstone varies from fine- to medium-grained and is mostly
trough cross-bedded sandstone at or near the base (Figure 65) and in some cases planartabular cross-beds toward the top. Multistory channels are commonly separated by half

159

Range Road 244

Range Road 243

519

Canadian pacific rail

Range Road 244

Range Road 243

500 m

Ol
an
dm

Ri
r
ve

Canadian pacific rail

D
100 m

Figure 64: Location of study outcrop and GPR lines. (

160

Outcrop exposure)

metre to a metre layer of mudstone. The channel lenses are capped by rippled siltstone to
very fine sandstone.

Figure 65: Outcrop photo showing St. Mary River Formation trough-crossbedded channel sandstone. (Photo taken looking north).
Facies Assemblage 2 - Small Sandstone Lenses
Sandstone lenses of Facies Assemlage 2 are observed between the northeast-southwest
oriented main channels. These smaller lenses are composed of fine to medium sandstone
with sharp, planar bases with no basal lag (Figure 66). Where paleocurrent indicators are
observable, they are generally perpendicular to the main channel (facies assemblage 1)
trend. Facies Assemblage 2 lenses are interpreted as crevasse splays. The small scale of
these sandstone bodies and their general orientation perpendicular to the thicker channel

161

sandstone lenses implies that these channels were conduits delivering sediment from the
main channels onto the floodplain.

2m

Figure 66: Outcrop photo showing St. Mary River Formation crevasse splay sandstone
(photo taken looking north).
Facies Assemblage 3 Overbank Fines
Facies Association 3 is a mud-dominated recessive unit typically found underlying the
channels of facies assemblage 1. The mudstone is in some cases interbedded with thin
sandstone beds interpreted as overbank levees. Paleosols are also preserved in this unit
with thickness varying from less than 20 cm to about a metre. Internal structures are
seldom visible because of the combination of grain-size uniformity, rooting, and
bioturbation. This facies assemblage represents environments on the floodplain of the St.
Mary river system.

162

Radar Facies
Radar reflections from GPR line along and orthogonal to one of the channels reveal a few
radar facies described below:
Concave upward reflector with high amplitude fill
This facies consists of concave upward reflectors about 55 metres wide and 3 metres
thick with high amplitude dipping and horizontal internal reflectors. The dipping
reflectors are interpreted as lateral accretion macroforms (LA) based on their scale and
dip into the channel (Figure 67).This radar facies is interpreted as the ribbon-shaped
channels of St. Mary River Formation with evidence of lateral accretion and vertical
aggradation.

SB

LA

FF

Figure 67: Radar image of St.Mary River Formation ribbon channel sandstone (Line CD);
see location of GPR line in Figure 64.
Horizontal Reflectors
High amplitude continuous horizontal reflectors between the concave upward reflectors
(channels) are interpreted as crevasse splays; they are characterized by high amplitude
upper reflectors and low amplitude basal reflectors (figure 68) indicative of crevasse
sands deposited on underlying muddy floodplain lithology.

163

CH

Figure 68 (A) Radar profile CD showing ribbon channels and adjoining horizontal
reflectors (B) Radar profile CD showing horizontal reflectors interpreted as crevasse
splay deposits
These deposits are expressed geomorphologically as flat surfaces adjoining moundshaped surface topography characteristic of the St. Mary River Formation channels.
Radar profile AB recorded along the channel axis contains a few slightly dipping
reflectors underlain by horizontal low reflectors interpreted as floodplain mudstone.

164

B
SB
FF

C
LA

SB

CH
FF

Figure 69: (A) Interpreted Radar profile AB (B) Uninterpreted profile AB in greyscale (C) Profile CD
annotated with major interpreted architectural elements (D) Uninterpreted profile CD ingreyscale. (LA
Lateral accretion, CH Channel, SB Sandy Bedforms, FF Floodplain mudstone)
165

3.5.5 Discussion
The overall facies distribution of the St. Mary River Formation at the study site near
Monarch Southern Alberta) shows that the unit is dominated by northeast-southwest
trending channels separated by either recessive inter-channel island mudstone (Figure 70)
or by horizontally-bedded crevasse splay deposits. The channels have ribbon geometry
with width: depth ratio of less than 25. Textural trends in the crevasse-splay sediments
are consistent with those described from Holocene anastomosed deposits (Weerts and
Bierkens, 1993). The geometry and internal structures of the main channel observed in
outcrop photos and radar profile show that they are products of vertical aggradation with
minor evidence of lateral accretion and possible downstream accretion.
Splay channels suggest dominance of vertical aggradation noted also in both modern and
Holocene studies (Smith and Smith 1983, 1986; Trnqvist et al., 1993). Absence of coals
within the St. Mary River Formation is probably due to repeated inundation of the
floodplain with clastic detritus. Presence of lateral accretion indicates sinuosity in St.
Mary River channels as also observed from modern rivers (Bridge 2006). Overall
observations from St. Mary River Formation suggest an anastomosed fluvial system
characterized by ribbon sandstones encased in finer floodplain sediments (figure 70).
North et al (2007) however, cautioned that single thread, non anabranching rivers could
create a similar sedimentary record in a setting where accommodation is large
(subsidence rate is high), avulsion frequent, or the sediment supply to the river system is
mud-rich; as might occur on a large-scale fluvial fan entering a foreland basin envisaged
as the context for the St. Mary River Formation.

166

While low width-depth ratio (less than 20) is typical of the St. Mary River Formation
channels observed in this study, a regional study of the dimensions of the St. Mary River
Formation channels utilizing GPR imaging might reveal an interesting trend in channel
dimension with possible correlation with syn-depositional regional tectonic processes.
Regional studies of other ancient fluvial deposits similar to St. Mary River Formation
aided by ground penetrating radar could help reconcile observations from modern
environments with those from ancient deposits.

167

3.5.6 Conclusion
Although thick bed of glacial drift on many St. Mary Formation outcrops in Southern
Alberta as well as outcrop accessibility posed significant difficulty to finding ideal sites
for GPR surveys, where the overburden is thin, GPR imaging is a promising technology
for three dimensional outcrop studies.

Ribbon channels

Crevasse splay

Muddy Inter-channel Island

Figure 70: Block diagram depicting depositional model of St. Mary River Formation at
the study site.

168

At the study outcrop, two GPR lines, one parallel and orthogonal to the St. Mary River
lenticular channels reveal architectural elements and reflection patterns typical of
anastomosing rivers. Lithological contrasts between channel fills, mud-rich crevasse
splays and inter-channel floodplain typical of anastomosing deposits indicate that other
ancient anastomosing fluvial deposit could be possible targets for GPR imaging. With
sub-metre accuracy survey equipment, other St. Mary Formation outcrop sites with
uneven surface topography could be imaged and a more regional study of the sandstone
conducted to understand regional depositional trends and how this correlates with
allogenic controls.
Low penetration depth is the major limitation of Ground Penetrating Radar at the study
site; lower frequency antennas (100 and 200 MHz) might improve signal penetration
without significant loss in resolution.

169

Chapter 4
Summary, conclusion and recommendation for future research

This study represents the first extensive test of the application of the GPR technique to
ancient clastic units under a wide range of field and sedimentological conditions. Based
on the high resolution data observed in GPR profiles from various outcrops studied in the
course of this research, there is no doubt that contrary to a widely held notion, GPR
imaging technology works on outcrop and it will play a pivotal role in three dimensional
outcrop studies as analogues for subsurface reservoirs. The clarity with which channel
and intra-channel scale features are revealed at the studied outcrops suggests that lack of
success with previous studies is a result of inappropriate choice of GPR instrumentation,
inefficient data acquisition and poorly informed data processing steps. This research
confirms the usefulness of outcrop-based GPR imaging as a tool for refining outcropbased interpretation of processes from ancient depositional environments and as an
imaging technology with the potential to revolutionize outcrop studies in the way
reflection seismology changed basin analysis.
This study as well as analysis of previous studies involving application of GPR imaging
technology on sediments and rocks shows that data obtained with shielded antennas are
less prone to extraneous noise and require less rigorous data processing than those
obtained with unshielded antennas. While most GPR surveys are prone to diffraction,
often seen as hyperbolas in radar profiles, these must be carefully removed by appropriate
migration algorithm. Diffraction migration algorithm worked quite well on data acquired
for this study but this requires a fairly accurate knowledge of the velocity profile through
170

the rocks being imaged; common-mid-point surveys or actual measurement of the


dielectric properties of the lithologies at the outcrop might be required to correctly
estimate velocity.
Topographic survey of the surface where the GPR mapping is conducted is a compulsory
prerequisite for all outcrop-based GPR studies; it helps prevent artificial structures
created in the profile by uneven surface elevation. Other than shallower depth of
penetration at shale-rich outcrops, GPR imaging works on all formations surveyed
including those with minimal lithological heterogeneity (intra-channel reflectors at the
Whirlpool Formation and Shinarump Conglomerate outcrops) widely considered as
necessary for providing electromagnetic impedance. With the 250 MHz antenna used for
this study, individual bedding units and sedimentary structures were clearly imaged;
higher resolution and deeper penetration can be achieved by using a range of antenna
frequency for future studies.
Seismic imaging technology has been one of the most potent tools that advanced our
understanding of regional and reservoir scale architecture of clastic deposits and the
results of its evolution into three dimensional imaging and visualization realm in the last
decade have been magical. Developments in three-dimensional seismic acquisition,
processing and data visualization as well as advances in seismic attribute analysis have
significantly reduced exploration risks and dry holes over the years (Brown, 2003) but the
resolution of conventional seismic data limits its applicability in detailed reservoir
characterization which is increasingly required in many primary field development
projects and enhanced oil recovery schemes. This resolution gap especially in predicting
inter-well permeability beyond seismic resolution will continue to drive three-

171

dimensional outcrop research as analogue data for the subsurface as well as serving as a
predictive tool for reservoir characterization.
Amongst the myriads of technology available to date, Ground Penetrating Radar imaging
technology holds the greatest promise for detailed three-dimensional studies. GPR
imaging, in combination with detailed outcrop sedimentology and architectural element
analysis constitute a powerful set of tools.GPR is currently the only high resolution, submetre non-destructive, non-invasive and relatively inexpensive penetrative imaging
technology available for outcrop studies and it has the potential to transform outcrop
studies, refine existing facies models and provide more realistic input for quantitative
modelling of reservoir architecture and heterogeneities. Although the current cost of GPR
equipment is still beyond the reach of an average outcrop researcher, the costs are
significantly cheaper than seismic data acquisition.
The high resolution nature of GPR data makes it susceptible to image degradation due to
subtle changes in topography; this makes detailed elevation survey a critical requirement
for GPR surveys. Equipment for sub-metre scale topographic surveys (differential GPS,
Total Station) are quite expensive and will likely keep GPR research only within the
reach of research institutions and companies. Because GPR data interpretation depends
significantly on the data processing steps employed; more involvement of geophysicists
in projects involving three-dimensional GPR studies is required, especially as GPR data
processing, GPR attribute analysis and visualization get increasingly advanced.
GPR imaging will significantly improve studies on sand body architecture. Attempts to
create statistical distribution of sediment-body geometry based on outcrop observations
(Fielding and Crane 1987; Dreyer, 1990; Gibling, 2006) have been criticized for inability

172

to provide unambiguous determinations of the 3-D geometry and orientation of channels,


channel belts and other sand bodies in most outcrops due to limited and often two
dimensional exposures (Bridge, 2006). More accurate sand body dimensions can now be
obtained in outcrops via GPR imaging as demonstrated clearly in this study.
Advances in GPR data acquisition, data processing and visualization still trail similar
developments in seismic imaging due to less interest in GPR technology by the petroleum
industry that has funded much of the advances in seismic data acquisition, processing and
three dimensional visualization; however, the current growing interest in GPR imaging
technology is stimulating rapid advances in GPR processing as well as data display and
the trend will likely continue.
The recent development and application of Seismic Geomorphology to exploration and
field development was triggered by the advent of high-quality and increasingly more
affordable and widespread 3D seismic data currently available. 3D seismic data affords
plan-view images of depositional elements and in some instances entire depositional
systems. Attribute analyses of such images have significantly enhanced predictions of the
spatial and temporal distribution of subsurface lithology (reservoir, source, and seal),
compartmentalization, and stratigraphic trapping capabilities, as well as enhanced
understanding of process sedimentology and sequence stratigraphy (Posamentier and
Kolla, 2003). Combining vertical and plan view images via integrated seismic sections
and stratal slices have aided extraction of geological insights and geomorphologic
analyses of depositional systems from three-dimensional seismic data. This approach has
also aided quantitative assessment of paleogeomorphic features observable in subsurface
reservoirs; such attributes as channel sinuosity, slope, width-thickness ratios, as well as

173

measures of channel bifurcation, internal architecture and distribution of architectural


elements in various clastic deposits are now readily measured.
In contrast to advances in three-dimensional seismic imaging, three dimensional full
resolution Ground Penetrating Radar surveys on outcrops are rare; this is mostly due to
very close line spacing required for unaliased recording of reflections and diffractions
coupled with 3D migration processing. Applying the Nyquist criterion requires 3D GPR
measurements to be spaced in all horizontal directions near a quarter-wavelength of the
highest signal and noise frequency content, because diffractions are hyperboloids.
The well-known half-wavelength spatial sampling formula (e.g., Sheriff and Geldart,
1995), valid for a seismic spread with one source and multiple receivers, often leads to
spatially aliased 3D data sets when applied to fixed-offset, single source-and-receiver
GPR surveys. Theoretical line spacing for 250 MHz antenna used in this study is about
10 cm; such close line spacing will significantly increase data acquisition time and
greatly increase the field cost for large-scale three-dimensional surveys; this explains
why full resolution three-dimensional GPR surveys are rarely conducted. Lower
frequency antennas require wider line spacing but even the lowest GPR antenna (12.5
MHz) will need less than 5 metre spacing, which still requires huge data acquisition time
for channel-belt-scale surveys. Due to the practical limitations of adhering to the
theoretical line spacing for GPR surveys, most surveys use coarser line spacing aided
with the knowledge of the outcrop stratigraphy and sedimentology.
Despite the limitations that line spacing imposes on outcrop GPR imaging; full resolution
three-dimensional GPR surveys hold significant promise for high-resolution outcrop
studies (Beres et al, 1995; Jol, 2003; Heinz and Aigner, 2003) and with refinement in

174

GPR data processing and visualization could herald the field of radar geomorphology and
provide outcrop-based data for paleohydraulics and paleoclimatic studies. Results of the
three-dimensional GPR survey at the Navajo Sandstone outcrop show that depositional
trends and facies distribution are better evaluated with time slices, as GPR amplitude
maps are often indicative of lithological trends (Figure 71).

Dry phase
interdune facies

Damp phase
interdune facies

10

Figure 71: Three-dimensional GPR data display from Navajo Sandstone interdune
deposit at Big Mesa, Utah (USA) showing low amplitude damp interdune facies
preserved in the dry phase dune flanks.

175

Where three-dimensional GPR data acquisition is not feasible, this study shows that twodimensional GPR lines integrated with outcrop data could provide architectural details at
the resolution required for reliable three-dimensional interpretation.
As observed from the outcrop locations studied in this research; significant signal
attenuation and low penetration depth are the primary limitations of GPR imaging
technology to outcrop studies. There is virtually nothing that can be done to reduce signal
attenuation due to shale content and the presence of diagenetic clay in the formation
being studied or overlying formations. Deeper signal penetration can be achieved by
using lower frequency antenna (12.5 MHz to 100 MHz) for GPR surveys, although lower
frequency antennas are usually bulkier which implies longer data acquisition time and in
some cases, additional survey crew. Resolution will also be sacrificed with lower
frequency antennas but the lower end (100MHz and 200MHz) of the high frequency
antennas is still ideal for resolving architectural elements and macroforms.
The ultimate goal of outcrop-based GPR studies is to obtain analog data on architecture
and distribution of lithological heterogeneities from ancient clastic deposits as analogue
data for subsurface aquifers and reservoirs. Future outcrop studies could utilize hand-held
gamma ray scintillometers (used for detecting natural radioactivity as proxy for clay
content) for a quantitative correlation of GPR reflection amplitudes with lithology. Probe
permeameters could also be used to correlate reflection amplitude with variation in
permeability.
In conclusion, four major findings from this research that are of immense significance to
outcrop studies are summarized below:

176

1. All the case studies show that clastic deposits especially of eolian, fluvial and
estuarine origing are significantly heterogeneous even at outcrop scale. This
implies that petroleum reservoir modeling using wells spaced several kilometers
apart can rarely yield a reliable prediction of reservoir properties in such
heterogeneous clastic deposits.
2. Observations from all the case studies suggest that sedimentological and
stratigraphic data (especially of decimeter-scale outcrop features) can be obtained
from outcrop-based GPR profiles; architectural elements and lithofacies observed
in outcrop were unequivocally correlated to their radar equivalent. This suggests
that the failure of outcrop GPR imaging in earlier studies is not due to diagenetic
overprints but inappropriate choice of GPR equipment, ill-informed data acquition
and processing steps. Appropriate choice of GPR instrumentation (with proper
attention paid to the choice of antenna frequency and antenna shielding) is
therefore critical to the success of GPR surveys, subsequent data processing and
geological interpretation. It is important to understand the mode of operations of
the various GPR products and their appropriateness for specific problems
addressed by the intended outcrop research.
3. Amplitude variation and reflection configuration of GPR images are closely
linked with lithology as well as depositional environment and they are vital
interpretive features critical to outcrop-based research. They are quite
instrumental for identifying architectural elements as well as revealing sandbody
geometry and internal architecture. Interpretation of radar images should however
be ground-truthed by outcrop observations.

177

4. GPR data degrading due to poorly collected data can rarely be fixed during data
processing; hence, adhering to best practices of GPR data collection spelt out in
this thesis is critical to the success of GPR surveys. In addition, topographic
survey data must accompany GPR surveys as these are required for static
correction during the data processing phase.
5. Properly acquired GPR data require less vigorous data processing; few processing
steps such as dewow, signal gain, elevation correction and migration are sufficient
in most cases. Whereever possible, shielded antenna should be used for surveys as
data collected with unshielded antenna are more prone to extraneous noise.

178

References
Ahlbrandt, T. S., and S. G. Fryberger, 1980, Eolian Deposits in the Nebraska Sand Hills:
U.S. Geological Survey, Professional Paper 1120-A, 24 p.
Ahlbrandt, T. S., and S. G. Fryberger, 1981, Sedimentary features and significance of
interdune deposits. Recent and ancient nonmarine depositional environments: models for
exploration, p 293-314
Aigner, T., U. Asprion, J. Hornung, W. D. Junghans, R. Kostrewa, 1996, Integrated
outcrop analogue studies for Triassic alluvial reservoirs: examples from southern
Germany. Journal of Peteroleum Geology 19, 393406.
Ainsworth, R.B, G. Roger, R.W. Dalrymple, R. Boyd, 1994, Control of estuarine valleyfill deposition by fluctuations of relative sea-level, Cretaceous Bearpaw-Horseshoe
Canyon Transition, Drumheller, Alberta, Canada in B.A., Zaitlin, Brian A. eds. Source:
Special Publication - SEPM (Society for Sedimentary Geology), 51, 159-174, Incisedvalley systems; origin and sedimentary sequences, 1994
Aitken, J. F., and S. S. Flint, 1996, Variable expression of interfluvial sequence
boundaries in the Breathitt Group (Pennsylvanian), eastern Kentucky, USA. In: High
Resolution Sequence Stratigraphy: Innovations and Applications (Eds J.A. Howell and
J.F. Aitken), Special Publication Geological Society, London, 104, 193206.
Allen, G. P., 1991, Sedimentary processes and facies in the Gironde estuary: a recent
model for macrotidal estuarine systems. In: Clastic Tidal Sedimentology (Eds D.G.
Smith, G.E. Reinson, B.A. Zaitlin and R.A. Rahmani), Canadian Society Petroleum
Geologists Memoir, 16, p. 2940.
Allen, J. R. L, 1983, Studies in fluviatile sedimentation: bars, bar-complexes and
sandstone sheets (low-sinuosity braided streams) in the Brownstones (L. Devonian),
Welsh Borders.Sedimentary Geology, 33: p. 237-293.

179

Alling, H. L., 1936, Petrology of the Niagara Gorge sediments; Proceedings of the
Rochester Acadmy of Science, vol. 7, p.189-207
Annan, A. P., W. M. Waller, D. W. Strangway, J. R. Rossiter, J. D. Redman, and R. D.
Watts, 1975, The electromagnetic response of a low-loss, 2-layer, dielectric earth for
horizontal electric dipole excitation. Geophysics 40, p. 285 298.
Annan, A. P., and J. L. Davis, 1976, Impulse radar sounding in permafrost, Radio Sci. 11,
p. 383394.
Annan, A. P., 1999, Practical Processing of GPR Data. Sensors and Software, Ontario.
Annan, A. P., 2002, Ground Penetrating Radar Workshop Notes. Sensors and Software,
Incorporated, Mississauga, Ontario.
Annan, A. P., 2004, Ground Penetrating Radar Principles, Procedures and Applications
Practical Processing of GPR Data. Sensors and Software, Ontario.
Annan, A. P., 2009, Electromagnetic principles of ground penetrating radar in Jol, H.M.,
(Ed) Ground Penetrating Radar Theory and Applications. 524 p
Annan, A. P., and S. W. Cosway, 1992, Ground Penetrating Radar Survey Design:
Proceedings of the Symposium on the Application of Geophysics to Engineering and
Environmental Problems, SAGEEP92, April 26-29, 1992,Oakbrook, IL, p. 329-351.
Annan, A. P., W. M. Waller, D. W. Strangway, J. R. Rossiter, J. D. Redman, and R. D.
Watts, 1975, The electromagnetic response of a low-loss, 2-layer, dielectric earth for
horizontal electric dipole excitation. Geophysics 40, p. 285 298.
Arcone, S. A., D. E. Lawson, A. J. Delaney, J. C. Strasser, and J. D Strasser, 1998,
Ground-penetrating radar reflection profiling of groundwater and bedrock in an area of
discontinuous permafrost. Geophysics 63, p. 15731584.

180

Asprion, U., T. Aigner, 1997, Aquifer architecture analysis using ground-penetrating


radar: Triassic and Quaternary examples (S. Germany). Environmental Geology. 31, p.
66 75.
Asprion, U., T. Aigner, 1999, Towards realistic aquifer models: three-dimensional
georadar surveys of Quaternary gravel deltas (Singen Basin, SW Germany). Sedimentary.
Geology. 129, p. 281 297.
Bailey, S. D., A. G. Wintle, G. A. T. Duller, and C. S. Bristow, 2001, Sand deposition
during the last millennium at Aberffraw, Anglesey, North Wales as determined by OSL
dating of quartz. Quaternary Science Reviews, 20, p. 701-704.
Baker, P. L., 1991, Response of ground-penetrating radar to bound- ing surfaces and
lithofacies variations in sand barrier sequences. A. Neal / Earth-Science Reviews 66
(2004) 261330 321, Exploration. Geophysics 22, p. 1922.
Baker, G. S., T. E. Jordan, and J. Pardy, 2007, An introduction to ground-penetrating
radar (GPR) in Baker, G.S and Jol, H.M (Eds.) Stratigraphic analyses using GPR
Geological Society of America, Special Paper vol. 432, p. 1-18, 2007
Baker, G. S., D. W. Steeples, C. Schmeissner, M. Pavlovic, and R. Plumb, 2001, Nearsurface imaging using coincident seismic and GPR data. Geophysical Research. Letter
28, p. 627630.
Batten K. L, 2000, Architecture of alluvial paleochannels in the Dunvegan Formation,
Northeastern British Columbia, Canada. Unpublished B.Sc thesis, University of Western
Ontario. London Ontario, Canada.
Bazard, D. R., and R. F. Butler, 1991, Paleomagnetism of the Chinle and Kayenta
formations, New Mexico and Arizona. Journal of Geophysical Research, vol. 96, no. B6,
p. 9847-9871
Beard, D. C., and P. K. Weyl, 1973, Influence of texture on porosity and permeability of
unconsolidated sand. American Association of Petroleum Geologists Bulletin, vol. 57,
NO 2, p. 349-369.
181

Beaumont, C., 1981, Foreland basins, Oceanography Dept., Dalhousie Univ., Halifax,
NS, Canada) Source: Geophysical Journal of the Royal Astronomical Society, vol. 65, n
2, p. 291-329
Beaumont, C., G. Quinlan, and J. Hamilton, 1988, Orogeny and stratigraphy: numerical
models of the Paleozoic in the eastern interior of North America. Tectonics, 7: p. 389416.
Beres, M., F. P. Haeni, 1991, Application of ground-penetrating radar methods in
hydrogeologic studies. Groundwater 29, p. 375 386.
Beres, M., A. G. Green, P. Huggenberger, H. Horstmeyer, 1995, Mapping the
architecture of glaciofluvial sediments with three dimensional georadar. Geology 23, p.
1087 1090.
Beres, M., P. Huggenberger, A. G. Green, and H. Horstmeyer, 1999, Using two and three
dimensional georadar methods to characterize glaciofluvial architecture. Sedimentary
Geology, 129, p. 1-24.
Best, J. L., P. J. Ashworth, C. S Bristow, and J. E. Roden, 2003, Three-dimensional
sedimentary architecture of a large, mid-channel sand braid bar, Jamuna River,
Bangladesh. Journal of Sedimentary Research.
Best J., J. Woodward, Ashworth, S. Sambrook, C. Simpson, 2006, Bar top hollows: a
new element in the architecture of sandybraided river. Sedimentary Geology. 190, p. 241255.
Bhattacharya, J. P., and R. G. Walker, 1991b, River and wave-dominated depositional
systems of the Upper Cretaceous Dunvegan Formation, northwestern Alberta. Bulletin of
Canadian Petroleum Geology, 39, p. 165-191.
Bhattacharya, 1994, Chapter 22. Cretaceous Dunvegan Formation of the Western Canada
Sedimentary Basin. In: Geological Atlas of the Western Canada Sedimentary Basin. G.D.
Mossop and I. Shetson (comps.). Canadian Society of Petroleum Geologists and Alberta
Research Council, p. 365-373.
182

Binley, A., P. Winship, R. Middleton, M. Pokar, and J. West, 2001, High-resolution


characterization of vadose zone dynamics using cross-borehole radar. Water Resources
Research 37, p. 2639 2652.
Blakey R., 20008, Stratigraphic columns of the Southern Western Interior, NAU Geology
Blakey, R. C., and R. Gubitosa, 1983, Late Triassic paleogeography and depositional
history of the Chinle Formation, southern Utah and northern Arizona, in Reynolds, M.W.,
and Dolly, E.D., eds., Mesozoic Paleogeography of the West-Central United States:
Denver, Rocky Mountain Section SEPM, p. 5776.
Blakey, R. C. and R. Gubitosa, 1984, Controls on sandstone body geometry and
architecture in the Chinle Formation (Upper Triassic), Colorado Plateau. Sedimentary
Geology, 38, p. 5186.
Blakey, R. C., F. Peterson, and G. Kocurek, 1988, Synthesis of late Paleozoic and
Mesozoic eolian deposits of the Western Interior of the United States. Sedimentary
Geology, vol. 56, no. 1-4, p. 3-125
Blaird, D., 1980, A prosaurod dinosaur trackway from the Navajo Sandstone (Lower
Jurrasic) of Arizona, in Jacobs L.L (Ed), Aspects of vertebrate history: Flagstaff,
Museum of Northern Arizona Press. p. 219 -230
Blodgett, R. H., 1988, Calcareous paleosols in the Triassic Dolores Formation,
southwestern Colorado: Geological Society of America Special Paper 216, p. 103 -121
Bolton, T. E., 1949, Stratigraphy and paleontology of the Silurian section at DeCew
Falls, Ontario; Unpublished M.Sc thesis, University of Toronto, Toronto, Ontario.
Bolton, T. E., 1953, Silurian formations of the Niagara Escarpment in Ontario;
Geological Survey of Canada, Paper 53-23.
Bolton, T. E., 1957, Silurian stratigraphy and paleontology of the Niagara Escarpment in
Ontario; Geological Survey of Canada, Memoir 289.

183

Botha, G. A., C. S. Bristow, N. Porat, G. Duller, S. J. Armitage, H. M. Roberts, B. M.


Clarke, M. W. Kota, and P. Schoeman, 2003, Evidence for dune reactivation from GPR
profiles on the Maputaland coastal plain, South Africa. In: Bristow, C.S., Jol, H.M.
(Eds.), Ground Penetrating Radar in Sediments. Geological Society of London Special
Publication. 211, p. 29 46.
Brenchley, P. J., and G. Newall, 1984, Late Ordovician environmental changes and their
effects on faunas. In Aspects of the Ordovician System. Edited by D.L. Bruton.
Paleontological Contributions from the University of Oslo, No. 295, Universitetsforlaget,
Oslo, Norway, p. 65-79.
Brett, C. E., W. M. Goodman, and S. T. Loduca, 1990, Sequences, cycles, and basin
dynamics in the Silurian of the Appalachian foreland basin: Sedimentary Geology, v. 69,
p. 191244.
Bridge, J. S., 2006, Fluvial Facies Models: Recent Developments, in., Posamentier H.W
and Walker R.G., eds., Facies Models Revisited, p. 85170.
Bridge, J. S., J. Alexander, R. E. L.Collier, R. L. Gawthorpe, and J. Jarvis, 1995, Ground
penetrating radar and coring used to document the large-scale structure of point-bar
deposits in 3-D: Sedimentology, v. 42, p. 839852.
Bridge, J. S., R. E. L.Collier, and J. Alexander, 1998, Large-scale structure of Calamus
River deposits (Nebraska, USA) revealed using ground penetrating radar: Sedimentology,
v. 45, p. 977986.
Bridge, J. S., and I. A. Lunt, 2006, Depositional models of braided rivers, in Sambrook
Smith, G.H., Best, J.L., Bristow, C.S., and Petts, G., eds., Braided Rivers; Process,
Deposits, Ecology, and Management: International Association of Sedimentologists,
Special Publication 36, p. 1150.
Bridge, J. S., and S. D. Mackey, 1993b, A theoretical study of fluvial sandstone body
dimensions, in S. S. Flint and I. D. Bryant, eds.,Geological modelling of hydrocarbon

184

reservoirs: International Association of Sedimentologists, Special Publication 15,


Utrecht, Netherlands, p. 213236.
Bristow, C. S., 1995, Facies analysis in the Lower Greensand using ground-penetrating
radar. Journal of the Geological Society, 152, p. 591 598.
Bristow, C. S., H. M. Jol, (Eds.), 2003, Ground Penetrating Radar in Sediments.
Geological Society of London Special Publication, vol. 211.
Bristow, C. S., J. Pugh, T. Goodall, 1996, Internal structure of aeolian dunes in Abu
Dhabi determined using ground-penetrating radar. Sedimentology 43, p. 995 1003.
Bristow, C. S., R. L. Skelly, F. G. Etheridge, 1999, Crevasse splays from the rapidly
aggrading, sand-bed, braided Niobrara River, Nebraska: effect of base-level rise.
Sedimentology 46, p. 1029 1047.
Bristow, C. S., S. D. Bailey, N. Lancaster, 2000a, The sedimentary structure of linear
sand dunes. Nature 406, p. 5659.
Bristow, C. S., J. L. Best, P. J. Ashworth, 2000b, The use of GPR in developing a facies
model for a large sandy braided river, Brahmaputra River, Bangladesh. In: Noon, D.A.,
Stickley, G.F., Longstaff, D. (Eds.), Proceedings of the Eighth International Conference
on Ground Penetrating Radar. SPIE, Billingham, vol. 4084, p. 95 100.
Bristow, C. S., P. N. Chroston, S. D. Bailey, 2000c, The structure and development of
foredunes on a locally prograding coast: insights from ground penetrating radar surveys,
Norfolk, UK. Sedimentology 47, p. 923 944.
Brown, L. F., and W. L. Fisher, 1980, Seismic stratigraphic interpretation and petroleum
exploration. American Association of Petroleum Geologists, Continuing Education
Course Note Series, 16
Brown, A. R., 2003, Interpretation of Three Dimensional Seismic Data, 6th edition,
AAPG Mem., 42, 541pp.,Tulsa,OK

185

Busby-Spera, C., 1988, Speculative tectonic model for the early Mesozoic arc of the
southwest Cordilleran United States: Geology, V. 16, p. 1121-1125
Calow, R., 1983, A petrographic and diagenetic study of the Whirlpool Sandstone from
outcrops in Hamilton and Niagara Gorge areas; Unpublished B.Sc thesis, Mc Master
University, Hamilton, Ontario.
Carcione, J. M., and F. Cavallini, 1995, On the acoustic-electromagnetic analogy. Wave
Motion 21, p. 149 162.
Cardenas, M. B., and V. A. Zlotnik, 2003, Three-dimensional model of modern channel
bend deposits. Water Resources Research 39, 1141.
Cartwright, J. A. and M. Huuse, 2005, 3D seismic technology: the geological Hubble.
Basin Research, 17, p. 120.
Castle, J. W., 1998, Regional Sedimentology and Stratal Surfaces of a Lower Silurian
Clastic Wedge in the Appalachian Foreland Basin, Journal of Sedimentary Research, vol.
68.
Castle, J. W., and A. P. Byrnes, 1998, Petrophysics of low-permeability Medina
Sandstone, northwestern Pennsylvania, Appalachian basin: The Log Analyst, v. 39, no. 4,
p. 3646.
Chan, M., D. Netoff, R. Blakey, G. Kocurek, and W. Alvarez, 2007, Clastic-injection
pipes and syndepositional deformation structures in Jurassic eolian deposits: Examples
from the Colorado Plateau, in A. Hurst and J. Cartwright, eds., Sand injectites:
Implications for hydrocarbon exploration and production: AAPG Memoir 87, p. 233
244.
Chaumillon, E., J. N. Proust, D. Meunier, and N. Weber, 2008, Incised-valley
morphologies and sedimentary fills within the inner shelf of the Bay of Biscaye (France):
a synthesis. J. Mar. Syst., 72, p. 383-396.

186

Cheel R. J., and G. V. Middleton, 1993, Directional scours on a transgressive surface:


Examples from the Silurian Whirlpool Sandstone of Southern Ontario, Canada . Journal
of Sedimentarv Research Vol 63, No. 3, p. 392-397
Christian, S., and N. Klaus-Peter, 1994, Eccentricity-migration: A method to improve the
imaging of pipes in radar reflection data. Proceedings of the 5th International Conference
on Ground Penetrating Radar (GPR94), Canada, p. 723733.
Clark, J. D., and K. T. Pickering, 1996, Architectural elements and growth patterns of
submarine channels; application to hydrocarbon exploration. AAPG Bulletin, vol.80,
no.2, p.194-221.
Clemmensen, L. B., K. Pye, A. Murray, and J. Heinemeier, 2001, Sedimentology,
stratigraphy and landscape evolution of a coastal dune system, Lodbjerg, NW Jutland,
Denmark. Sedimentology 48, p. 3 27
Cleveland, D. M., S. C. Atchley, and L. C. Nordt, 2007, Continental Sequence
Stratigraphy of the Upper Triassic (Norian-Rhaetian) Chinle Strata, Northern New
Mexico, U.S.A.: Allocyclic and Autocyclic Origins of Paleosol-Bearing Alluvial
Successions. Journal of Sedimentary Research Vol. 77, no. 11, p.909-924,
Collinson, J. D., 1978, Vertical sequence and sand body shape in alluvial sequences, in
A. D. Miall, ed., Fluvial sedimentology: Canadian Society of Petroleum Geologists
Memoir 5, p. 577586.
Collinson, J. D., and D. B. Thompson, 1989, Sedimentary Structures, 2nd edition.
Chapman and Hall, London.
Conyers, L. B., D. Goodman, 1997, Ground-Penetrating Radar: An Introduction for
Archaeologists. Altamira Press, London.
Coogan, A. H., 1990, Whirlpool (Early Silurian) marine transgression in eastern Ohio;
Northeastern Geology, v. 12, p. 138-142

187

Cook, J. C. 1975, Radar transparencies of mine and tunnel rocks. Geophysics, 40, p. 865885.
Corbeanu, R. M., K. Soegaard, R. B. Szerbiak, J. B. Thurmond, G. A. McMechan, D.
Wang, S. Snelgrove, C. B. Forster, and A. Menitove, 2001, Detailed internal architecture
of a fluvial sandstone determined from outcrop, cores, and 3-D ground penetrating radar:
example from the middle Cretaceous Ferron Sandstone, eastcentral Utah. AAPG
Bulletin. 85, p. 15831608.
Corbeanu, R. M., G. A. McMechan, R. B. Szerbiak, K. Soegaard, 2002, Prediction of 3-D
fluid permeability and mudstone distributions from ground-penetrating radar (GPR)
attributes: examples from the Cretaceous Ferron Sandstone Member, east central Utah.
Geophysics 67, p. 14951504.
Cotter E., 1978, The evolution of fluvial style, with special reference to the Central
Appalachian Paleozoic. In Miall, A. D. (Ed) Memoir - Canadian Society of Petroleum
Geologists, no.5, p.361-383
Crerar, E. Erin, and R. W. C. Arnott, 2007, Facies distribution and stratigraphic
architecture of the Lower Cretaceous McMurray Formation, Lewis Property, northeastern
Alberta. Bulletin of Canadian Petroleum Geology, vol. 55, no. 2, p.99-124.
Daniels, D. J, 2004, Ground Penetrating Radar, 2nd edition, Radar, Sonar, Navigation
and Avionics Series 15, Institute of Electrical Engineers, London, UK.
Davies, N. S., and M. R. Gibling, 2010, Cambrian to Devonian evolution of alluvial
systems: The sedimentological impact of the earliest land plants. Earth-Science Reviews,
Volume 98, Issues 3-4, p. 171-200
Davis, J. L., and A. P. Annan, 1989, Ground-penetrating radar for high resolution
mapping of soil and rock stratigraphy. Geophysical Prospecting. 3, p. 531551
Dennison, J. M. and J. W. Head, 1975, Sea level variations interpreted from the
Appalachian Basin Silurian and Devonian; American Journal of Science, vol. 275, p.
1089-1120
188

Dennison, J. M., 1976, Appalachian Queenston delta related to eustatic sea-level drop
accompanying Late Ordovician glaciation centred in Africa, in Bassett, M.G., ed., The
Ordovician System: Proceedings of a Paleontological ontological Association
symposium, Birmingham: Cardiff, University of Wales Press and National Museum of
Wales, p. 107120.
Deptuck, M. E., D. J. W. Piper, B. Savoye, and A. Gervais, 2008, Dimensions and
architecture of late Pleistocene submarine lobes off the northern margin of East Corsica:
Sedimentology, vol. 55, p. 869898.
Deutsch, C. V., and T. T. Tran, 2002, FLUVSIM: A program for object-based stochastic
modeling of fluvial depositional systems: Computers and Geosciences, vol. 28, p. 525
535.
Di, Q. Y., and M. Y. Wang, 2004, Migration of ground-penetrating radar data method
with a finite element and dispersion, Geophysics, vol. 69, No. 2, p. 472477.
Dickinson, W. R., 1981, Plate tectonic evolution of the southern Cordillera, in Dickinson,
W.R., and Payne, W.D., eds., Relations of Tectonic to Ore Deposits in the Southern
Cordillera: Arizona Geological Society Digest, v.14, p. 113 135.
Doelling, H. H., 1985, Geology of Arches National Park. Accompanied by Map 74, Utah
Geological and Mineral Survey, Salt Lake City, 15 pp.
Dreyer, T., 1990, Sand body dimensions and infill sequences of stable, humid-climate
delta plain channels, in Buller, A. T., Berg, E., Hjelmeland, O., Kleppe, J., Torsaeter, O.,
and Aasen, J. O., eds., North Sea oil and gas reservoirsII: Graham and Trotman,
London, p. 337-351.
Dreyer, T., 1993a, Geometry and facies of large-scale flow units in fluvial-dominated
fan-delta-front sequences, in Ashton, M., ed., Advances in reservoir geology, Geological
Society, London, Special Publication 69, p. 135-174.

189

Dreyer, T., 1993b, Quantified fluvial architecture in ephemeral stream deposits of the
Esplugafreda Formation (Palaeocene), Tremp-Graus Basin, northern Spain, in Marzo, M.,
and Puigdefbregas, C., eds., Alluvial sedimentation: International Association of
Sedimentologists Special Publication 17, p. 337-362. QE581 .A458 1993
Dreyer, T., L. M. Flt, T. Hy, R. Knarud, R. J. Steel, and J. L. Cuevas, 1993,
Sedimentary architecture of field analogues for reservoir information (SAFARI): a case
study of the fluvial Escanilla Formation, Spanish Pyrenees, in Flint, S. S., and Bryant, I.
D., eds., The geological modelling of hydrocarbon reservoirs and outcrop analogues:
International Association of Sedimentologists Special Publication 15, p. 57-80.
Dubiel, R. F., ed., 1983, Stratigraphic sections of the Shinarump, Monitor Butte, and
Moss Back members of the Upper Triassic Chinle Formation in the northern part of the
White Canyon, Red Canyon, and Blue Notch Canyon area, southeastern Utah, Source:
Open-File Report - U. S. Geological Survey, 32.
Dubiel, R. F., 1987, Sedimentology of the Upper Triassic Chinle Formation, southeast
Utah: Ph.D. thesis, University of Colorado at Boulder, Boulder, Colorado, 132 p.
Dubiel, R. F., 1989, Paleoclimatic cycles and tectonic controls on fluvial lacustrine, and
eolian strata in Upper Triassic Chinle Formation, San Juan Basin, In: AAPG Bulletin, 73,
9, p. 1153-1154.
Dubiel, R. F., J. T. Parrish, J. M. Parrish, and S. C. Good, 1991, The Pangaean
megamonsoon Evidence from the Upper Triassic Chinle Formation, Colorado Plateau:
Palaios, v. 6, p. 347370.
Ekes, C., and P. Friele, 2003, Sedimentary architecture and post-glacial evolution of
Cheekye fan, southwestern British Columbia, Canada. In: Bristow, C.S., Jol, H.M. (Eds.),
Ground Penetrating Radar in Sediments. Geological Society of London Special
Publication.. 211, p. 87 98.

190

Ellis, D., 1993. The Rough Gas Field; distribution of Permian aeolian and non-aeolian
reservoir facies and their impact on field development. Geological Society Special
Publications, 73, 265-277, Characterization of fluvial and aeolian reservoirs.
Emmett, W. R., K. W. Beaver, and J. A. McCaleb, 1971, Little Buffalo Basin Tensleep
heterogeneity its influence on infilling drilling and secondary recovery. Journal of
Petroleum Technology, p. 161-168.
Energy Resource Conservation Board, Alberta, 2009. Alberta Table of Formations
Ethridge, F. G. and S. A Schumm, 1978, Reconstructing paleochannel morphologic and
flow characteristics; methodology, limitations and assessment. In: Miall, A. D., ed.
Memoir - Canadian Society of Petroleum Geologists, 5, 703-721, Fluvial sedimentology.
Ettensohn, F. R., 1991, Flexural interpretation of relationships between Ordovician
tectonism and stratigraphic sequences, Central and Southern Appalachians, U.S.A. Paper
- Geological Survey of Canada, Report: 90-09, p.213-224
Fairchild, H. L., 1901, Beach structure in the Medina Sandstone, American Geologist,
vol. 28, p. 9-14
Fielding, C. R., and R. C. Crane, , 1987, An application of statistical modelling to the
predictionof hydrocarbon recovery factors in fluvial reservoir sequences, in Ethridge, F.
G., Flores, R. M., and Harvey, M. D., eds., Recent developments in fluvial
sedimentology: Society of Economic Paleontologists and Mineralogists Special
Publication 39, p. 321-327.
Fisher, D. W., 1954, Stratigraphy of Medina Group, New York and Ontario; American
Association of Petroleum Geologist, Bulletin, vol. 38, p. 1979-1996
Fisher, E., G. A. McMechan, A. P. Annan, and S. W. Cosway, 1992a, Examples of
reverse-time migration of single-channel, ground penetrating radar profiles. Geophysics
57, p. 577 586.

191

Fisher, S. C., R. R. Stewart, and H. M. Jol, 1994, Processing ground penetrating radar
data. Proceedings of the 5th International Conference on Ground Penetrating Radar
(GPR94), Canada, p. 661675.
Fisher, S. C., R. R. Stewart, and H. M. Jol, 1996, Ground penetrating radar (GPR) data
enhancement using seismic techniques. J. Environ. Eng. Geophys. 1, p. 8996.
Fowler, J. C., and W. L. Still, 1977, Surface and subsurface profiling using groundprobing radar. Geophysics 42, p. 15021503.
Freeman, W. E., and G. S. Visher, 1975, Stratigraphic analysis of the Navajo Sandstone:
Jour. Sed. Petrology, vol. 45, p. 651--668.
Fryberger, S. G., 1990a, Great Sand Dunes depositional systeman overview, in
Fryberger, S.G., Krystinik, L.F., and Schenk, C.J., eds., Modern and Ancient Aeolian
Deposits: Petroleum Exploration and Production: SEPM, Denver, p. 1-11-9. Fryberger,
1990
Gawthorpe, R. L., R. E. L. Collier, J. Alexander, M. Leeder, and J. S. Bridge, 1993,
Ground penetrating radar: application to sandbody geometry and heterogeneity studies.
In: North, C.P., Prosser, D.J. (Eds.), Characterization of Fluvial and Aeolian Reservoirs.
Geological Society of London Special Publication, vol. 73, p. 421432.
Geehan, G., 1993, The use of outcrop data and heterogeneity modelling in development
planning, in Eschard, R. and Doligez, B., eds., Subsurface reservoir characterization from
outcrop observations: Institut Franais du Petrole. ditions Technip, Paris, p. 53-64.
Geitz, O. 1952. The Whirlpool Sandstone; Unpublished M.Sc thesis, McMaster
University, Hamilton, Ontario.
Gerlitz, K., M. D. Knoll, G. M. Cross, R. D. Luzitano, and R. Knight, 1993, Processing
ground penetrating radar data to improve resolution of near-surface targets, Proceeding of
the Symposium on the Application of Geophysics to Engineering and Environmental
Problems, San Diego, California

192

Gibling, M. R., 2006, Width and thickness of fluvial channel bodies and valley fills in the
geological record: A literature compilation and classification Journal of Sedimentary
Research, vol. 76, no.5-6, p.731-770.
Gilbert, G. K., 1899, Ripple-marks and cross-bedding. Geological Society of America,
Bulletin, vol. 10, p. 135-140
Gingras, M. K., J. A. MacEachern, and S. G. Pemberton, 1998, A comparative analysis of
the ichnology of wave- and river-dominated allomembers of the Upper Cretaceous
Dunvegan Formation. Bulletin of Canadian Petroleum Geology, 46, p. 51-73.
Glennie, K. W., 1972, Permian Rotliegendes of Northwest Europe Interpreted in Light of
Modern Desert Sedimentation Studies. The American Association of Petroleum
Geologists Bulletin, 56, 6, p. 1048-1071
Glennie, K. W., 1983, Lower Permian Rotliegend desert sedimentation in the North Sea
area. p. 521-541
Glennie, K. W., 1990, Lower PermianRotliegend, in Glennie, K.W., ed., Introduction
to the Petroleum Geology of the North Sea: Oxford, U.K., Blackwell Science, p. 120
152.
Goodman, D., Y. Nishimura, and J. D. Rogers, 1995, GPR Time Slices in Archaeological
Prospection, Archaeological Prospection, 2: p. 85-89
Grabau, A. W., 1913, Early Paleozoic delta deposits of North America; Geological
Society of America, Bulletin, vol. 24, p. 399-528.
Grammer, G. M., P. M. Harris, , 2004, Integration of outcrop and modern analogs in
reservoir modeling; overview with examples from the Bahamas, In: Eberli,G. P. AAPG
Memoir, 80, 1-22, Integration of outcrop and modern analogs in reservoir modeling.
Gran, K., and C. Paola, 2001, Riparian vegetation controls on braided stream dynamics.
Water Resour. Res., 3702), p. 3275-3283.

193

Greaves, R. J., D. P. Lesmes, J. M. Lee, and M. N. Toksoz, 1996, Velocity variations


and water content estimated from multi-offset, ground-penetrating radar. Geophysics 61,
p. 683 695.
Hammon III, W.S., X. Zeng, R. M. Corbeanu, and G. A. McMechan, 2002, Estimation of
the spatial distribution of fluid permeability from surface and tomographic GPR data and
core, with a 2-D example from the Ferron Sandstone, Utah. Geophysics 67, p.15051515.
Harari, Z., 1996, Ground-penetrating radar (GPR) for imaging stratigraphic features and
groundwater in sand dunes. J. Appl.Geophys. 36, p. 4352.
Hatton, L., M. H. Worthington, and J. Makin, 1986, Seismic Data Processing: Theory and
Practice. Blackwell, Oxford.
Havholm, K. G., N. D. Bergstrom, H. M. Jol, and G. L. Running, IV. 2003, GPR survey
of a Holocene aeolian/fluvial/lacustrine succession, Lauder Sandhills, Manitoba, Canada.
In: Bristow, C. S. and Jol, H. M. (Eds) Ground Penetrating Radar in Sediments.
Geological Society, London, Special Publications, 211, p. 47-54.
Hay, W.W., D. L. Eicher, and R. Diner, 1993, Physical oceanography and water masses
in the Cretaceous Western Interior Seaway. Geological Association of Canada Special
Paper, vol. 39, p. 297-318
Hayakawa, H. and A. Kawanaka, 1998, Radar imaging of underground pipes by
automated estimation of velocity distribution versus depth. Journal of Applied
Geophysics, vol. 40, p. 3748.
Heincke, B., A. G. Green, J. van der Kruk, and H. Willenberg, 2006, Semblance-based
topographic migration (SBTM): A method for identifying fracture zones in 3D georadar
data. Near Surface Geophysics, vol. 4, No. 2, p. 7988.
Heinz, J., 2001, Sedimentary geology of glacial and periglacial gravel bodies (SWGermany): dynamic stratigraphy and aquifer sedimentology. Unpublished. PhD Thesis,
Univ Tubingen.

194

Heinz, J., and T. Aigner, 2003, Three-dimensional GPR analysis of various Quaternary
gravel-bed braided river deposits (southwestern Germany). In: Bristow, C.S., Jol, H.M.
(Eds.), Ground Penetrating Radar in Sediments. Geological Society of London Special
Publication. 211, p. 99110.
Heller P. L., K. Dueker, and M. E. McMillan, 2003, Post-Paleozoic alluvial gravel
transport as evidence of continental tilting in the U.S. Cordillera. GSA Bulletin;
September 2003; vol. 115; no. 9; p. 11221132
Herries, R. D., 1993, Contrasting styles of fluvialaeolian interaction at a downwind erg
margin: Jurassic KayentaNavajo transition, northeastern Arizona, USA, in North, C.P.,
and Prosser, J.D., eds., Characterization of Fluvial and Aeolian Reservoirs: Geological
Society of London, Special Publication 73, p. 199218.
Hickin, A. S., P. T. Bobrowsky, R. C. Paulen, and M. Best, 2007, Imaging fluvial
architecture within a paleovalley fill using ground-penetrating radar, Maple Creek,
Guyana Special Paper - Geological Society of America, vol. 432, p. 133-153
Holden, J., T. P. Burt, and M. Vilas, 2002, Application of ground-penetrating radar to the
identification of subsurface piping in blanket peat. Earth Surf. Processes Landf. 27, p.
235 249.
Holstein, A., 1936, Heavy minerals in Silurian rocks of the Niagara Escarpment, Ontario;
Unpublished MSc. Thesis, University of Toronto, Toronto, Ontario.
Hornung, J., and T. Aigner, 2002, Reservoir architecture in a terminal alluvial plain: an
outcrop analogy study (Upper Triassic, southern Germany): Part 1. Sedimentology and
petrophysics. J. Petrol. Geol. 25, p. 3 30.
Hurlbert, J. C., 1995, Stratigraphic analysis of the Shinarump member of the Triassic
Chinle Formation, northern Arizona and southern Utah. Unpublished thesis, Northern
Arizona University

195

Hugenholtz, C. H., B. J. Moorman, and S. A. Wolfe, 2007, Ground-penetrating radar


(GPR) imaging of the internal structure of an active parabolic sand dune. Special Paper Geological Society of America, vol. 432, p. 35-45
Huggenberger, P., 1993, Radar facies: recognition of facies patterns and heterogeneities
within Pleistocene Rhine gravels, NE Switzerland. In: Best, J.L., Bristow, C.S. (Eds.),
Braided Rivers. Geological Society of London Special Publication. 75, p. 163 176.
Huggenberger, P., E. Meier, and A. Pugin, 1994, Ground-probing radar as a tool for
heterogeneity estimation in gravel deposits: advances in data-processing and facies
analysis. Journal Applied Geophysics. 31, p. 171184.
Hummel, G., and G. Kocurek, 1984, Interdune areas of the back-island dune field, North
Padre Island, Texas. Sedimentary Geology, vol. 39, no. 1-2, p. 1-26
Ilya, V., 2006, Coastal environmental changes revealed in geophysical images of
Nantucket Island, Massachusetts, U.S.A. Environmental and Engineering Geoscience,
vol.12, no.3, p. 227-234, Jan 2006
Johnson, H., 1934, The stratigraphy and paleontology of the Cataract Formation in
Ontario; Unpublished Ph.D. thesis, University of Toronto, Toronto, Ontario.
Jol, H. M., 1995, Ground penetrating radar antenna frequencies and transmitter powers
compared for penetration depth, resolution and reflection continuity. Geophysical
Prospecting 43, p. 693-709.
Jol, H. M., and C. S. Bristow, 2003, GPR in sediments: advice on data collection, basic
processing and interpretation, a good practice guide. In: Bristow, C.S., Jol, H.M. (Eds.),
Ground Penetrating Radar in Sediments. Geological Society of London Special
Publication. 211, p. 9 27.
Jol, H. M., and D. G. Smith, 1991, Ground penetrating radar of northern lacustrine deltas.
Canadian Journal of Earth Sciences. 28, p. 1939 1947.

196

Jol H. M, C. S. Bristow, D. G. Smith, M. B. Junck, and P. Putnam, 2003, Stratigraphic


imaging of the Navajo Sandstone using ground-penetrating radar in The Leading Edge p.
882-887
Jol, H. M., D. G. Smith, and R. A. Meyers, 1996a, Digital Ground Penetrating Radar
(GPR): a new geophysical tool for coastal barrier.
Jol, H. M., S. Vanderburgh, and K. G. Havholm, 1998, GPR studies of coastal aeolian
(foredune and crescentic) environments: examples from Oregon and North Carolina,
U.S.A. Proceedings of the Seventh International Conference on Ground Penetrating
Radar (GPR '98), Lawrence, Kansas, USA, May 27-30, 2, p. 681-686.
Jordan, D. W., W. A. Pryor, 1987, Vertical and lateral reservoir inhomogeneities of a bigriver bar deposit, Mississippi River, Missouri. AAPG Bulletin, 71, 9, 1107, Association
round table; AAPG Eastern Section meeting; abstracts.
Kauffman, E. G., and W. G. E. Caldwell, 1993, The Western Interior Basin in space and
time. Geological Association of Canada Special Paper, vol. 39, p 1-30.
Kaulbach, C., 1995, A Petrographic Comparison of Lateral and Axial Depositional
Systems in a Foreland Basin The Dunvegan Formation Alberta. Canada, Bachelor of
Science Thesis, University of Western Ontario, Earth Sciences Department, 85 P.
Kearey, P., and M. Brooks, 1991, An Introduction to Geophysical Exploration.
Blackwell, Oxford.
Kocurek, G., 1981, Significance of interdune deposits and bounding surfaces in aeolian
dune sands: Sedimentology, vol. 28, p. 753780.
Kocurek, G., and R. H. Dott Jr., 1981, Distinctions and uses of stratification types in the
interpretation of eolian sand. Journal of Sedimentary Petrology, vol. 51, n 2, p. 579-596
Kocurek, G. and G. Fielder, 1982, Adhesion structures. Journal of Sedimentary
Petrol.ogy, 52: p. 1229-1241.

197

Kocurek, G., and K. G. Havholm, 1993, Eolian sequence stratigraphy-a conceptual


framework, in Weimer, P., and Posamentier, H.W., eds., Siciliclastic Sequence
Stratigraphy: American Association of Petroleum Geologists, Memoir 58, p. 393409.
Kocurek, G., and R. E. Hunter, 1986, Origin of polygonal fractures in sand, uppermost
Navajo and Page Sandstones, Page, Arizona: Journal of Sedimentary Petrology, vol. 56,
p. 895904.
Koepke, W., 1960, The Whirlpool Formation as building stone, Georgetown area,
Ontario; Unpublished B.Sc. thesis, Univesity of Western Ontario, London Ontario
Kraus, M. J., L. T. Middleton, 1987, Dissected paleotopography and base-level changes
in a Triassic fluvial sequence. Geology (Boulder), 15, 1, p. 18-21
Lahann, R. W., J. A. Ferrier, and S. Corrigan, 1993, Reservoir heterogeneity in the
Vanguard Field, in Ashton, M. (ed.), 1993, Advances in Reservoir Geology, Geological
Society Special Publication No. 69, p. 33-56.
LaFleche, P. T., J. P. Todoeschuck, O. G. Jensen, and A. S. Judge, 1991, Analysis of
ground-probing radar data: predictive deconvolution. Canadian Geotechnical Journal. 28,
p. 134 139.
Larue, D. K., and J. Hovadik, 2006, Connectivity of channelized reservoirs: A modeling
approach: Petroleum Geoscience, vol. 12, p. 291308.
Laughrey, C. D., 1984, Petrology and reservoir characteristics of the Lower Silurian
Medina Group sandstones, Athens and Geneva fields, Crawfield County, Penslyvania;
Pensylvania Geological Survey, Mineral Resources Report 85
Leatherman, S. P., 1987, Coastal geomorphological applications of ground-penetrating
radar. Journal of Coastal Research. 3, p. 397 399.
Leckebusch, J., 2003, Ground-penetrating radar: a modern three-dimensional prospection
method Archaeological Prospection, vol. 10, no. 4, p. 213 240

198

Leclerc, R. F., and E. J. Hickin, 1997, The internal structure of scrolled floodplain
deposits based on ground-penetrating radar, North Thompson River, British Columbia.
Geomorphology, 21, p. 17 38.
Lee, K., R. Szerbiak, G. A. McMechan, N. A. Hwang, 2009, 3-D ground-penetrating
radar and wavelet transform analysis of the morphology of shoreface deposits in the
Upper Cretaceous Ferron Sandstone Member, Utah. AAPG Bulletin, vol. 93, no. 2,
p.181-201
Lehmann, F. and A. G. Green, 2000, Topographic migration of georadar data:
Implications for acquisition and processing. Geophysics, vol. 65, No. 3, p. 836848.
Lericolais, G., S. Bern, and H. Fnis, 2001, Seaward pinching out and internal
stratigraphy of the Gironde incised valley on the shelf. Marine Geology, 175, p. 183197.
Leuschen, C. and R. Plumb, 2000, A matched-filter approach to wave migration. Journal
of Applied Geophysics, vol. 43, p. 271280.
Lindquist, S. J, 1988, Practical characterization of eolian reservoirs for development;
Nugget Sandstone, Utah-Wyoming thrust belt. Sedimentary Geology, vol.56, no.1-4, p.
315-339
Lockley, M. G., and K. Conrad, 1989, The paleoenvironmental context, preservation and
paleoecological significance of dinosaur tracksites in the western USA, in Gillette, D.D.,
and Lockley M.G., (Eds) Dinosaur tracks and traces: Cambridge, England, Cambridge
University Press, p. 121 134.
Lockwood, W. N., 1942, Sedimentation problem of the White Medina or Whirlpool
Sandstone; Buffalo Society of Natural Sciences, Bulletin, vol. 17. p. 64
Long, D. G. F., 1978, Proterozoic stream deposits; some problems of recognition and
interpretation of ancient sandy fluvial systems. In Miall, A. D (Ed) Memoir - Canadian
Society of Petroleum Geologists, no.5, p. 313-341

199

Lorenz, J. C., D. M. Heinze, J. A. Clark, and C. A. Searls, 1985, Determination of widths


of meander-belt sandstone reservoirs from vertical downhole data, Mesaverde Group,
Piceance Creek basin, Colorado: AAPG Bulletin, vol. 69, p. 710 -721.
Lorenz, J. C., N. R. Warpinski, and P. T. Brannagan, 1991, Subsurface characterization
of Mesaverde reservoirs in Colorado: geophysical and reservoir-engineering checks on
predictive sedimentology, in A. D. Miall and N. Tyler, eds., The three-dimensional facies
architecture of terrigenous clastic sediments and its implications for hydrocarbon
discovery and recovery: SEPM Concepts in Sedimentology and Paleontology 3, p. 5779.
Lucas, S. G., 1993, The Chinle Group; revised stratigraphy and biochronology of Upper
Triassic nonmarine strata in the Western United States. Museum of Northern Arizona
Bulletin, vol. 59, p.27-50
Lumsdon, M. P., 2000, Alluvial architecture and palaeoenvironments in a proximal
foreland basin setting: Upper Cretaceous (Cenomanian) Dunvegan Formation, BC.
Unpubl. MSc Thesis, University of Western Ontario, London, Canada, 299 pp.
Lumsdon-West, M. P. and A. G. Plint, 2005, Changing alluvial style in response to
changing accommodation rate in a proximal foreland basin setting: Upper Cretaceous
Dunvegan Formation, NE British Columbia, Canada. In: Fluvial Sedimentology VII.
M.D. Blum, S.B. Marriott and S. Leclair (eds.). International Association of
Sedimentologists, Special Publication No. 35, p. 493515.
Lunt, I. A., and J. S. Bridge, 2004, Evolution and deposits of a gravelly braid bar and a
channel fill, Sagavanirktok river, Alaska: Sedimentology, vol.51, p. 415432.
Lupe, R., T. S. Ahlbrandt, E. D. McKee, 1979. Sediments of the ancient eolian
environment; reservoir inhomogeneity. U. S. Geological Survey Professional Paper, 241251, A study of global sand seas.
Lutz, P., S. Garambois, and H. Perrouo, 2003, Influence of antenna configurations for
GPR survey: information from polarization and amplitude versus offset measurements.

200

In: Bristow, C. S. and Jol, H. M. (eds) Ground Penetrating Radar in Sediments.


Geological Society, London, Special Publications, 211, p. 299-313.
Maijala, P., 1992, Application of some seismic data processing methods to ground
penetrating radar data. In: Hanninen, P., Autio, S. (Eds.), Fourth International
Conference on Ground Penetrating Radar. Geol. Surv. Finland Special Papers. 16, p.
103 110.
Martini, I. P., 1966, The sedimentology of the Medina Formation along the Niagara
Escarpment (Ontario and New York State); Unpublished Ph.D thesis, McMaster
University, Hamilton, Ontario.
Martini, I. P., and J. K. P. Kwong, 1985, Depositional Characteristics and resources
potential of the Whirlpool Sandstone, Lower Silurian, Ontario; Ontario Geological
Survey, Open File Report 5363.
Martini, I. P., and C. J. Salas, 1983, Depositional characteristics of the Whirlpool
Sandstone, Lowe Silurian, Ontario; Ontario Geological Survey, Open File Report 5549
Marzolf, J. E., 1988, Controls on late Paleozoic and early Mesozoic eolian deposition of
the Western United States. Sedimentary Geology, vol.56, no.1-4, p. 167-191
McCann, D. M., P. D. Jackson, and P. J. Fenning, 1988, Comparison of the seismic and
ground probing radar methods in geological surveying. Proceedings of Institute of
Electrical Engineers 135, p. 380 390.
McCarthy, P. J., 2002, Micromorphology and development of interfluve paleosols: a case
study from the Cenomanian Dunvegan Formation, NE British Columbia, Canada.
Bulletin of Canadian Petroleum Geology, 50, p. 158177.
McCarthy, P. J., and A. G. Plint, 1999, Floodplain palaeosols of the Cenomanian
Dunvegan Formation, Alberta and British Columbia, Canada: micromorphology,
pedogenic processes and palaeoenvironmental implications. In: Floodplains:
Interdisciplinary Approaches (Eds J. Alexander and S.B. Marriott), Special Publication.
Geological Society of London, 163, p. 289310.
201

McCarthy, P. J., U. F. Faccini, and A. G. Plint, 1999, Evolution of an ancient coastal


plain: palaeosols, interfluves and alluvial architecture in a sequence stratigraphic
framework, Cenomanian Dunvegan Formation, NE British Columbia, Canada.
Sedimentology, 46, p. 861891.
McDonough, K. J., and T. A. Cross, 1991, Late Cretaceous sea level from a
paleoshoreline Journal of Geophysical Research, vol. 96, p 6591-6607.
Meats, C., 1996, An appraisal of the problems involved in three-dimensional ground
penetrating radar imaging of archaeological features. Archaeometry, vol. 38, No. 2, p.
359379.
Metzeger, S. L., 1982, Subsurface paleoenvironmental analysis of gas-producing Medina
Group (Lower Silurian), Chautauqua County, New York; American Association of
Petroleum Geologist, Bulletin, v. 55, p. 1249-1261
Meyers, R. A., D. G. Smith, H. M. Jol, and C. D. Peterson, 1996, Evidence for eight great
earthquake-subsidence events detected with ground-penetrating radar, Willapa barrier,
Washington. Geology 24, p. 99102.
Miall, A. D., 1978c, Lithofacies types and vertical profile models in braided river
deposits: a summary. In: A.D. Miall (Editor), Fluvial Sedimentology. Canadian Society
of Petroleum Geology Memoir. 5: p. 597-604.
Miall, A. D., 1985, Architectural-element analysis: A new method of facies analysis
applied to fluvial deposits: Earth Science Reviews, vol. 22, p. 261308.
Miall, A. D., 1996, The geology of fluvial deposits: Sedimentary facies, basin analysis
and petroleum geology: Heidelberg, Springer-Verlag Inc., 582 p.
Miall, A. D., 2006, Reconstructing the architecture and sequence Stratigraphy of the
preserved fluvial record as a tool for reservoir development: A reality check. AAPG
Bulletin, vol. 90, no. 7(July2006), p. 9891002

202

Middleton, L.T., and R. C. Blakey, 1983, Processes and controls on the intertonguing of
the Kayenta and Navajo formations, northern Arizona: Eolianfluvial interactions, in
Brookfield, M.E., and Ahlbrandt, T.S., eds., Eolian Sediments and Processes:
Amsterdam, Elsevier, Developments in Sedimentology, no. 38, p. 613634
Mitchum, R. M., P. R. Vail, and J. B. Sangree, 1977, Stratigraphic interpretation of
seismic reflection patterns in depositional sequences. In: Payton, C.E. (Ed.), Seismic
StratigraphyApplications to Hydrocarbon Exploration. AAPG Mem. 16, p. 117 123.
Mjos, R., O. Walderhaug, and E. Prestholm, 1993, Crevasse splay sandstone geometries
in the Middle Jurassic Ravenscar Group of Yorkshire, UK. Special Publication of the
International Association of Sedimentologists, vol.17, p.167-184
Molina-Garza, R., S. Geissman, W. John, R. Van der Voo, S. G. Lucas, S. N. Hayden,
1991, Paleomagnetism of the Moenkopi and Chinle formations in central New Mexico;
implications for the North American apparent polar wander path and Triassic
magnetostratigraphy. Journal of Geophysical Research, vol. 96, no. B9, p. 14,239-14,262
Mller, I., and D. Anthony, 2003, A GPR study of sedimentary structures within a
transgressive coastal barrier along the Danish North Sea Coast. In: Bristow, C.S., Jol,
H.M. (Eds.), Ground Penetrating Radar in Sediments. Geological Society of London
Special Publication. 211, p. 55 65.
Moran, M., S. A. Arcone, A. J. Delaney, and R. Greenfield, 1998, 3-D migration/array
processing using GPR data. Proceedings of the 7th International Conference on Ground
Penetrating Radar (GPR98), USA, p. 225231.
Nadon, G. C., 1991, Thrust controlled foreland basin fluvial architecture, St. Mary River
Formation, southwestern Alberta, Canada. AAPG Bulletin, vol.75, no.3, p.644.
Nadon, G. C., 1994, The genesis and recognition of anastomosed fluvial depositsdata
from the St-Mary River Formation, Southwestern Alberta, Canada: Journal of
Sedimentary Research, vol. 64, p. 451463.

203

Neal, A., 1994, Ground-penetrating radar and its use in sedimentology: principles,
problems and progress. Earth-Science Reviews 66, p. 261330
Neal, A., R. V. Dackombe, C. L. Roberts, 2001, Applications of ground-penetrating radar
(GPR) to the study of coarse clastic (shingle) coastal structures. In: Packham, J.R.,
Randall, R.E., Barnes, R.S.K., Neal, A. (Eds.), Ecology and Geomorphology of Coastal
Shingle. Westbury Academic and Scientific, Yorkshire, p. 77 106.
Neal, A., and C. L. Roberts, 2000, Applications of ground-penetrating radar (GPR) to
sedimentological, geomorphological and geoarchaeological studies in coastal
environments. In: Pye, K., Allen, J.R.L. (Eds.), Coastal and Estuarine Environments:
Sedimentology, Geomorphology and Geoarchaeology. Geological Society of London
Special Publication 175, p. 139 171.
Neal, A., and C. L. Roberts, 2001, Internal structure of a trough blowout, determined
from migrated ground-penetrating radar profiles. Sedimentology, 48, p. 791 810.
Nichols, G. J., and J. A. Fisher, 2007, Processes, facies and architecture of fluvial
distributary system deposits: Sedimentary Geology, vol. 195, p. 7590.
Nio, S., and D. Yang, 1993, High frequent cycles and sequence stratigraphy; new aspects
for hydrocarbon field development and re-evaluation. AAPG Bulletin, vol. 77. no 2., p.
338
Nobes, D. C., R. J. Ferguson, and G. J. Brierley, 2001, Ground-penetrating radar and
sedimentological analysis of Holocene floodplains: insight from the Tuross valley, New
South Wales. Australian Journal of Earth Sciences. 48, p. 347 355.
North C. P., G. C. Nanson, and S. D. Fagan, 2007, Recognition of the sedimentary
architecture of dryland anabranching (Anastomosing) Rivers. Journal of Sedimentary
Research, 2007, vol. 77, p. 925938
Oden, C. P., M. H. Powers, D. L.Wright, and G. R. Olhoeft, 2007, Improving GPR image
resolution in lossy ground using dispersive migration. IEEE Transactions on Geoscience
and Remote Sensing, vol. 45, No. 8, p. 24922500.
204

ONeal, M. L., and S. McGeary, 2002, Late Quaternary stratigraphy and sea-level history
of the northern Delaware Bay margin, southern New Jersey, USA: a ground penetrating
radar analysis of composite Quaternary coastal terraces. Quaternary Science Reviews. 21,
p. 929 946.
Olsen, H., P. H. Due, and L. B. Clemmensen, 1989, Morphology and genesis of
asymmetric adhesion wartsa new adhesion surface structure: Sedimentary Geology,
vol. 61, p. 277285.
Parrish, J. T., and F. Peterson, 1988, Wind directions predicted from global circulation
models and wind directions determined from eolian sandstones of the Western United
States a comparison: Sedimentary Geology, vol. 56, p. 261 - 282
Payan, I., and N. Kunt, 1982, Subsurface radar signal deconvolution. Signal Process. 4, p.
249262.
Pedley, H. M., I. Hill, P. Denton, and J. Brasington, 2000, Three-dimensional modelling
of a Holocene tufa system in the Lathkill Valley, north Derbyshire, using groundpenetrating radar. Sedimentology. 47, p. 721 737.
Pees, S. T., 1986, Geometry and Petroleum Geology of the Lower Silurian Whirlpool
Formation, portion of NW Pennsylvania and NE Ohio; Northeastern Geology, vol.8, p.
171-200
Pettinelli, J. D., J. D. Redman, A. L. Endres, A. P. Annan, and G. B. Johnston, 1994,
GPR response quantification: Initial processing and model testing. Proceedings of the 5th
International Conference on Ground Penetrating Radar (GPR94), Canada.
Pindell, J. and J. F. Dewey, 1982, Permo-Triassic reconstruction of western Pangea and
the evolution of the Gulf of Mexico/Carribbean region: Tectonics, vol. 1, p. 179 - 211
Pipan, M., I. Finetti, and F. Ferigo, 1996, Multi-fold GPR techniques with applications to
highresolution studies: Two case histories. European Journal of Environmental and
Engineering Geophysics, vol. 1, p. 83103.

205

Pitrowski, R. G., 1981, Geology and natural gas production of the Lower Silurian Medina
Group and equivalent rock unit in Pensylvania; Pensylvania Geological Survey, Mineral
Resources Report 82.
Plint, A. G., 1996, Marine and nonmarine systems tracts in fourth-order sequences in the
Early-Middle Cenomanian, Dunvegan Alloformation, northeastern British Columbia,
Canada. In: High Resolution Sequence Stratigraphy: Innovations Applications (Ed. by
J.A. Howell and J.F. Aitken), Geological Society of London Special Publication, 104, p.
159-191.
Plint, A. G., 1997, Interacting tectonic and eustatic controls on regional paleogeography,
incised valley systems and sequence geometry: Cenomanian Dunvegan Formation,
Alberta foreland basin, Canada. In: Proceedings of the 1st Latin American Congress of
Sedimentology, 16-19 November, Margarita Island, Venezuela. Abstract volume, p. 299306.
Plint, A. G., 2000, Sequence stratigraphy and paleogeography of a Cenomanian deltaic
complex: the Dunvegan and lower Kaskapau formations in subsurface and outcrop,
Alberta and British Columbia, Canada. Bulletin of Canadian Petroleum Geology, 48, p.
4379.
Plint, A. G., 2002, Paleo-valley systems in the Upper Cretaceous Dunvegan Formation,
Alberta and British Columbia. Bulletin of Canadian Petroleum Geology, 50, p. 277296.
Plint, A. G. and D. Nummedal, 2000, The falling stage systems tract: recognition and
importance in sequence stratigraphic analysis. In: Sedimentary Responses to Forced
Regression (Eds D. Hunt and R.L. Gawthorpe), Spec. Publ. Geol. Soc. London, 172, p.
117.
Plint, A. G., and J. A. Wadsworth, 2003, Sedimentology and palaeogeomorphology of
four large valley systems incising delta plains, western Canada Foreland Basin:
Implications for mid-Cretaceous sea-level changes. Sedimentology, vol. 50, n 6, p 11471186

206

Plint, A. G., A. Tyagi, M. J. Hay, B. L. Varban, H. Zhang, X. Roca, 2009, Clinoforms,


paleobathymetry, and mud dispersal across the Western Canada Cretaceous foreland
basin; evidence from the Cenomanian Dunvegan Formation and contiguous strata, In:
Journal of Sedimentary Research, 79, 3, p. 144-161
Posamentier, H. W., and V. Kolla, 2003, Seismic geomorphology and stratigraphy of
depositional elements in deep-water settings. Journal of Sedimentary Research vol. 73,
no. 3, p. 367-388
Posamentier, H. W., R. J. Davies, J. A. Cartwright, and L. Wood, 2007, Seismic
geomorphology - an overview. Geological Society, London, Special Publications 277: p.
1-14.
Pratt, B. R., and A. D. Miall, 1993, Anatomy of a bioclastic grainstone megashoal
(Middle Silurian, southern Ontario) revealed by ground-penetrating radar. Geology 21, p.
223 236.
Prochnow, S. J., L. C. Nordt, S. C. Atchley, M. R. Hudec, 2006b, Multi-proxy paleosol
evidence for Middle and Late Triassic climate trends in eastern Utah, In:
Palaeogeography, Palaeoclimatology, Palaeoecology, 232, 1, p. 53-72
Prochnow, S. J., S. C. Atchley, L. C. Nordt, T. E. Boucher, and M. R. Hudec, 2006, The
influence of salt withdrawal subsidence on palaeosol maturity and cyclic fluvial
deposition in the Upper Triassic Chinle Formation: Castle Valley, Utah. Sedimentology,
vol. 53, no. 6, p. 1319-1345
Pryor, W. A., 1973, Permeability-porosity patterns and variations in some Holocene sand
bodies. AAPG Bulletin, vol. 57, Number 1, p. 162-189, January 1973.
Pyrcz, M. J., O. Catuneanu, and C. V. Deutsch, 2005, Stochastic surface-based modeling
of turbidite lobes: American Association of Petroleum Geologists, Bulletin, vol. 89, p.
177191.

207

Quinlan, G. M. and C. Beaumont, 1984, Appalachian thrusting, lithospheric flexure, and


the Paleozoic stratigraphy of the eastern interior of North America. Can. J. Earth Sci., 21:
p. 973-996.
Rahmani, R. A., and V. Schmidt, 1982, Oldman River Section (Locality A) in Shawa, M.
S, Sedimentology of Upper Cretaceous fluviatile, deltaic and shoreline deposits (CSPG
Field Trip Guidebook.
Reading, H. G., 1996, Introduction. In: Reading, H.G. (Ed.), Sedimentary Environments:
Processes, Facies and Stratigraphy, 3rd edition. Blackwell, Oxford.
Reynolds, J. M., 1997, An Introduction to Applied and Environmental Geophysics.
Wiley, Chichester.
Riediger, C. L., R. Macdonald, M. G. Fowler, L. R. Snowdon, and M. D. Sherwin, 1999,
Origin and alteration of Lower Cretaceous Mannville Group oils from the Provost oil
field, east central Alberta, Canada. Bulletin of Canadian Petroleum Geology, vol. 47, no.
1, p. 43-62.
Roberts, R. L., and J. J. Daniels, 1996, Analysis of GPR polarization phenomena. J.
Environ. Eng. Geophys. 1, p. 139 157.
Robinson, J. W., and P. J. McCabe, 1997, Sandstone-body and shale-body dimensions in
a braided fluvial system: Salt Wash Sandstone Member (Morrison Formation), Garfield
County, Utah: American Association of Petroleum Geologists Bulletin, vol. 81, p. 112671291.
Rubin, D. M., and R. E. Hunter, 1983, Reconstructing bedform assemblages from
compound crossbedding (geometry-reconstruction). Eolian sediments and processes, p
407-427
Rubin, D. M., and R. E.Hunter, 1985, Why deposits of longitudinal dunes are rarely
recognized in the geologic record. Sedimentology, 32, 1, p. 147-157

208

Rubin, Y. and S. S. Hubbard, 2005, Introduction to Hydrogeophysics in Y. Rubin and S.


S. Hubbard (eds.) Hydrogeophysics, 321. Springer, Netherlands.
Russell, A. J., O. Knudsen, H. Fay, P. M. Marren, J. Heinz, and J. Tronicke, 2001,
Morphology and sedimentology of a giant supraglacial, ice-walled, jokulhlaup channel,
Skeidararjokull, Iceland: implications for esker genesis. Global Planet. Chan. 28, p.
193 216.
Rutka, M. A., 1986, Sedimentology and Petrography of the Whirlpool Sandstone (Lower
Silurian) in outrop and the subsurface in southern Ontario and upper New York State;
unpublished Msc thesis, McMaster University, Hamilton, Ontario 355 p
Rutka, M. A., R. J. Cheel, G. V. Middleton, and C. J. Salas, 1991, The Lower Silurian
Whirlpool Sandstone; in Sedimentary and Depositional Environments of Silurian Strata
of the Niagara Escarpment, Ontario and New York. Geological Association of CanadaMineralogical Association of Canada-Society of Economic Geologists, Joint Annual
Meeting, Field Trip B4; Guidebook, p. 27-34.
Ryder, R. T., K. L. Aggen, R. D. Hettinger, B. E. Law, J. J. Miller, V. F. Nuccio, W. J.
Perry Jr., S. E. Prensky, J. R. Sanfilipo, and C. J. Wandrey, 1996, Possible continuoustype (unconventional) gas accumulation in the Lower Silurian Clinton sands, Medina
Group, and Tuscarora Sandstone in the Appalachian basin: a progress report of 1995
project activities: U.S. Geological Survey, Open-File Report 96-42, 82 p
Sahagian, D., and M.Jones, 1993, Quantified Middle Jurassic to Paleocene eustatic
variations based on Russian Platform stratigraphy: stage level resolution. Geological
Society of America Bulletin, vol. 105, n 8, p. 1109-1118, 1993
Sanford, B.V., 1969, Silurian of southwestern Ontario; Ontario Petroleum Institute,
Proceedings, vol. 8
Sangree, J. B., and J. M. Widmier, 1979, Interpretation of depositional facies from
seismic data. Geophysics 44, p. 131160.

209

Schumm, S. A., 1968, Speculations concerning paleohydrologic controls of terrestrial


sedimentation: Geological Society of America, Bulletin, vol. 79, p. 15731588.
Schumm, S. A., 1993, River response to base level change: implications for sequence
stratigraphy: Journal of Geology, vol. 101, p. 279294.
Sena, A. R., P. L. Stoffa, and M. K. Sen, 2006, Split-step Fourier migration of GPR data
in lossy media. Geophysics, vol. 71, No. 4, p. K77K91.
Snchal, P., H. Perroud, G. Snchal, 2000, Interpretation of reflection attributes in a 3D GPR survey at Vallee dOssau, western Pyrenees, France. Geophysics 65, p. 1435
1445
Seyler, B., 1981. The Whirlpool Sandstone (Medina Group) in outcrop and the
subsurface; a desription and interpretation of environment and deposition. Unpublished
M.Sc thesis, State University of New York at Fredonia, New York.
Shanley, K.W. and P. J. McCabe, 1993, Alluvial architecture in a sequence stratigraphic
framework: a case history from the Upper Cretaceous of southern Utah, USA, The
geological modelling of hydrocarbon reservoirs and outcrop analogues, p. 21-55
Sheriff, R. E., and L. P. Geldart, 1995, Exploration Seismology, 2nd edition. University
Press, Cambridge.
Skelly, R. L., C. S. Bristow, and F.G. Ethridge, 2003, Architecture of channel-belt
deposits in an aggrading shallow sandbed braided river: the lower Niobrara River,
northeast Nebraska. Sedimentary Geology, 158, p. 249270
Smith, D. G., 1983, Anastomosed fluvial deposits: modern examples from Western
Canada, in Collinson, J. D., and J. Lewin, eds., Modern and Ancient Fluvial Systems:
International Association of Sedimentologists, Special Publication 6, p. 155168.
Smith, D. G., 1986, Anastomosing river deposits, sedimentation rates and basin
subsidence, Magdalena River, northwestern Colombia, South America: Sedimentary
Geology, vol. 46, p. 177196.
210

Smith, D. G., and H. M. Jol, 1992, Ground-penetrating radar investigation of a Lake


Bonneville delta, Provo level, Bingham City, Utah. Geology 20, p. 1083 1086.
Smith, D. G., R. A. Meyers, and H. M. Jol, 1999, Sedimentology of an upper-mesotidal
(3.7 m) Holocene barrier, Willapa Bay, SW Washington, U.S.A. J. Sediment. Res. 69, p.
12901296.
Smith, D. G., and N. D. Smith, 1980, Sedimentation in anastomosed river systems:
examples from alluvial valleys near Banff, Alberta: Journal of Sedimentary Petrology,
vol. 50, p. 157164.
Smith, D.G., 1983, Anastomosed fluvial deposits: modern examples from Western
Canada, in Collinson, J.D., and Lewin, J., eds., Modern and Ancient Fluvial Systems:
International Association of Sedimentologists, Special Publication 6, p. 155168.
Smith, D.G., 1986, Anastomosing river deposits, sedimentation rates and basin
subsidence, Magdalena River, northwestern Colombia, South America: Sedimentary
Geology, v. 46, p. 177196.
Sambrook Smith, G., , J.Best, C. Bristow,and G., Petts, 2006. Braided Rivers: Where
have we Come in 10 Years? Progress and Future Needs in Sambrook Smith, G., , J.Best,
C. Bristow,and G., Petts, Editor(s): Special Publication Number 36 of the International
Association of Sedimentologists.
Sneh, A., 1983, Desert stream sequences in the Sinai Peninsula. Journal of Sedimentary
Petrology 53: p. 1271-1280
Steele, R. P., 1983, Longitudinal draa in the Permian Yellow Sands of north-east England
(aeolian bedforms). Eolian sediments and processes, p. 543-550
Stephens, M., 1994, Architectural element analysis within the Kayenta Formation (Lower
Jurassic) using ground-probing radar and sedimentological profiling, southwestern
Colorado. Sedimentary Geology, 90, p. 179 211.
211

Stewart, J. H., F. G. Poole, and R. F. Wilson, 1972, Stratigraphy and origin of the Chinle
Formation and related Upper Triassic strata in the Colorado Plateau region. U.S.
Geological Survey Professional Paper, 690:336 pp.
Stewart, J. H. F. G. Poole, and R. F. Wilson, 1972a, Changes in Nomenclature of the
Chinle Formation on the Southern Part of the Colorado Plateau; 1850s-1950s, Museum of
Northern Arizona Bulletin, 47, p. 75-103, Investigations in the Triassic Chinle Formation.
Stewart, J. H., T. H. Anderson, and G. B. Haxel, 1986, Late Triassic paleogeography of
the southern cordillera: the problem of a source for voluminous volcanic detritus in the
Chinle formation of the Colorado Plateau region. Geology, vol. 14, p. 567-570
Stokes, W. L., 1950, Pediment concept applied to Shinarump and similar conglomerates.
Geological Society of America Bulletin, 61 : p. 91-98.
Stott, D. F., 1982, Lower Cretaceous Fort St. John Group and Upper Cretaceous
Dunvegan Formation of the Foothills and Plains of Alberta, British Columbia, District of
Mackenzie and Yukon Territory. Bull. Geological Survey of Canada, 328.
Stow, D. A. V., and M. Johansson, 2000, Deep-water massive sands: nature, origin and
hydrocarbon implications: Marine and Petroleum Geology, vol. 17, p. 145174.
Streich, R., J. van der Kruk, and A. G. Green, 2007, Vector-migration of standard
copolarized 3D GPR data. Geophysics, vol. 72, No. 5, p. J65J75.
Sun, J. and R. A. Young, 1995, Recognizing surface scattering in ground-penetrating
radar data. Geophysics, vol. 60, No. 5, p. 13781385.
Szerbiak, R. B., G. A. McMechan, R. Corbeanu, C. Forster, S. H. Snelgrove, 2001, 3-D
characterization of a clastic reservoir analogue: from 3-D GPR data to a 3-D fluid
permeability model. Geophysics 66, p. 10261037.
Tanner, L. H., S. G. Lucas, S. C. Semken, W. R. Berglof, D. S. Ulmer-Scholle eds.,
2003a, Possible tectonic controls on Late Triassic sedimentation in the Chinle Basin,

212

Colorado Plateau In: Guidebook - New Mexico Geological Society, 54, p. 263-267,
Geology of the Zuni Plateau.
Tanner, L. H., 2003b, Pedogenic evidence of Late Triassic Chinle Group, Four Corners
region: evidence of Late Triassic aridification, in Lucas, S.G., ed., Geology of the Zuni
Plateau: New Mexico Geological Society, Guidebook, 54th Field Conference, p. 269
280.
Telford, P. G., 1978, Silurian stratigraphy of the Niagara Escarpment, Niagara Falls to
the Bruce Peninsula
Tesmer, I. H., 1981, Colossal Cataract, the geologic history of Niagara Falls. New York
press. 219 p.
Therrien, F., and D. E. Fastovsky, 2000, Paleoenvironments of Early Theropods, Chinle
Formation (Late Triassic), Petrified Forest National Park, Arizona: Palaios, vol. 15, p.
194211.
Thomas, R. G., D. G. Smith, J. M. Wood, J. Visser, E. A. Calverley- Range, and E. H.
Koster, 1987, Inclined heterolithic stratification - terminology, description, interpretation
and significance. Sedimentary Geology, vol. 53, no. 1-2, p. 123-179,
Tillard, T., and J. C. Dubois, 1992, Influence and lithology on radar echoes: analysis with
respect to electromagnetic parameters and rock anisotropy: Fourth International
Conference on Ground Penetrating Radar June 8-13, 1992. Rovaniemi, Finland.
Geological Survey of Finland, Special Paper 16, 365 pages.
Tillard, S., and J. C. Dubois, 1995, Analysis of GPR data: wave propagation velocity
determination. Journal of Applied Geophysics. 33, p. 77 91.
Todoeschuck, J. P., P. T. LaFle`che, O. G. Jensen, A. S. Judge, and J. A. Pilon, 1992,
Deconvolution of ground probing radar. In: Pilon, J. (Ed.), Ground Penetrating Radar.
Geological Survey of Canada Paper. 90-4, p. 227 230.

213

Tornqvist, T. E., M. H. M. Vanree, and E. L. J. H. Faessen, 1993, Longitudinal facies


architectural changes of a Middle Holocene anastomosing distributary system (Rhine
Meuse delta, central Netherlands): Sedimentary Geology, vol. 85, p. 203219.
Topp, G. C., J. L. Davis, and A. P. Annan, 1980, Electromagnetic determination of soil
water content: measurements in coaxial transmission lines. Water Resources Research 16,
p. 574 582.
Tronicke, J., P. Dietrich, U. Wahlig, and E. Appel, 2002a, Integrated surface georadar
and crosshole radar tomography: a validation experiment in braided stream deposits.
Geophysics 67, p. 1516 1523.
Turnbridge I. P., 1981, Sandy high-energy flood sedimentation some criteria for
recognition, with an example from the Devonian of Southwest England. Sedimentary
Geology 28: p. 79-96
Turner, G., 1994, Subsurface radar propagation deconvolution. Geophysics 59, p. 215
223.
Tyler, N., and R. J. Finley, 1991, Architectural controls on the recovery of hydrocarbons
from sandstone reservoirs. Concepts in Sedimentology and Paleontology, vol.3, p. 1-5
Ursin, B., 1983, Review of elastic and electromagnetic wave propagation in horizontally
layered media. Geophysics 48, p. 10631081.
Van Dam, R. L., 2001, Causes of ground-penetrating radar reflections in sediment.
Unpubl. PhD Thesis, University of Amsterdam.
Van Dam, R. L., S. L. Nichol, P. C. Augustinus, K. E. Parnell, P. L. Hosking, and R. F.
McLean, Sep. 2003, GPR stratigraphy of a large active dune on Parengarenga Sandspit,
New Zealand. The Leading Edge, vol. 22: p. 865 - 870.
Van Dam, R. L., and W. Schlager, 2000, Identifying causes of ground-penetrating radar
reflections using time-domain reflectometry and sedimentological analyses.
Sedimentology 47, p. 435449.
214

Van Dam, R. L., W. Schlager, M. J. Dekkers, and J. A. Huisman, 2002a, Iron oxides as a
cause of GPR reflections. Geophysics 67, p. 536545.
Van Dam, R. L., E. H. van den Berg, M. G. Schaap, L. H. Broekema, and W. Schlager,
2003, Radar reflections from sedimentary structures in the vadose zone. In: Bristow, C.
S., and H. M. Jol, (Eds.), Ground Penetrating Radar in Sediments. Geological Society of
London Special Publication. 211, p. 257273.
Van Gestel, J. and P. L. Stoffa, 2000, Migration using multi-configuration data.
Proceedings of the 8th International Conference on Ground Penetrating Radar (GPR
2000), Australia, p. 448452.
Van Heteren, S., D. M. Fitzgerald, D. C. Barber, J. T. Kelley, and D. F. Belknap, 1996,
Volumetric analysis of a New England barrier system using ground-penetrating radar and
coring techniques. J. Geol. 104, p. 471 483.
Van Heteren, S., D. M. Fitzgerald, P. A. McKinlay, and I. V. Buynevich, 1998, Radar
facies of paraglacial barrier systems: coastal New England, USA. Sedimentology 45, p.
181200.
Van Overmeeren, R. A., 1998, Radar facies of unconsolidated sediments in The
Netherlands: a radar stratigraphy interpretation method for hydrogeology. Journal of
Applied Geophysics. 40, p. 1 18.
Van Overmeeren, R. A., S. V. Sariowan, and J. C. Gehrels, 1997, Ground penetrating
radar for determining volumetric soil water content; results of comparative measurements
at two test sites. Journal of Hydrology. 197, p. 316 338.
Weber, K. J., 1987, Computation of initial well productivities in aeolian sandstone on the
basis of a geological model, Leman gas field, U.K. Special Publication - Society of
Economic Paleontologists and Mineralogists, 40, p. 333-354, Reservoir sedimentology
Weerts, H. J. T., and M. F. P. Bierkens, 1993, Geostatistical analysis of overbank
deposits of anastomosing and meandering fluvial systems; RhineMeuse delta, The
Netherlands: Sedimentary Geology, vol. 85, p. 221232.
215

Wescott, W. A., 1993, Geomorphic thresholds and complex response of fluvial systems
some implications for sequence stratigraphy. Bull. AAPG, 77, p. 12081218.
Sedimentology, Statistics, and Flow Behavior for a Tide-influenced Deltaic Sandstone,
Frontier Formation, Wyoming, United States Christopher D. White,Brian J. Willis,
Shirley P. Dutton, Janok P. Bhattacharya, Keshav Narayanan.
White, C. D., B. J. Willis, S. P. Dutton, J. P. Bhattacharya, and K. Narayanan, 2004,
Sedimentology, statistics, and flow behavior for a tide-influenced deltaic sandstone,
Frontier Formation, Wyoming, United States. AAPG Memoir, vol.80, p. 129-152
Williams, G. D., and C. R. Stelck, 1975, Speculations on the Cretaceous palaeogeography
of North America. Special Paper - Geological Association of Canada, 13, p. 1-20
Williams, M. Y., 1919, The Silurian geology and faunas of Ontario Peninsula, and
Manitoulin and adjacent islands; Geological Survey of Canada Memoir 111
Williams, G.D. and Stelck, C.R. 1975. Speculations on the Cretaceous palaeogeography
of North America. In: The Cretaceous System of the Western Interior of North America.
W.G.E. Caldwell (ed.). Geological Association of Canada, Special Publication No. 13, p.
1-20.
Willis, B. J., and S. L. Gabel, 2003, Formation of deep incisions into tide-dominated river
deltas: Implications for the stratigraphy of the Sego Sandstone, book Cliffs, Utah, U.S.A.
Journal of Sedimentary Research. vol. 73, no. 2, p. 246-263
Wilson, A. W. G., 1903, The theory of the formation of sedimentary deposits; Canada
Record of Science, vol. 9. p. 112-132
Woodward, J., P. J. Ashworth, J. L. Best, G. H. Sambrook Smith, and C. J. Simpson,
2003, The use and application of GPR in sandy fluvial environments: methodological
considerations. In: Bristow, C. S., H. M. Jol (Eds.), Ground Penetrating Radar in
Sediments. Geological Society of London Special Publication. 211, p. 127142.

216

Wooldridge, C. L., and E. J. Hickin, 2005, Radar architecture and evolution of channel
bars in wandering gravel-bed rivers: Fraser and Squamish Rivers, British Columbia,
Canada. Journal of Sedimentary Research. vol. 75, p. 844-860
Xia, J., E. K. Franseen, R. D. Miller, T. V. Weis, and A. P. Byrnes, 2003, Improving
ground-penetrating radar data in sedimentary rocks using deterministic deconvolution. J.
Appl. Geophysics.
Yilmaz, O., 1987, Seismic Data Processing. Investigations in Geophysics, vol. 2. Society
of Exploration Geophysics, Tulsa.
Yilmaz, O., 2001, Seismic Data Analysis, vol. 1. Society of Exploration Geophysics,
Tulsa.
Yu, H., X. Ying, and Y. Shi, 1996, The use of FK techniques in GPR processing.
Proceedings of the 6th International Conference on Ground Penetrating Radar (GPR96),
Japan, p. 595600.
Zeng, X., G. A. McMechan, J. P. Bhattacharya, C. L. V. Aiken, X. Xu, W. S. Hammon,
III, and R. M. Corbeanu, 2004, 3-D imaging of a reservoir analogue in point bar deposits
in the Ferron Sandstone, Utah, using ground-penetrating radar: Geophysical Prospecting,
vol. 52, p. 151163.
Ziegler, A. M., K. S. Hansen, M. E. Johnson, M. A. Kelly, C. R. Scotese, and R. Van Der
Roo, 1977, Silurian continental distributions, paleogeography, climatology, and
biogeography. Tectonophysics, 40: p. 1351.
Ziegler, A. M., C. R. Scotese, and S. F. Barrett, 1983, Mesozoic and Cenozoic
Paleogeographic maps: in Brioshe, P., and Sundermann, J., eds., Tidal friction and the
earths rotation. II: Spring-Verlag, Berlin, p. 240 - 252

217

218

Appendix 1
Table A1: Typical conductivity, relative permittivity and radar velocity data in sediment
(Annan, 2004)

219

Appendix 2
GPR Instrumentation
Considering the surging interest in GPR imaging of modern and ancient sedimentary
deposit, a brief discussion of GPR instrumentation is pertinent to understanding the
capabilities and limitations of the various GPR systems available in order to achieve
efficiently- planned GPR surveys. The three major manufacturers of GPR equipment are
Sensors and Software, MAL Geoscience, and Geophysical survey Systems Inc. (GSSI).
Sensors and Software and MAL Geoscience equipment are used more for GPR surveys
in Canada as these companies have offices/distributing agents in Canada. Each of these
companies manufactures array of GPR equipment designed for specific survey type.
Sensors and Softwares Inc., a Mississuaga (Canada) based GPR company manufactures
Noggin, PulseEKKO Pro, Conquest and SnowScan. Conquest is used mostly for concrete
imaging while Snowscan is used to evaluate snow thickness. Noggin and PulseEKKO
Pro are often used for geological applications. Noggin Smart Systems are integrated
ground penetrating radar (GPR) data acquisition platforms designed for shallow target,
high resolution imaging; they range in frequency from 250 MHz to 1000 MHz. The
Noggin systems are available in three different configurations: the SmartCart system, the
SmartHandle system and the Rock Noggin. Data acquisition is done simply by pushing
(or pulling) the Smart System along the survey line. This is usually done with an
odometer used as the triggering device, however, it is possible to change the triggering
method and run the Smart System in continuous operation or using the trigger button
(step mode). Noggin systems, because they are either pulled or dragged save survey time;
they are however only adapted for continuous mode data collection and Common Offset
220

surveys, not Common Midpoint (CMP) surveys usually required for time-depth
conversion of GPR profiles.

Figure A1: Mode of operation of Noggin system (Image taken from


http://www.sensoft.ca/products/noggin/noggin.html - Accessed August 7, 2010)
PulseEKKO Pro antennas are low frequency, unshielded, bistatic antennas designed for
lower resolution, deeper target surveys. The antennas come in pairs; one transmitting
antenna and one receiving antenna. The shortest antennas, the 200 MHz, are 0.5 metres
long while the longest; the 12.5 MHz are 8 metres long. The 12.5 MHz antenna is the
lowest frequency currently available for GPR imaging and offers the deepest target
capability possible for any GPR survey. The length of the antennas (8 metres each)
however, makes surveys with these antennas cumbersome and time consuming (Figure
4). The pulseEKKO PRO can also be used with high frequency (100MHz, 250MHz, 500
MHz), shielded, bistatic antennas designed for high resolution shallow target imaging.
Each antenna box is really a transducer because it consists of both an antenna and the
electronics (Figure A2). The transmitting transducer is indicated with a T on the label,
e.g. T500, while the receiving transducer is indicated with an R on the label, e.g. R500.

221

Figure A2: High frequency, shielded, bistatic antennas available for the pulseEKKO
PRO. The antennas come in pairs, one transmitting transducer and one receiving
transducer. These are indicated by a T and an R on the labels. The 1000 MHz
transducers are 14.5 cm (6 in) across, the 500 MHz transducers are 22.5 cm (9 in) across
and the 250 MHz transducers are 38 cm (15 in) across. (Image taken from
http://www.sensoft.ca/products/pulseekko/pulseekkopro.html - Accessed August 7, 2010)
MAL Geoscience manufactures unshielded antenna in the 25MHz to 200 MHz and
shielded antenna in the 100MHz to 1 GHz range. MALs unshielded antennas consist of
separate transmitter and receiver elements and electronics, allowing them to be operated
in several modes for different survey techniques, such as reflection profiling, velocity
profiling (common mid-point (CMP), or wide angle reflection and refraction (WARR)),
and cross-scanning (tomography). MALs unshielded antennas are designed for use in
applications that require maximum depth penetration. However, due to the fact that these
antennas are unshielded, they are more suitable for use in areas with little or no
background noise. Wooden handles are provided for use with the 50, 100 and 200 MHz
antennas; these are used to carry the antennas and to also keep them separated and
stabilized during measurements (Figure A3)

222

Figure A3: MAL Geoscience unshielded Bistatic antennas (Image taken from
http://www.malags.com/Downloads/Product-Brochures.aspx - Accessed August 7, 2010)
Due to the size of the 25 MHz unshielded antenna elements, wooden handles are not
practical, so strapping is provided to aid in separation and carrying. MAL also
manufactures low frequency (25 MHz , 50 MHz and 100 MHz) Rough Terrain Antennas.
The Rough Terrain Antenna (RTA) series from MAL affords flexibility required in
low-frequency Ground Penetrating Radar (GPR) surveying in rough terrains typical of
outcrops. The antenna design provides improved performance for deeper penetration and
the flexible snake like design allows the antenna to be maneuvered easily and
efficiently through dense vegetation or most uneven of terrain without affecting ground
contact, providing optimum results in the most difficult of environments; the most
important benefit being that the operator doesnt have to clear an access path or route

223

prior to the profile or survey. MAL Geoscience claims reductions in survey time and
manpower requirements to a third of those experienced when using traditional unshielded
antennas. However, as the RTA is an antenna of unshielded type, unwanted noise most
often in form of air reflections can also occur in the data due to the fact that the antenna
emits electromagnetic waves in any direction and also receive them from any direction.

Figure A4: Mode of operation of MAL Geoscience Rough Terrain Antenna (Image
taken from http://www.malags.com/Downloads/Product-Brochures.aspx - Accessed
August 7, 2010)
MAL shielded antennas are designed for use in urban areas or sites with a background
noise. A shielded antenna consists of both transmitter and receiver antenna elements in a
single housing. The design ensures that the transmitted radar energy is only emitted from
the bottom of the antenna housing where it is in contact with the ground and protects the
receiver element from externals signals (noise) from directions other than the bottom of
the housing where it is located (Figure A5)

224

800 MHz
500 MHz

250 MHz

100 MHz

Figure A5: MAL Geoscience shielded antennas; note increase in antenna size with
frequency (Image taken from http://www.malags.com/Downloads/ProductBrochures.aspx - Accessed August 7, 2010)
GSSI manufactures adjustable multiple low-frequency (15-80 MHz) antennas designed
for deep penetration depth. The antenna design consists of interchangeable elements; by
changing the length of the antenna, the transmission frequency is changed. This antenna
can be deployed in discrete measurements (step mode) or continuous profile data
collection modes; this affords the user some flexibility in resolution/target depth during
GPR surveys. GSSI also manufactures 100 MHz, 200 MHz and 270 MHz antennas in
addition to the high resolution antenna (>500 MHz) useful for low target high resolution
outcrop studies. The 100 MHz antenna is used for deep (10-20 m depending on the
dielectric properties of the transmitting medium) subsurface applications. The 100 MHz
antennas are produced in monostatic and bistatic format. The 100 MHz monostatic (left)
combines the transmitter and receiver electronics in a single antenna housing. The 100
MHz bistatic (right) is a versatile antenna pair that can operate in different configurations
(common offset and common mid-point) to optimize performance.
225

surface

Figure A6: GSSI 100 MHz antennas available in monostatic and bistatic mode (Image
taken from http://www.geophysical.com/antennas.htm - Accessed August 7, 2010)

226

Appendix 3
Table A3- List of Published GPR studies on Sediments from 1980 - 2010

Deposit /
Environment

Carbonate

Age /

Equipment

Formation

Type

Cretaceous

PulseEKKO

Lowest

Survey

Max Depth

Survey

Survey

Successful

Mode

of

Type

Area

Antenna

(Common

Frequenc

Offset /

y (MHz)

CMP/Both)

100

Common

Investigation

(2D

(m)

/3D /

Study Title

Publication

Authors

Year

(m )

Both)
10

Offset

2D

GPR and marine

Lars Ole

carbonate mound

Boldreel, and

structures in

Jette Vindum

southwest Sweden: A
case study of imaging
structures at different
scales

227

Lars Nielsen,

seismic imaging of

Denmark and

Table A3 continued

2003

Carbonate

Cretaceous

Unknown

400

Common

10

2D

Offset

Geoelectrical

2010

Damayanti

constraints on radar

Mukherjee,

probing of shallow

Essam Heggy,

water-saturated zones

Shuhab D. Khan

within karstified
carbonates in semiarid environments
Carbonate

Jurassic

Ramac

250

Shielded

Common

3D

Offset

Photographing layer

2008

Selma Kadioglu

2000

Guy Dagallier,

thicknesses and
discontinuities in a
marble quarry with
3D GPR visualisation

Carbonate

Jurassic

GSSI SIR-2

100

Common

40

Offset

Table A3 continued

228

2D

Ground penetrating
radar application in a

Ari I. Laitinen,

shallow marine

Fabrice Malartre,

Oxfordian limestone

Ignace P. A. M.

sequence located on

Van

the eastern flank of

Campenhout,

the Paris Basin, NE

Paul C. H.

France

Veeken

Carbonate

Mississippia

Ramac

n/

Unshielded

50

Common

40

2D

Mapping fractures

Offset

2006

Ulrich Theune,

with GPR; a case

Dean Rokosh,

Livingstone

study from Turtle

Mauricio D.

Formation

Mountain

Sacchi, and
Douglas R.
Schmitt

Carbonate

Ordovician

PulseEKKO

50

Common

14

2D

Characterization of a

Offset

2002

George A.

coalesced, collapsed

McMechan,

paleocave reservoir

Robert G.

analog using GPR

Loucks, Paul

and well-core data

Mescher, and
Xiaoxian Zeng

Carbonate

Ordovician /

Unknown

50

Ellenburger

Common

10

2D

Ground penetrating

Offset

dolomite

Carbonate

Triassic

Unknown

100

Common

20

Offset

3D

1000

1998

George A.

radar imaging of a

McMechan,

collapsed paleocave

Robert G.

system in the

Loucks,

Ellenburger dolomite,

Xiaoxian Zeng,

central Texas

Paul Mescher

Full-resolution 3D
GPR imaging

2005

Mark
Grasmueck, Ralf
Weger, and
Heinrich
Horstmeyer

Table A3 continued
229

Carbonate

Carbonate

Upper Farley

Unknown

400

Both

12

2D

Improving resolution

2007

Evan K.

(Pennsylvani

and understanding

Franseen, Alan

an)

controls on GPR

P. Byrnes,

response in carbonate

Jianghai Xia, and

strata: Implications

Richard D.

for attribute analysis

Miller

Triassic

GSSI SIR

200

10B

Common

2D

Visualization and

Offset

2003

Stefan Reiss,

characterization of

Klaus R.

active normal faults

Reicherter, and

and associated

Claus-Dieter

sediments by high-

Reuther

resolution GPR
Carbonate

Paleocene

PulseEKKO

100

Common

11

Offset

Both

3150

Application of GPR

230

Thrainn

for 3-D visualization

Sigurdsson,

of geological and

Torben

structural variation in

Overgaard

a limestone formation

Table A3 continued

1998

Turbidite

Carbonifero-

PulseEKKO

50

Both

20

Both

30000

us

3D high-resolution

2004

J. K. Pringle, A.

digital models of

R. Westerman, J.

outcrop analogue

D. Clark, N. J.

study sites to

Drinkwater, and

constrain reservoir

A. R. Gardiner

model uncertainty: an
example from Alport
Castles, Derbyshire,
UK
Deltaic

Cretaceous /

PulseEKKO

100

Both

12

3D

660

3-D characterization

2001

Robert B.

Ferron

of a clastic reservoir

Szerbiak, George

Sandstone

analog; from 3-D

A. McMechan,

GPR data to a 3-D

Rucsandra

fluid permeability

Corbeanu, Craig

model

B. Forster, and
Stephen H.
Snelgrove

Table A3 continued

231

Fluvial

Jurassic /

PulseEKKO

50

Both

15

2D

Architectural element

Kayenta

1994

Mark Stephens

2004

Ryan P. Stepler,

analysis within the


Kayenta Formation
(Lower Jurassic)
using ground-probing
radar and
sedimentological
profiling,
southwestern
Colorado

Marine

Dad

Ramac

Sandstone

Unshielded

100

Common

20

Offset

3D

3780

Three-dimensional
imaging of a deep

Alan J. Witten,

marine channel-

and Roger M.

levee/overbank

Slatt

sandstone behind
outcrop with EMI
and GPR

Table A3 continued

232

Turbidite

Carboniferou

PulseEKKO

225

Common

30000

Offset

The use of GPR to

2003

J. K. Pringle, J.

image three-

D. Clark, A. R.

dimensional (3-D)

Westerman, and

turbidite channel

A. R. Gardiner

architecture in the
Carboniferous Ross
Formation, County
Clare, western
Ireland
Marine

Lewis Shale

PulseEKKO

100

Common

2D

Offset

A hybrid laser-

2002

tracking/GPS

Roger A. Young
and Neal Lord

location method
allowing GPR
acquisition in rugged
terrain
Turbidite

Cretaceous /

PulseEKKO

100

Both

10

Lewis Shale

2D

Application of

M. Slatt, J. G.

radar imaging to

Staggs

outcrops

233

R. A. Young, R.

ground penetrating
deepwater (turbidite)

Table A3 continued

2003

Carbonate

Ordovician /

GSSI

Ellenburger

SIR10A

300

Common

15

Both

288

Offset

GPR data processing

1996

for 3D fracture

dolomite

G. Grandjean, J.
C. Gourry

mapping in a marble
quarry (Thassos,
Greece)

Deltaic

Cretaceous /

PulseEKKO

50

Ferron

Common

20

3D

Offset

Sandstone

230

Integration of GPR

2007

Keumsuk Lee,

with stratigraphic and

Mark Tomasso,

lidar data to

William A.

investigate behind-

Ambrose, and

the-outcrop 3D

Renaud

geometry of a tidal

Bouroullec

channel reservoir
analog, upper Ferron
Sandstone, Utah
Deltaic

Cretaceous /

PulseEKKO

100

Both

16

2D

Estimation of the

2002

William S.

Ferron

spatial distribution of

Hammon,

Sandstone

fluid permeability

Xiaoxian Zeng,

from surface and

Rucsandra M.

tomographic GPR

Corbeanu, and

data and core, with a

George A.

2-D example from

McMechan

the Ferron Sandstone,


Utah

Table A3 continued
234

Deltaic

Cretaceous /

PulseEKKO

100

Both

15

3D

660

Prediction of 3-D

2002

Rucsandra M.

Ferron

fluid permeability

Corbeanu,

Sandstone

and mudstone

George A.

distributions from

McMechan, and

ground-penetrating

Robert B.

radar (GPR)

Szerbiak

attributes; example
from the Cretaceous
Ferron Sandstone
Member, east-central
Utah
Barrier spit

Holocene

PulseEKKO

100

Both

15

2D

Ground water surface

2007

Curt D. Peterson,

and Beach

trends from ground-

Harry M. Jol,

plain

penetrating radar

David Percy, and

(GPR) profiles taken

Eric L. Nielsen

across late Holocene


barriers and beach
plains of the
Columbia River
littoral system,
Pacific Northwest
coast, USA

Table A3 continued
235

Coastal

Holocene

Unknown

200

Both

2D

Barrier

Coastal

Holocene

Unknown

100

Both

10

2D

Annual Layers

2004

L.J. Moore,

Revealed By GPR in

H.M. Jol, S.

the Subsurface of a

Kruse, S.

Prograding Coastal

Vanderburgh,

Barrier,Southwest

and G.M.

Washington, U.S.A.

Kaminsky

Applications of

2000

ground-penetrating

Adrian Neal and


Clive L. Roberts

radar (GPR) to
sedimentological,
geomorphological
and
geoarchaeological
studies in coastal
environments
Lake

Holocene

Ramac

50

Both

40

Unshielded

2D

Rapid lake infill

O. Sass, M.

following major

Krautblatter, and

rockfall (bergsturz)

D. Morche

events revealed by
ground-penetrating
radar (GPR)
measurements,
Reintal, German Alps

236

2007

Glacio-fluvial

Holocene

PulseEKKO

50

Both

30

2D

Gpr-derived

2007

Jonathan L.

Sedimentary

Carrivick, Jamie

Architecture and

K. Pringle,

Stratigraphy of

Andrew J.

Outburst Flood

Russell, and

Sedimentation Within

Nigel J. Cassidy

a Bedrock Valley
System, Hraundalur,
Iceland
Coastal

Holocene

Barrier

GSSI SIR-3

120

Common

Offset

2D

High-Resolution

Buynevich and

Imaging and

Duncan M.

Sedimentology of

Fitzgerald

Maine, U.S.A.:
Implications for
Holocene BackBarrier Evolution

237

Ilya V.

Subsurface (GPR)

Coastal Ponds,

Table A3 continued

2003

Coastal

Holocene

PulseEKKO

100

Barrier

Common

10

2D

Offset

A GPR study of

2003

Ingelise Moller

sedimentary

and Dennis

structures within a

Anthony

transgressive coastal
barrier along the
Danish North Sea
coast
Eolian

Holocene

Common

2D

Offset

GPR survey of a

2003

K. G. Havholm,

Holocene

N. D. Bergstrom,

aeolian/fluvial/lacustr

Harry M. Jol,

ine succession,

and G. L.

Lauder Sandhills,

Running

Manitoba, Canada
Eolian

Holocene

PulseEKKO

100

Common

15

Offset

2D

Evidence for dune

2003

Greg A. Botha,

reactivation from

Charlie S.

GPR profiles on the

Bristow, Naomi

Maputaland coastal

Porat, Geoff

plain, South Africa

Duller, Simon J.
Armitage, Helen
M. Roberts,
Brendan M.
Clarke, Mxolisi
Kota, and Philo
Schoeman

Table A3 continued
238

Pyroclastic

Holocene

PulseEKKO

200

Both

3D

GPR studies of

deposits

2001

pyroclastic deposits:

B. Cagnoli and
T. Ulrych

Subsurface
information where
there are no outcrops
Glacial

Pyroclastic

Holocene

Holocene

PulseEKKO

PulseEKKO

50

100

VRP

Both

43

2D

Downhole GPR for

2D

Lewis E. Hunter,

high-resolution

Allan J. Delaney,

analysis of material

Daniel E.

properties near

Lawson, and Les

Fairbanks, Alaska

Davis

Joint time-frequency

deposits

2003

2008

S. Guha, S.

analysis of GPR data

Kruse, and P.

over layered

Wang

sequences
Mud Volcano

Holocene

GSSI SIR-II

200

Both

Both

1000

GPR reflection

Chow, Su-Kai

depositional models

Chang, and Ho-

of mud volcanic

Shing Yu

Wushanting mud
volcano field,
southwestern Taiwan

239

Joseph Jinder

characteristics and

sediments;

Table A3 continued

2006

Carbonate

Holocene

PulseEKKO

50

Both

10

2D

The recognition of

2003

barrage and paludal

Martyn Pedley
and Ian Hill

tufa systems by GPR;


case studies in the
geometry and
correlation of
Quaternary
freshwater carbonates
Lake

Holocene

Common

2D

Offset

Amplitude analysis

2003

of repetitive GPR

Sarah E. Kruse
and Harry M. Jol

reflections on a Lake
Bonneville Delta,
Utah
Coastal

Holocene

Barrier

Ramac

500

Both

Unshielded

2D

Cyclical Evolution of

Alejo, F. Rial, H.

Transgressive Sand

Lorenzo, and

Barrier in

M.A. Nombela

Elucidated by GPR
and Aerial Photos

240

S. Costas, I.

a Modern

Northwestern Spain

Table A3 continued

2006

Deltaic

Holocene

PulseEKKO

100

Both

12

2D

A comparison of the

2000

Paulette Tercier,

correlation structure

Rosemary

in GPR images of

Knight, and

deltaic and barrier-

Harry Jol

spit depositional
environments
Coastl beach

Holocene

PulseEKKO

100

Both

2D

Quantifying rates of

2006

coastal progradation

C. S. Bristow
and K. Pucillo

from sediment
volume using GPR
and OSL; the
Holocene fill of
Guichen Bay, southeast South Australia
Coastal

Holocene

Barrier

Ramac
Unshielded

100

Common

Offset

2D

Internal structure of a

Bennett, Nigel J.

revealed by ground

Cassidy, and

penetrating radar

Jeremy Pile

UK

241

Matthew R.

barrier beach as

(GPR); Chesil Beach,

Table A3 continued

2009

Coastal

Holocene

PulseEKKO

200

Both

2D

Barrier

Digital ground

1996

Harry M. Jol,

penetrating radar

Derald G. Smith,

(GPR); a new

and Richard A.

geophysical tool for

Meyers

coastal barrier
research (examples
from the Atlantic,
Gulf and Pacific
coasts, U.S.A.)
Eolian

Holocene

GSSI SIR-8

100

Common

30

2D

Offset

Ground-penetrating

1996

Zaki Harari

2003

Remke L. Van

radar (GPR) for


imaging stratigraphic
features and
groundwater in sand
dunes

Eolian

Holocene

GSSI SIR
2000

35

Common

20

Offset

2D

GPR stratigraphy of a
large active dune on

Dam, Scott L.

Parengarenga

Nichol, Paul C.

Sandspit, New

Augustinus,

Zealand

Kevin E. Parnell,
Peter L.
Hosking, and
Roger F.
McLean

Table A3 continued

242

Eolian

Holocene

GSSI SIR

400

3000

Common

3D

Mapping the internal

Offset

2008

Ademola Q.

structure of sand

Adetunji,

dunes with GPR: A

Abdullatif Al-

case history from the

Shuhail, and

Jafurah sand sea of

Gabor Korvin

eastern Saudi Arabia


Fluvial

Holocene

225

Common

3D

Offset

400

Architecture and

2003

Andrew J.

sedimentology of an

Mumpy, Harry

active braid bar in the

M. Jol, William

Wisconsin River

F. Kean, and

based on 3-D ground-

John L. Isbell

penetrating radar
Coastal

Holocene

Unknown

100

Both

12

2D

Imaging fluvial

Hickin, Peter T.

paleovalley fill using

Bobrowsky,

ground-penetrating

Roger C. Paulen,

radar, Maple Creek,

and Mel Best

Stratigraphic analyses
using GPR)

243

Adrian S.

architecture within a

Guyana (in

Table A3 continued

2008

Eolian

Holocene

PulseEKKO

100

Both

15

2D

Combining ground

2005

C.S. Bristow, N.

penetrating radar

Lancaster, and

surveys and optical

G.A.T. Duller

dating to determine
dune migration in
Namibia
Glacio-fluvial

Holocene

PulseEKKO

100

Common

2D

Offset

The Architecture of

2005

Simone Engels

Prograding Sandy-

and Michael C.

Gravel Beach Ridges

Roberts

Formed During the


Last Holocene
Highstand:
Southwestern British
Columbia, Canada
Coastal

Holocene

Barrier

PulseEKKO

100

Common

Offset

2D

Minimum Runup

Kenneth M.

Paleotsunami from

Cruikshank,

Evidence of Sand

Harry M. Jol,

Ridge Overtopping at

and Robert B.

Cannon Beach,

Schlichting

Cascadia Margin,
U.S.A.

244

Curt D. Peterson,

Heights of

Oregon, Central

Table A3 continued

2008

Carbonate

Holocene

Unknown

50

Common

30

2D

Offset

Ground-penetrating

1995

Christopher L.

radar; a near-face

Liner and Jeffrey

experience from

L. Liner

Washington County,
Arkansas
Coastal

Holocene

Barrier

GSSI SIR-

200

2000

Common

10

2D

Offset

Coastal

2006

Environmental

Ilya V.
Buynevich

Changes Revealed in
Geophysical Images
of Nantucket Island,
Massachusetts,
U.S.A.
Coastal

Holocene

PulseEKKO

100

Common

15

Offset

2D

The structure and

Bristow, P. Neil

foredunes on a

Chroston, Simon

locally prograding

D. Bailey

ground-penetrating
radar surveys,
Norfolk, UK

245

Charlie S.

development of

coast: insights from

Table A3 continued

2000

Coastal

Holocene

GSSI SIR-3

120

Common

15

2D

Radar facies of

Offset

1988

Sytze Van

paraglacial barrier

Heteren, Duncan

systems: coastal New

M. Fitzgerald,

England, USA

Paul A.
Mckinlay, Ilya
V. Buynevich

Coastal

Holocene

Barrier

Ramac

100

Unshielded

Common

2D

Internal structure of a

Offset

2009

Matthew R.

barrier beach as

Bennett, Nigel J.

revealed by ground

Cassidy, Jeremy

penetrating radar

Pile

(GPR): Chesil beach,


UK
Eolian

Holocene

GSSI SIR-

100

Common

35

2D

Ground-penetrating

Offset

1996

Zaki Harari

2010

James D.

radar (GPR) for


imaging stratigraphic
features and
groundwater in sand
dunes

Eolian

Holocene

Ramac
Unshielded

200

Common

Offset

Both

300

Parabolic dune
reactivation and

Girardi, Dan M.

migration at

Davis

Napeague, NY, USA:


Insights from aerial
and GPR imagery

Table A3 continued

246

Fluvial

Holocene

PulseEKKO

50

Both

10

2D

Ground penetrating

1999

J. Vandenberghe,

radar images of

R. A. van

selected fluvial

Overmeeren

deposits in the
Netherlands
Eolian

Holocene

PulseEKKO

50

Both

22

2D

Unveiling past

2005

Karsten

aeolian landscapes: A

Pedersen, Lars

ground-penetrating

B. Clemmensen

radar survey of a
Holocene coastal
dunefield system,
Thy, Denmark
Alluvial Fan

Holocene

PulseEKKO

50

Common

40

Offset

2D

Ground penetrating
radar facies of the
paraglacial Cheekye
Fan, southwestern
British Columbia,
Canada

Table A3 continued

247

2001

Csaba kes,
Edward J. Hickin

Coastal

Holocene

PulseEKKO

100

Both

15

2D

Barrier

Integrating ground-

2009

L. Nielsen, I.

penetrating radar and

Mller, L.H.

borehole data from a

Nielsen, P.N.

Wadden Sea barrier

Johannessen, M.

island

Pejrup, T.J.
Andersen, J.S.
Korshj

Fluvial

Holocene

PulseEKKO

100

Both

14

2D

The internal structure

1997

of scrolled floodplain

Rene F. Leclerc,
Edward J. Hickin

deposits based on
ground-penetrating
radar, North
Thompson River,
British Columbia
Fluvial,

Holocene

PulseEKKO

50

Both

30

2D

Radar facies of

Eolian,

unconsolidated

lacustrine

sediments in The
Netherlands: A radar
stratigraphy
interpretation method
for hydrogeology

Table A3 continued

248

1998

R. A. van
Overmeeren

Deltaic

Holocene

PulseEKKO

25

Both

35

2D

Radar structure of a

1997

Gilbert-type delta,

Derald G. Smith,
Harry M. Jol

Peyto Lake, Banff


National Park,
Canada
Eolian

Holocene

GSSI

200

SIR3000

Coastal

Holocene

PulseEKKO

Common

2D

Offset

25

Both

10

Barrier

2D

The internal structure

T. Martn-

dunes of the Ebro

Crespo, I.

River Delta (Spain)

Rodrguez, M.J.

from ground

Snchez, I.

penetrating radar

Montoya

Ground penetrating

2003

Harry M. Jol,

radar: 2-D and 3-D

Don C. Lawton,

subsurface imaging

Derald G. Smith

spit, Long Beach,


WA, USA

249

D. Gmez-Ortiz,

of modern barchan

of a coastal barrier

Table A3 continued

2009

Beach

Holocene

PulseEKKO

100

Both

10

2D

Ground-penetrating

2008

Toru Tamura,

radar profiles of

Fumitoshi

Holocene raised-

Murakami,

beach deposits in the

Futoshi

Kujukuri strand plain,

Nanayama,

Pacific coast of

Kazuaki

eastern Japan

Watanabe,
Yoshiki Saito

Beach

Holocene

PulseEKKO

100

Common

10

2D

Offset

Internal architecture

2010

Lars B.

of a raised beach

Clemmensen,

ridge system (Anholt,

Lars Nielsen

Denmark) resolved
by groundpenetrating radar
investigations
Alluvial Fan

Holocene

PulseEKKO

50

Common

16

Offset

2D

Long-term bed load

Pelpola, Edward

on aerial-photo and

J. Hickin

radar surveys of fandelta growth, Coast


Mountains, British
Columbia

250

Channa P.

transport rate based


ground penetrating

Table A3 continued

2004

Coastal

Holocene

PulseEKKO

100

Both

2D

Barrier

Investigation of

2006

Adam D.

large-scale washover

Switzer, Charles

of a small barrier

S. Bristow, Brian

system on the

G. Jones

southeast Australian
coast using ground
penetrating radar
Eolian

Holocene

PulseEKKO

100

Common

20

2D

Offset

Investigation of the

2010

C.S. Bristow,

age and migration of

P.C. Augustinus,

reversing dunes in

I.C. Wallis, H.M.

Antarctica using GPR

Jol, E.J. Rhodes

and OSL, with


implications for GPR
on Mars
Eolian

Holocene

PulseEKKO

450

Both

2D

Slipfaceless

H.M. Jol, P.

a polar desert,

Augustinus, I.

Victoria Valley,

Wallis

from ground
penetrating radar

251

C.S. Bristow,

whaleback dunes in

Antarctica: Insights

Table A3 continued

2010

Eolian

Holocene

GSSI SIR-

400

2000

Coastal

Holocene

GSSI SIR 3

Common

10

2D

Offset

120

Common

15

2D

Offset

Glacio-fluvial

Fluvial

Holocene

Holocene

PulseEKKO

Unknown

100

400

Both

Common

12

Offset

Both

2D

Dune advance into a

Buynevich,

equatorial Brazil: A

Pedro Walfir M.

subsurface

Souza Filho, Nils

perspective

E. Asp

Sedimentary records

2004

Ilya V.

of intense storms in

Buynevich,

Holocene barrier

Duncan M.

sequences, Maine,

FitzGerald, Sytze

USA

van Heteren

Using two- and three-

1999

Milan Beres,

dimensional georadar

Peter

methods to

Huggenberger,

characterize

Alan G. Green,

glaciofluvial

Heinrich

architecture

Horstmeyer

Combining
sedimentological and
high-resolution 3-D
mapping of fluvial
architectural elements
in the Quaternary Po
plain (Italy)

252

Ilya V.

coastal forest,

geophysical data for

Table A3 continued

2009

2007

R. Bersezio, M.
Giudici, M. Mele

Eolian

Holocene

GSSI SIR

200

3000

Common

2D

Offset

Internal structure of

2009

Inmaculada

the aeolian sand

Rodrguez

dunes of El Fangar

Santalla, Mara

spit, Ebro Delta

Jos Snchez

(Tarragona, Spain)

Garca, Isabel
Montoya
Montes, David
Gmez Ortiz,
Tomas Martn
Crespo, Jordi
Serra Raventos

Fluvial

Holocene

PulseEKKO

12.5

Both

36

2D

Characterizing large

Derald G. Smith,

shallow geophysics:

David T.

Middle Yukon River,

Clement

Alaska

253

Duane G. Froese,

river history with

Yukon Territory and

Table A3 continued

2005

Beach

Holocene

PulseEKKO

50

Common

15

2D

Offset

Dating Of Late

2010

Curt D. Peterson,

Holocene Beach

Harry M. Jol,

Shoreline Positions

Sandy

By Regional

Vanderburgh,

Correlation of

James B. Phipps,

Coseismic Retreat

David Percy,

Events In The

Guy Gelfenbaum

Columbia River
Littoral Cell, USA
Coastal

Holocene

Barrier

GSSI SIR

200

3000

Common

2D

Offset

The sedimentary

2008

Sebastian

architecture of a

Lindhorst,

Holocene barrier spit

Christian

(Sylt, German Bight):

Betzler, H.

Swash-bar accretion

Christian Hass

and storm erosion


Fluvial

Holocene

PulseEKKO

200

Both

Table A3 continued

2D

Bar-top hollows: A

2006

Jim Best, John

new element in the

Woodward, Phil

architecture of sandy

Ashworth, Greg

braided rivers

Sambrook Smith,
Chris Simpson

Table A3 continued
254

Fluvial

Holocene

PulseEKKO

100

Both

2D

Architecture of

2003

Raymond L.

channel-belt deposits

Skelly, Charlie

in an aggrading

S. Bristow,

shallow sandbed

Frank G.

braided river: the

Ethridge

lower Niobrara River,


northeast Nebraska
Eolian

Holocene

Unknown

Unknown

Common

25

Offset

2D

A Holocene history

2004

Walter L. Loope,

of dune-mediated

Timothy G.

landscape change

Fisher, Harry M.

along the

Jol, Ronald J.

southeastern shore of

Goble, John B.

Lake Superior

Anderton,
William L.
Blewett

Table A3 continued
255

Glacio-fluvial

Holocene

PulseEKKO

100

Both

12

2D

Sedimentary and

2004

Matthew R.

tectonic architecture

Bennett, David

of a large push

Huddart, Richard

moraine: a case study

I. Waller, Nigel

from Hagafellsjkull-

Cassidy,

Eystri, Iceland

Alexandre
Tomio, Paul
Zukowskyj,
Nicholas G.
Midgley, Simon
J. Cook, Silvia
Gonzalez, Neil
F. Glasser

Glacio-fluvial

Beach

Holocene

Holocene

PulseEKKO

PulseEKKO

100

200

Both

Both

15

2D

2D

Architecture and

256

Kurt H. Kjr,

sedimentation of

Lina Sultan,

outwash fans in front

Johannes Krger,

of the Mrdalsjkull

Anders

ice cap, Iceland

Schomacker

Geomorphic evidence

2009

Adam D.

for midlate

Switzer, Craig R.

Holocene higher sea

Sloss, Brian G.

level from

Jones, Charles S.

southeastern

Bristow

Australia

Table A3 continued

2004

Lake

Holocene

PulseEKKO

200

Common

10

2D

Offset

The influence of

2006

Bryan Shuman,

seasonal precipitation

Jeffrey P.

and temperature

Donnelly

regimes on lake
levels in the
northeastern United
States during the
Holocene
Glacio-fluvial

Coastal

Holocene

Holocene

Barrier

PulseEKKO

Radar
Systems Inc

100

300

Both

Common

20

Offset

- Zond 12c

2D

2D

The formation of

2009

Valentin Burki,

sawtooth moraine

Eiliv Larsen, Ola

ridges in Bdalen,

Fredin, Aninna

western Norway

Margreth

Solar-forced 2600 BP

2007

S.B.

and Little Ice Age

Kroonenberg,

highstands of the

G.M.

Caspian Sea

Abdurakhmanov,
E.N. Badyukova,
K. van der Borg,
A. Kalashnikov,
N.S. Kasimov,
G.I. Rychagov,
A.A. Svitoch,
H.B. Vonhof,
F.P. Wesselingh

Table A3 continued

257

Beach

Holocene

GSSI

120

Common

10

2D

Offset

Sand budgets at

2005

Joseph T. Kelley,

geological, historical

Donald C.

and contemporary

Barber, Daniel F.

time scales for a

Belknap, Duncan

developed beach

M. FitzGerald,

system, Saco Bay,

Sytze van

Maine, USA

Heteren, Stephen
M. Dickson

Eolian

Holocene

PulseEKKO

100

Common

12

2D

Offset

Formation of aeolian

2007

Lars B.

dunes on Anholt,

Clemmensen,

Denmark since AD

Mette Bjrnsen,

1560: A record of

Andrew Murray,

deforestation and

Karsten Pedersen

increased storminess
Coastal

Holocene

Barrier

Ramac

100

Both

12

Unshielded

2D

Lithological

258

Jonathan

heterogeneity in a

Hodgkinson,

back-barrier sand

Malcolm E. Cox,

island: Implications

Stephen

for modelling

McLoughlin,

hydrogeological

Gary J. Huftile

frameworks

Table A3 continued

2008

Fluvial

Holocene

PulseEKKO

50

Both

12

2D

Integrated

2007

Jerry C.

geophysical and

Bowling, Dennis

geological

L. Harry,

investigation of a

Antonio B.

heterogeneous fluvial

Rodriguez,

aquifer in Columbus

Chunmiao Zheng

Mississippi
Coastal

Holocene

Unknown

200

Common

2D

Offset

Evidence for late

2009

Hamid Alizadeh

Holocene highstands

Ketek Lahijani,

in Central Guilan

Hossain

East Mazanderan,

Rahimpour-

South Caspian coast,

Bonab, Vahid

Iran

Tavakoli, Muna
Hosseindoost

Eolian

Holocene

PulseEKKO

100

Both

10

2D

Origin of a complex

Gary Kocurek,

dune-field pattern,

Ryan C. Ewing,

Algodones,

Charlie Bristow

California

259

Dana Derickson,

and spatially diverse

southeastern

Table A3 continued

2008

Eolian

Holocene

GSSI

270

Shielded

Common

15

2D

Offset

An extended field of

2006

Philippe Paillou,

crater-shaped

Bruno Reynard,

structures in the Gilf

Jean-Marie

Kebir region, Egypt:

Malzieux, Jean

Observations and

Dejax, Essam

hypotheses about

Heggy, Pierre

their origin

Rochette, Wolf
Uwe Reimold,
Patrick Michel,
David Baratoux,
Philippe Razin,
Jean-Paul Colin

Beach

Holocene

PulseEKKO

50

Both

20

2D

Composition, age,

2010

Curt D. Peterson,

and depositional rates

Sandy

of shoreface deposits

Vanderburgh,

under barriers and

Michael C.

beach plains of the

Roberts, Harry

Columbia River

M. Jol, Jim

littoral cell, USA

Phipps, David C.
Twichell

Table A3 continued
260

Lake

Holocene

GSSI SIR

200

2000

Common

15

2D

Offset

Evidence of

2009

Paige E. Newby,

centennial-scale

Jeffrey P.

drought from

Donnelly, Bryan

southeastern

N. Shuman,

Massachusetts during

Dana

the

MacDonald

Pleistocene/Holocene
transition
Glacio-fluvial

Holocene

Unknown

80

Common

15

Offset

2D

Early Holocene
regressive spitplatform and
nearshore
sedimentation on a
glaciofluvial complex
during the Yoldia Sea
and the Ancylus Lake
phases of the Baltic
Basin, SW Finland

Table A3 continued

261

2003

Joni Mkinen,
Matti Rsnen

Coastal

Holocene

Unknown

100

Common

2D

Offset

Lagoa da Aplia: A

2009

H. Granja, F.

residual lagoon from

Rocha, M.

the Late Holocene

Matias, R.

(NW coastal zone of

Moura, F.

Portugal)

Caldas, J.
Marques, H.
Tareco

Coastal

Holocene

GSSI SIR-3

120

Common

25

Offset

2D

Influence of relative

VandePlassche

tidal-inlet

(1997)

barrier-spit
stratigraphy, Sandy
Neck, Massachusetts

262

Van Heteren and

sea-level change and


development on

Table A3 continued

1997

Glacio-fluvial

Holocene

PulseEKKO

200

Common

3D

Offset

Estimating 3D

2009

Troy R. Brosten;

variation in active-

John H.

layer thickness

Bradford; James

beneath arctic

P. McNamara;

streams using

Michael N.

ground-penetrating

Gooseff; Jay P.

radar

Zarnetske;
William B.
Bowden;
Morgan E.
Johnston

Fluvial

Holocene

PulseEKKO

200

Both

2D

The use and

2003

John Woodward,

application of GPR in

Philip J.

sandy fluvial

Ashworth, James

environments;

L. Best, Gregory

methodological

H. Sambrook

considerations

Smith, and
Christopher J.
Simpson

Table A3 continued
263

Glacio-fluvial

Holocene

GPR derived

2003

N. J. Cassidy, A.

architecture of

J. Russell, P. M.

November 1996

Marren, H. Fay,

jokulhlaup deposits,

O. Knudsen, E.

Skeidararsandur,

L. Rushmer, and

Iceland

T. A. G. P. van
Dijk

Flood deposit

Pleistocene

PulseEKKO

100

Common

Both

Offset

Eolian

Pleistocene

PulseEKKO

100

Common

GPR imaging of
3,600

2D

Offset

2007

William P.

clastic dikes at the

Clement and

Hanford Site,

Christopher J.

Hanford, Washington

Murray

GPR surveys of

2007

C. S. Bristow, B.

vegetated linear dune

G. Jones, G. C.

stratigraphy in central

Nanson, C.

Australia; evidence

Hollands, M.

for linear dune

Coleman, and D.

extension with

M. Price

vertical and lateral


accretion
Eolian

Pleistocene

Nogging
Plus

250

Common

Offset

2D

Ground-penetrating

264

C. H.

radar (GPR) imaging

Hugenholtz, B. J.

of the internal

Moorman, and S.

structure of an active

A. Wolfe

parabolic sand dune

Table A3 continued

2007

Glacio-fluvial

Pleistocene

GSSI SIR-

300

Both

10

3D

Three-dimensional

10A

2003

GPR analysis of

J. Heinz and T.
Aigner

various Quaternary
gravel-bed braided
river deposits
(southwestern
Germany)
Estuarine

Pleistocene

Nogging

250

Common

2D

GPR investigation of

Offset

2003

Michael L.

multiple stage-5 sea-

O'Neal and

level fluctuations on

Richard K. Dunn

a siliciclastic
estuarine shoreline,
Delaware Bay,
southern New Jersey,
USA
Coastal

Pleistocene

PulseEKKO

100

Both

18

3D

400

A GPR analysis of a

and Jandyr M.

stratigraphy at the

Travassos

Janeiro

265

Maria C. Pessoa

Quaternary
coast of Rio de

Table A3 continued

2007

Glacial

Pleistocene

Ramac

100

Both

2D

Unshielded

GPR images of

Matthias

periglacial slope

Leopold and

deposits beneath peat

Joerg Voelkel

bogs in the Central


European Highlands,
Germany
Coastal

Glacial

Pleistocene

Pleistocene

PulseEKKO

GSSI SIR-2

50

100

Both

Common

25

2D

2D

Offset

Glacial

Pleistocene

Common

2D

Offset

GPR stratigraphy

Christopher J.

transgressive

Simpson, Harry

deposition of spits

M. Jol, Richard

and a barrier, Lake

A. Meyers, and

Bonneville, Stockton,

Donald R.

Utah, USA

Currey

GPR-based detection

2007

Rolf Gerber,

of Pleistocene

Christina Salat,

periglacial slope

Andreas Junge,

deposits at a shallow-

and Peter Felix-

depth test site

Henningsen

Structure of a

2003

Marcel A. J.

Pleistocene push

Bakker and Jaap

moraine revealed by

J. M. van der

GPR; the eastern

Meer

Netherlands
266

Derald G. Smith,

used to infer

Veluwe Ridge, the

Table A3 continued

2003

Glacial

Pleistocene

PulseEKKO

50

Both

15

2D

Morphology and

2006

S. Sadura, I. P.

GPR stratigraphy of a

Martini, A. L.

frontal part of an end

Endres, and K.

moraine of the

Wolf

Laurentide ice sheet;


Paris Moraine near
Guelph, ON, Canada
Carbonate

Pleistocene

Unknown

100

Both

3D

1104

3D GPR reveals

2002

complex internal

Mark Grasmueck
and Ralp Weger

structure of
Pleistocene oolitic
sandbar
Carbonate

Pleistocene

Unknown

250

3D

350

Karst and Early

2007

Sean A. Guidry,

Fracture Networks in

Mark

Carbonates, Turks

Grasmueck,

and Caicos Islands,

Daniel G.

British West Indies

Carpenter,
Andrew M.
Gombos, Jr.,
Steven L.
Bachtel, and
David A.
Viggiano

Table A3 continued
267

Coastal

Pleistocene

Ramac

200

Both

2D

Unshielded

Significance of

2010

Z. Kanbur, M.

shallow seismic

Gormus, S.

reflection (SSR) and

Kanbur, Z.

ground penetrating

Durhan

radar (GPR) profiling


on the Modern Coast
line History of the
Bedre Area, Eirdir
Lake, Isparta, Turkey
Carbonate

Pleistocene

GSSI

100

Both

SIR10A

2D

Application of
ground-penetrating
radar, digital optical
borehole images, and
cores for
characterization of
porosity hydraulic
conductivity and
paleokarst in the
Biscayne aquifer,
southeastern Florida,
USA

Table A3 continued
268

2004

Kevin J.
Cunningham

Fluvial

Pleistocene

GSSI SIR-

200

20

Common

2D

Offset

Fluvial

2007

A.K. Patidar,

geomorphology and

D.M. Maurya,

neotectonic activity

M.G. Thakkar,

based on field and

L.S. Chamyal

GPR data, Katrol hill


range, Kachchh,
Western India
Volcaniclastic

Pleistocene

GPR Zond
12c

300

Common

Offset

2D

Correlation of near-

Freyre, Mariano

and physical

Cerca, Martn

properties of clayey

Hernndez-

sediments from

Marn

Mexico, using
Ground Penetrating
Radar

269

Dora Carren-

surface stratigraphy

Chalco Basin,

Table A3 continued

2003

Coastal

Pleistocene

PulseEKKO

50

Both

20

2D

Late Quaternary

2002

Michael L.

stratigraphy and sea-

ONeal, Susan

level history of the

McGeary

northern Delaware
Bay margin, southern
New Jersey, USA:: a
ground penetrating
radar analysis of
composite
Quaternary coastal
terraces
Coastal

Pleistocene

PulseEKKO

900

Both

2D

Sedimentology of

2003

Adrian Neal,

coarse-clastic beach-

Julie Richards,

ridge deposits, Essex,

Ken Pye

southeast England
Deltaic

Pleistocene

PulseEKKO

25

Both

57

2D

Ground penetrating
radar: antenna
frequencies and
maximum probable
depths of penetration
in Quaternary
sediments

Table A3 continued
270

1995

Derald G. Smith,
Harry M. Jol

Glacio-fluvial

Pleistocene

Oyo

250

Both

15

2D

Ground-probing radar

Georadar 1

1994

Peter

as a tool for

Huggenberger,

heterogeneity

Edi Meier,

estimation in gravel

Andr Pugin

deposits: advances in
data-processing and
facies analysis
Glacio-fluvial

Pleistocene

GSSI SIR

100

10A

Common

10

2D

Contribution of

Offset

2003

Jean-Christophe

geophysics to the

Gourry,

study of alluvial

Francoise

deposits: a case study

Vermeersch,

in the Val d'Avaray

Manuel Garcin,

area of the River

Denis Giot

Loire, France
Glacio-fluvial

Pleistocene

GSSI SIR
10A

300

Common

Offset

Both

308

Towards realistic
aquifer models:
three-dimensional
georadar surveys of
Quaternary gravel
deltas (Singen Basin,
SW Germany)

Table A3 continued

271

1999

U. Asprion, T.
Aigner

Glacio-fluvial

Pleistocene

Ramac

100

Unshielded

Common

20

2D

Offset

Sequence

2004

Stein Kjetil Helle

1995

Henrik Olsen,

stratigraphy in a
marine moraine at the
head of
Hardangerfjorden,
western Norway:
evidence for a highfrequency relative
sea-level cycle

Glacio-fluvial

Pleistocene

GSSI SIR

300

10

Common

10

2D

Offset

Sedimentology and
ground-penetrating

Frank Andreasen

radar characteristics
of a Pleistocene
sandur deposit
Coastal

Pleistocene

Barrier

GSSI SIR 3

120

Common

12

Offset

Table A3 continued

272

2D

Evidence for storm-

2004

Amy J.

dominated early

Dougherty,

progradation of

Duncan M.

Castle Neck barrier,

FitzGerald, Ilya

Massachusetts, USA

V. Buynevich

Alluvial Fan

Pleistocene

GSSI SIR

100

Both

10

2D

3-D architecture,

2010

J. Hornung, D.

2000 &

depositional patterns

Pflanz, A.

GSSI SIR

and climate triggered

Hechler, A.

3000

sediment fluxes of an

Beer, M.

alpine alluvial fan

Hinderer, M.

(Samedan,

Maisch, U. Bieg

Switzerland)
Glacio-fluvial

Deltaic

Pleistocene

Pleistocene

Unknown

GSSI SIR

200

100

Both

Both

12

2000

2D

2D

Controls on the

Burke, John

architecture of a

Woodward,

single event englacial

Andrew J.

esker:

Russell, P. Jay

Skeiarrjkull,

Fleisher, Palmer

Iceland

K. Bailey

3-D sedimentary

2005

Boris Kostic,

architecture of a

Andreas Becht,

Quaternary gravel

Thomas Aigner

Implications for
hydrostratigraphy

273

Matthew J.

sedimentary

delta (SW-Germany):

Table A3 continued

2008

Glacio-fluvial

Pleistocene

200

Both

10

2D

Aquifer

2002

Ugur Yaramanci,

characterisation using

Gerhard Lange,

Surface NMR jointly

Marian Hertrich

with other
geophysical
techniques at the
Nauen/Berlin test site
Glacio-fluvial

Pleistocene

Oyo

250

Both

10

2D

Georadar 1

Ground-probing radar

1994

Peter

as a tool for

Huggenberger,

heterogeneity

Edi Meier,

estimation in gravel

Andr Pugin

deposits: advances in
data-processing and
facies analysis
Deltaic

Pleistocene

Ramac
Unshielded

100

Common

Offset

2D

An integrated

L. D. Slater, R.

investigation of the

Versteeg

anisotropic
unconfined aquifer

274

S. K. Sandberg,

geophysical
hydrogeology of an

Table A3 continued

2002

Glacio-fluvial

Pleistocene

GSSI SIR-

300

Both

3D

10A

975

Morphology and

2001

A. J. Russell, .

sedimentology of a

Knudsen, H.

giant supraglacial,

Fay, P. M.

ice-walled,

Marren, J. Heinz,

jkulhlaup channel,

J. Tronicke

Skeiarrjkull,
Iceland: implications
for esker genesis
Alluvial Fan

Pleistocene

PulseEKKO

100

Common

2D

Offset

Relationship between

1999

J. O. Campos-

extensional tectonic

Enriquez, J.

style and the

Ortega-Ramrez,

paleoclimatic

D. Alatriste-

elements at Laguna

Vilchis, R. Cruz-

El Fresnal,

Gtica, E.

Chihuahua Desert,

Cabral-Cano

Mexico
Carbonate

Unknown

Ramac

400

Both

Unshielded

Table A3 continued
275

2D

Application of

2004

Jianghai Xia,

deterministic

Evan K.

deconvolution of

Franseen,

ground-penetrating

Richard D.

radar data in a study

Miller, Thomas

of carbonate strata

V. Weis

Gypsum

Unknown

GSSI

500

SIR10A

Common

2D

GPR and seismic

Offset

2000

imaging in a gypsum

Xavier Drobert,
Odile Abraham

quarry
Alluvial Fan

Unknown

Unknown

25

Both

40

2D

Assessing fault

2009

M. Christie, G.P.

displacement and off-

Tsoflias, D.F.

fault deformation in

Stockli, R. Black

an extensional
tectonic setting using
3-D groundpenetrating radar
imaging
Carbonate

Unknown

PulseEKKO

50

Common

2D

Ground penetrating

Offset

2000

S. E. Kruse, J. C.

radar imaging of cap

Schneider, D. J.

rock, caliche and

Campagna, J. A.

carbonate strata

Inman, T. D.
Hickey

Fluvial

Unknown

Unknown

200

Common

Offset

Table A3 continued
276

3D

1670

Results of 3-D

2003

Alan Green, Ralf

georadar surveying

Gross, Klaus

and trenching the San

Holliger,

Andreas fault near its

Heinrich

northern landward

Horstmeyer,

limit

John Baldwin

Carbonate

Unknown

200

Both

3D

110

GPR imaging of

2008

dual-porosity rocks:

Georgios P.
Tsoflias

Insights to fluid flow


Architecture and

2004

Jeffrey E.

evolution of a fjord-

Gutsell, John J.

head delta, western

Clague, Melvyn

Vancouver Island,

E. Best, Peter T.

British Columbia

Bobrowsky, Ian
Hutchinson

Fluvial

PulseEKKO

100

Both

2D

Radar signatures and

277

Michael C.

structure of an

Roberts, Jean-

avulsed channel:

Paul Bravard,

Rhne River, Aoste,

Harry M. Jol

France

Table A3 continued

1997

Number of published GPR studies

Number of published GPR studies

Figure A7: (A) Analyses of 150 published GPR


studies on sediments showing bias for Quaternary
sediments (B) Analysis of studies in (A) showing
comparatively more GPR studies on carbonate
rocks than clastics

278

Appendix 4

GPR Data Processing Software and Display


Various GPR processing are now commercially available; although many GPR users still use
seismic data processing software (such as VISTA, ProMax 3-D) for GPR data processing after
conversion of GPR data to SEGY (seismic software compatible) format.
The major GPR product manufacturers (Sensors and Software, GSSI, Mala Geoscience) have
developed software for processing data obtained with their products although many of these
applications do not accept data format generated by GPR equipment from other manufacturers
(except GPR data converted to SEGY format). Data acquired with Sensors and Software systems
are usually processed with the PulsEkko or Noggin processing system including data view, data
processing and 3D visualization modules. GSSI GPR equipment are typicallly processed with
RADAN software while GPR equipment from Mala Geoscience are processed with
RadExplorer. Few other commercially available software like Reflex W from Sandmeier
Scientific Software and GPR-Slice software can be used to process data from most GPR
equipment manufacturers without the need for data conversion. Many GPR users do export
processed GPR data into 3D data visualization software such as Gocad, petrel, Slicer Dicer, Easy
3D, Stratimagic 3-D due to their advanced visualization capability.
For this study, ReflexW was used for GPR processing as it has robust data processing features
and GPR-Slice software used for 3D data visualization and interpretation.

279

Appendix 5
GPR profiles from study locations

Low

Depth (m)
High

6
Amplitude Scale

B
St

SC

Sp

3
Sp

St

SB

2
1

SB

Sp

Queenston Shale
3m

6th Order surface


5th order surface

5m

3rd order surfcas


Figure A8: (A) GPR line AB in grayscale (B) Bounding surfaces and architectural elements in radar line AB (SC Scour; SB Sandy Bedform ;St Trough crossbedded sandstone ; Sp Planar coss-bedded sandstone)

280

Interdune freshwater carbonates

Interdune wavy beds


Line 136

Wet interdune carbonate

Dry phase dune


Duine flank damp phase deposit
Line 137

Wet interdune carbonate

Dune flank damp phase deposit


Line 138

Figure A8: (A) GPR line 136, 137 and 138 from Big Mesa (Utah) showing major radar facies see GPR line location in Figure 40

281

Вам также может понравиться