Sei sulla pagina 1di 16

Journal of Structural Geology 69 (2014) 163e178

Contents lists available at ScienceDirect

Journal of Structural Geology


journal homepage: www.elsevier.com/locate/jsg

Review article

Ground-based and UAV-Based photogrammetry: A multi-scale, high-


resolution mapping tool for structural geology and paleoseismology
Sean P. Bemis a, *, Steven Micklethwaite b, Darren Turner c, Mike R. James d, Sinan Akciz e,
Sam T. Thiele b, Hasnain Ali Bangash b
a
Department of Earth and Environmental Sciences, University of Kentucky, 101 Slone Research Building, Lexington, KY 40506, USA
b
CET (M006), School of Earth and Environment, The University of Western Australia, 35 Stirling Highway, Crawley, WA 6009, Australia
c
School of Land and Food, University of Tasmania, Hobart, Tasmania 7001, Australia
d
Lancaster Environment Centre, Lancaster University, Lancaster LA1 4YQ, UK
e
Department of Earth, Planetary, and Space Sciences, University of California, Los Angeles, Los Angeles, CA 90095, USA

a r t i c l e i n f o a b s t r a c t

Article history: This contribution reviews the use of modern 3D photo-based surface reconstruction techniques for high
Received 21 May 2014 fidelity surveys of trenches, rock exposures and hand specimens to highlight their potential for paleo-
Received in revised form seismology and structural geology. We outline the general approach to data acquisition and processing
13 October 2014
using ground-based photographs acquired from standard DSLR cameras, and illustrate the use of similar
Accepted 14 October 2014
Available online 27 October 2014
processing approaches on imagery from Unmanned Aerial Vehicles (UAVs). It is shown that digital map
and trench data can be acquired at ultra-high resolution and in much shorter time intervals than would
be normally achievable through conventional grid mapping. The resulting point clouds and textured
Keywords:
Photogrammetry
models are inherently multidimensional (x, y, z, point orientation, colour, texture), archival and easily
Structural geology transformed into orthorectified photomosaics or digital elevation models (DEMs). We provide some
Neotectonics examples for the use of such techniques in structural geology and paleoseismology while pointing the
3D surface modelling interested reader to free and commercial software packages for data processing, visualization and 3D
UAVs interpretation. Photogrammetric models serve to act as an ideal electronic repository for critical outcrops
Structure-from-Motion and observations, similar to the electronic lab book approach employed in the biosciences. This paper
also highlights future possibilities for rapid semi-automatic to automatic interpretation of the data and
advances in technology.
© 2014 Elsevier Ltd. All rights reserved.

1. Introduction photogrammetric techniques; most notably the ability to visualize


the Earth's surface and extract topographic data from stereo aerial
High-resolution three dimensional (3D) data capture is required photographs (e.g., Birdseye, 1940; Eardley, 1942). With the funda-
at all scales in the geosciences, from hand specimen to landscapes, mental principles of photogrammetry now combined with robust
and a range of tools are available for addressing different portions algorithms from the computer vision community, collections of
of the scale spectrum (e.g., McCaffrey et al., 2005). In particular, overlapping photographs can be automatically processed to rapidly
recent advances in high-resolution digital 3D data collection are extract the relative 3D coordinates of millions of surface points
dominated by active source sensors, predominantly based upon (Lowe, 2004; Snavely et al., 2008a, 2008b, 2006). Therefore, the
laser scanning technologies (e.g., LiDAR), which measure distance only specialized resource required for acquisition of 3D data
to a target based upon the travel time of reflected light (e.g., through photogrammetric techniques is access to suitable software
Hodgetts, 2013). However, a new development in high-resolution which depending on computer skills and requirements, is available
3D data collection exploits a very common and widely accessible through both commercial and open-source options (Table 1).
passive imaging source e digital photography. Geoscientists have The limited infrastructure requirements of modern photo-
long utilized the 3D information available through grammetric techniques present a wide range of opportunities for
geoscience research and education. In the most basic form, the raw
data consist of only digital photographs, which can be collected
with any commonly available digital camera, including those on
* Corresponding author.
E-mail address: sean.bemis@uky.edu (S.P. Bemis).
smartphones and tablets. As such, the technique facilitates rapid

http://dx.doi.org/10.1016/j.jsg.2014.10.007
0191-8141/© 2014 Elsevier Ltd. All rights reserved.
164 S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178

Table 1
Examples of open source and commercial software for photo-based 3d reconstruction.

Software Url (valid on 17 May, 2014) Notes

Freely available
Bundler Photogrammetry http://blog.neonascent.net/archives/bundler- Used in James and Robson (2012). Script-based, no graphical user interface
Packagea,b photogrammetry-package/ (GUI). Windows OS only.
SFMToolkita,b http://www.visual-experiments.com/demos/sfmtoolkit/ Similar software to above.
Python Photogrammetry http://code.google.com/p/osm-bundler/ Formerly OSM-bundler. Python-driven GUI and scripts, with a Linux
Toolbox (PPT)a,b distribution.
VisualSFMb http://www.cs.washington.edu/homes/ccwu/vsfm/ Advanced GUI with Windows, Linux and Mac. OSX versions. Georeferencing
options, but camera model is more restricted than that used in Bundler.
3DF Samantha http://www.3dflow.net/technology/samantha-structure- SfM only, but with more advanced camera models than all above (Farenzena
from-motion/ et al., 2009). Provides output compatible with several dense matching
algorithms.
Web sites and services
Photosynth http://photosynth.net/ Evolved from Bundler. SfM only, no dense reconstruction. Can incorporate a
very wide variety of images, but does so at the cost of reconstruction
accuracy.
Arc3D http://www.arc3d.be/ Vergauwen and Van Gool [2006]
CMP SfM Web servicea http://ptak.felk.cvut.cz/sfmservice/
Autodesk 123D Catch http://www.123dapp.com/catch/
Pix4D http://pix4d.com/ Also available as standalone software.
My3DScanner http://www.my3dscanner.com/
Commercial
PhotoScan http://www.agisoft.ru/products/photoscan/ Full SfM-MVS-based commercial package.
Acute3D http://www.acute3d.com/
PhotoModeler http://www.photomodeler.com/ Software, originally based on close-range photogrammetry, now also
implements SfM.
3DF Zephyr Pro http://www.3dflow.net/ Underlying SfM engine is 3DF Samantha

Note: Table modified from http://www.lancaster.ac.uk/staff/jamesm/research/sfm.htm.


SfM ¼ Structure from Motion; MVS ¼ Multi-View Stereo.
a
Uses Bundler (http://phototour.cs.washington.edu/bundler/) to compute structure from motion.
b
Uses PMVS2 (http://grail.cs.washington.edu/software/pmvs/) as a dense multi-view matcher.

collection of large amounts of data in remote settings where studies. Furthermore, photogrammetry has been directly employed
portability and efficiency may be critical. Because the source data for many years in the geosciences through stereoscopic viewing
are photographs, the derivative 3D data can readily be coloured or and analysis of overlapping pairs of aerial photographs (e.g.,
draped with the source photography to produce realistic 3D models Eardley, 1942; Pillmore, 1964). This stereoscopic viewing provides
of the feature of interest, which can then be exported to 3D visu- the researcher with the ability to visualize and map a study area
alization environments for analysis. In addition, with free viewers remotely and from a perspective that is impossible to attain in the
for many 3D formats (including the ubiquitous Adobe Reader for 3D field. These photogrammetric techniques were adapted for working
PDFs), models can be widely shared for exploration and enhanced at the outcrop scale for simple visualization purposes by taking
understanding of the 3D nature of many geoscience examples. 3D photo pairs of the outcrop of interest for later viewing through a
photogrammetric models and their accompanying digital photo- mirror stereoscope (Kuenen, 1950). This approach is valuable for a
graphs are inherently archival, easily shared and provide a record more representative visualization of a site once the researcher
that is faithful to the primary observations, thereby allowing future returns to the office, but requires tedious procedures to enable
generations of geoscientists to extract additional 3D and oriented extraction of data that are fully 3D and oriented (Hagan, 1980).
data, or reinterpret the outcrop/samples. With the advent of modern digital camera technology, re-
The purpose of this review is to introduce these easy to use strictions around the number of photos that can be collected have
photogrammetric techniques to the general structural geology and been relaxed and picture quality can be quickly and easily assessed
paleoseismology communities while illustrating a range of appli- in the field. Now, the greater limitation lies in achieving optimal
cations. A variety of 3D data collection tools are now available to the camera positioning relative to the object of interest, whether due to
geoscientist, including a range of laser scanning options and vegetation, topography, objective hazards, etc. However, even this
traditional surveying methods, and each method has different ad- limitation is being reduced through the use of digital photography
vantages/disadvantages in terms of resolution, scalability, porta- from balloons, kites, and UAVs, enabling drastically improved
bility, and computer processing requirements. Photo-based 3D synoptic views from overhead (e.g., Smith et al., 2009; Niethammer
reconstruction techniques can provide a comparable resolution to et al., 2012; Stumpf et al., 2013).
laser scanning tools with a significant reduction in cost, infra-
structure, and processing requirements. To assist with adoption of
such techniques, we provide an overview of the technical basis 3. Principles and capability of photo-based 3D reconstruction
along with workflows and discussion of best practices for photo-
graph collection that will provide optimal results. 3.1. Basic principles

New methods of photo-based 3D reconstruction reflect the


2. Photogrammetry in the geosciences evolution of photogrammetry from a highly specialized technique,
requiring expensive software and restrictive image collection re-
For most of the 20th century, photogrammetry principles quirements, to a user-friendly and scalable methodology. Funda-
implemented through stereoscopic instruments were the primary mentally, photogrammetry works on the basis that, from two
means of the construction of topographic maps (e.g., Birdseye, overlapping photographs, it is possible to calculate the unique
1940) e a critical element of traditional field-based geologic three-dimensional (3D) location of a set of given points shared in
S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178 165

both photographs, relative to the cameras (Fig. 1). The unknowns integrated georeferencing and error analysis. The rapidly widening
comprise a ‘camera model’ that describes how the camera repre- availability of such software is the motivation for this paper. On that
sents the 3D world as a 2D image, the relative camera positions and basis, a typical photo-based reconstruction workflow is described
pointing directions, and the 3D point coordinates. In ‘conventional’ below and summarized in Fig. 2.
photogrammetry, which has evolved through the surveying, engi-
neering and remote sensing communities, initial estimates are
3.2. Workflow for high precision data collection
generally derived by providing additional control data, such as the
positions of known control points within images, prior to pro-
3.2.1. Image acquisition
cessing. This approach allows error estimates (e.g., the accuracy of
Images can be collected with almost any digital camera, and
the control measurements) and a real-world coordinate system to
there are few limitations in number or types of digital cameras used
be embedded from the outset. Processing then consists of refining
for a single set. As expected, a higher quality model output is
parameters through a ‘bundle adjustment’ (Granshaw, 1980), which
facilitated by higher quality input images. Additional data collec-
simultaneously optimizes all variables to produce a self-consistent
tion flexibility derives from the ability of SfM algorithms to match
3D model, with minimized overall residual error. Such software
images taken at varying scales and perspectives, provided the
enables a highly rigorous approach and provides accurate results in
photos still produce sufficient overlap. Stereoscopic aerial
which error estimates are widely visible. However, the software
often has complexities and error intolerance that can present dif-
ficulties for inexperienced users, and the requirements for collec-
Establish control

Fieldwork
tion of suitable imagery and control data can be arduous.
Over the last decade, parallel advances in an area of computer points or scale
vision research called ‘structure from motion’ (SfM) has been
driven by a different rationale e to enable automated model pro-
duction from unconstrained imagery, for which metric accuracy is
not the primary goal. Some of the most significant advances in SfM Collect photographs
have arisen from the development of image feature descriptors that
are tolerant to changes in view point (e.g., Lowe, 2004), and robust
matching algorithms that can identify and reject errors when they
occur (e.g., Fischler and Bolles, 1981). With these, bundle adjust-
ment can be initialized from automated image measurements
Mask non-stationary
alone. For example, to work efficiently, coarse estimates for camera portions of images
model parameters are also required, e.g., focal length, but most
software will just automatically extract appropriate values from
image file metadata. Thus, 3D models can be effectively constructed Feature detection,
from a wide variety of imagery, with no user intervention (e.g., bundle adjustment,
SfM

Snavely et al., 2008a, 2008b, 2006). However, the results of such a and 3D scene Sparse
3D reconstruction will be in an arbitrary coordinate system so, to point cloud
reconstruction
reference to a real-world system, the model needs to be trans-
formed through the use of some control data. The control re-
quirements are usually significantly less arduous than for
Dense
MVS

‘conventional’ photogrammetry, but control data are not neces- Pixel grid based
sarily incorporated throughout model construction and their error matching point cloud
estimates are more weakly integrated.
The ongoing convergence of photogrammetry- and computer
vision-based workflows is now providing powerful tools for geo- Build mesh/
science use (Favalli et al., 2012), enabling automated model pro- interpolate surface
duction from flexible image input and with increasing access to

Georeferencing and DEM


camera scale generation
positions era
cam f view
o
field
target surface

Texture mapping Reprojection

photo-
realistic model orthomosaic

Fig. 1. Basic principles of photogrammetry for 3D reconstruction. Two overlapping Fig. 2. General workflow illustrating the photo-based 3D reconstruction process.
photographs taken from different positions allow each feature in the overlapping area Diamond fields indicate potential output products at different stages during recon-
to be defined by a unique 3D position. Dashed lines illustrate convergent views of struction. MVS (Multi-view Stereo) and SfM (Structure-from-Motion) illustrate the
discrete features from overlapping photographs. portions of the workflow that specifically related to these processes described in text.
166 S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178

photography is typically collected with 50e60% overlap between corresponding features in overlapping photos. For good results, any
adjacent images, under near-parallel viewing conditions (e.g., effects that reduce textural variability within images or increase
Krauss, 1993; Abdullah et al., 2013). SfM-based 3D reconstruction feature variability between images should be minimized during
methods also require overlapping images but, because they can acquisition. Common issues preventing algorithms from resolving
operate on unordered collections of photographs, the overlap coincident points include homogeneous surface texture, changes in
requirement is best considered in terms of coverage and angular the target, and changes in illumination. The latter two have the
change between overlapping images. In terms of coverage, every same effect of making a unique feature appear differently between
surface that will be reconstructed needs to be covered by at least 2 images, although both require different strategies for reducing the
images taken from different positions, and preferably more (Fig. 3). deleterious effect on model construction. Poor or variable image
Increasing angles of convergence between overlapping images will texture is often due to surface reflections, flat surfaces with little
tend to increase reconstruction accuracy up to a point, but will textural variation, and the occurrence of deep shadows. The target
eventually prevent matching due to the surface texture appearing itself may appear to change between images due to wind shifting
too dissimilar in images from different directions. Moreels and vegetation, or the movement of people and vehicles. Changes in
Perona (2007) found that popular feature detectors used for auto- illumination can result from accidental shading by the photogra-
mated image matching did not perform well with angular changes pher, changes in the sun position, or filtering by clouds. Strategies
greater than 25e30 between images. Thus, while angular changes for circumnavigating these issues in structural studies are outlined
between photos can increase the accuracy of reconstructed 3D further in Section 5.2.
surfaces, differences should be limited to 10e20 for overlapping
photos.
The simplest approach for image acquisition over relatively 3.2.2. Scale and coordinates
planar surfaces mimics the approach of collecting traditional ste- The scaling and georeferencing requirements for the target
reoscopic aerial photographs where images are collected in a surface will vary with the intended use of the 3D data/model and
continuous line with a camera position orthogonal to the surface of should be considered prior to image acquisition. An object or sur-
interest and with a frequency to produce image overlap >60% face can be fully reconstructed in 3D without any scale or position
(Fig. 3). However, the scale and layout of the target surface will information but, to extract oriented and scaled data, additional
frequently necessitate adjustments to this simple approach. control data must be provided. Scale can be added to a model
Furthermore, James and Robson (2014) document systematic errors simply by knowing the distance between two points on an input
that can be introduced across models derived from image collec- image or on the model. These distance measurements for scale are
tions where all photos are collected with parallel viewing di- best taken over the width of the target area rather than smaller,
rections. This error is manifest in the axis parallel to the image view isolated lengths. Greater accuracy in scale and full georeferencing
direction, for example producing broad-scale elevation error from to local or geographic coordinates requires three or more ground
images collected with a vertical orientation. James and Robson control points, usually collected with survey-grade, carrier phase
(2014) demonstrate that one approach to mitigate this systematic differential GPS or total station surveying. These ground control
error by combining additional images with a view direction that is points should be distributed widely across the target area, not
inclined relative to the view direction of the rest of the image neglecting the margins. An alternate option for georeferencing is
collection (Fig. 3). the use of high-precision camera locations for the input images, but
In addition to image coverage, the other key consideration in the current GPS units in hand-held cameras are insufficient for the
planning of any survey is how the texture of the target will resolve typical accuracy requirements of high resolution structural map-
in individual photographs. Automated feature matching relies upon ping and paleoseismology projects conducted over a few hundred
the ability of computer algorithms to identify unique meters or less.

portion of surface resolvable by photo-based reconstructions

Fig. 3. Representation of simple image acquisition for photo-based 3D reconstruction. Grey triangles represent the field of view for each camera position and the increasing
darkness of the triangles corresponds with the number of camera positions that the surface is visible from. A greater number of overlapping images is likely to produce a denser
point cloud because more features should be resolvable, leading to a higher resolution model for that portion of the reconstruction. The sensitivity of image overlap to the image
separation, inclination, and the distance to the surface is illustrated by the size of the areas visible from 3 or more camera positions (darkest grey). Although the inclined camera
positions are not required, they are recommended for reducing possible systematic errors due to camera calibration error (James and Robson, 2014).
S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178 167

3.2.3. Determining the imaging geometry: structure from motion measurement precision, a stereo image pair with only two near-
The first stage of a SfM-based 3D reconstruction involves the parallel images would represent a weaker network than a conver-
analysis of the individual images for distinct image features that gent multi-image arrangement. For projects using digital SLR im-
can be matched to their corresponding features in other images agery covering sub-meter to kilometer scales processed with an
within the collection (e.g., Fig. 1). The SfM process then uses the SfM-based approach, James and Robson (2012) estimated relative
resulting network of matched points to establish the relative lo- precision ratios of ~1:1000 or greater. These ratios were shown to
cations of each camera and simultaneously determines the camera be similar to those of theoretical estimates for stereo photogram-
parameters for each image, the collection position and orientation metry, but approximately an order of magnitude poorer than
for each input image, and the 3D coordinates for each matched equivalent theoretical estimates for close-range (convergent)
feature e commonly referred to as the bundle adjustment. Algo- networks.
rithms typically adopt an incremental approach where bundle In most structural geology and neotectonic applications, we are
adjustment of an initial image pair is sequentially repeated, with interested in a bare surface model. In areas of light to dense
more images incorporated at each iteration. The matched features vegetation, this preference for bare surfaces becomes the primary
thus constitute a sparse 3D point cloud that represents the struc- disadvantage to photo-based 3D reconstruction relative to active
ture of the target surface, defined within a local coordinate system. source 3D data collection tools. LiDAR collects a 3D point location
This point cloud is limited in precision and point density because it with a single pulse of light, thus is capable of collecting ground
is derived from robust feature matching, which has lesser accuracy surface points wherever the pulse of light is able to penetrate the
(e.g., ~0.5 pixel) than some other matching approaches vegetation, reflect off the ground surface, and return to the in-
(Remondino, 2006; Barazzetti et al., 2010). strument. The requirement in photo-based 3D reconstructions for
multiple images collected from different perspectives for 3D point
3.2.4. Densifying the measurements: multi-view stereo geometry reconstruction dramatically reduces the number of
With known camera models and orientations, a multi-view resolvable ground surface points due to the occlusion of the ground
stereo algorithm will produce a dense point cloud representation surface by vegetation when moving from one camera position to
of the surface. Typically, this technique will be implemented as a the next. However, because the SfM process creates a point cloud,
systematic search over a pixel grid to identify best matches be- as long as the vegetation is sparse enough to allow a sufficient
tween images, with the results providing significantly more 3D number of ground features to be identified and matched, some of
points, having greater precision than the feature matching of the the approaches developed for classification of LiDAR data can be
initial SfM step. Multi-view stereo is particularly intensive applied to SfM-derived point clouds. Furthermore, one of the
computationally if the full image collection is processed simulta- popular commercial photo-based 3D reconstruction packages
neously. However, most photo-based 3D reconstruction programs (Table 1), Agisoft PhotoScan, has implemented a point cloud clas-
and algorithms have the option to subset image collections (e.g., sification scheme for classification of ground surface points and
Furukawa et al., 2010) or to adjust the grid-cell size at which multi- vegetation (http://www.agisoft.ru/tutorials/photoscan/08).
view stereo is performed so as to manage the resolution and time
required to produce the resultant dense point cloud. 4. Applications to paleoseismology and neotectonics

3.2.5. 3D model and orthophoto generation Major advances in the ability to collect, process, and visualize
For many applications within structural geology and paleo- high resolution 3D topographic data in the form of LiDAR has
seismology, 3D models and orthophotos are useful for high reso- revolutionized paleoseismology and neotectonics research over the
lution mapping of outcrop, rock faces or trenches. Using past ~15 years (e.g., Haugerud et al., 2003; Hudnut et al., 2002). The
triangulation or grid interpolation, it is relatively straightforward to ability to visualize bare-earth topography on a regional scale with
generate a digital elevation model (DEM) and ortho-rectified pho- sub-meter resolution in a wide variety of landscapes using airborne
tomosaics for any selected orientation from the dense, georefer- laser scanning (ALS) is currently unparallelled. ALS has been
enced point cloud. True 2D orthophotos of any portion of the 3D implemented in neotectonic studies for fault mapping (e.g.,
model can then be derived, with the known 3D model and imaging Arrowsmith and Zielke, 2009; Bevis et al., 2005; Oskin et al., 2007),
geometry, allowing correction for the viewing characteristics of the measurement of geomorphic offsets (e.g., Zielke et al., 2010), and
input images and the images providing the texture for the ortho- measurement of coseismic surface displacements (e.g., Borsa and
photo mosaic. Furthermore, depending on the software used, the Minster, 2012; Duffy et al., 2013; Nissen et al., 2012; Oskin et al.,
3D model itself can be textured and exported in common 3D 2012). Terrestrial laser scanning (TLS) provides a higher resolu-
visualization formats (e.g.,.obj,.ply,.pdf, etc.) for visualization and tion data collection (sub-decimeter), but usually over more
interactivity. restricted distances due to the limited range of most instruments
and the reduced visibility from working near ground-level. This
3.3. Precision and applicability technique has been similarly used to measure geomorphic offsets,
determine coseismic surface displacements (Gold et al., 2012), and
In general terms, the accuracy of a photo-based model depends scan paleoseismic excavations (Haddad et al., 2012). The high-
upon the scale and resolution of the input images, the distribution resolution capability of photo-based 3D reconstruction accommo-
and accuracy of control data (whether ground control points, scale dates many of the same applications as ALS and TLS surveying, but
measurements or camera positions), the precision and distribution also presents several site-specific and technical advantages in its
of matched image points, and the network geometry, which in- implementation due to portability, low power consumption, rapid
cludes the number of photos, how much they overlap and how data collection, and low cost (Morelan et al., 2010; Johnson et al.,
convergent the views are. In close-range photogrammetry (e.g., as 2014). We provide some examples illustrating these advantages.
often used for high accuracy engineering applications), where im-
age networks are usually multi-image and highly convergent, the 4.1. Neotectonics applications
strength of a network can be described by its relative network
precision, which is a ratio of the mean 3D point uncertainty esti- One of the critical roles for geoscientists following a major
mate to the longest dimension of the network. For a given image surface-rupturing earthquake is the systematic measurement of
168 S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178

offset features along the length of the surface rupture. Even though capture the offset feature from all directions (360 with 20 rota-
these measurements may underestimate total coseismic displace- tion between individual photographs) and additional photographs
ment due to distributed deformation off the main fault trace (Dolan from higher and lower perspectives will reduce shadowing and
and Haravitch, 2014), these offset measurements provide critical improve the model geometry. Fig. 5 shows an example of a photo-
information about the earthquake rupture and provide markers for based 3D reconstruction for an ephemeral stream channel offset
assessment of post-seismic processes such as afterslip, which is the 6.6 m during the 1857 Fort Tejon earthquake on the San Andreas
increasing displacement that occurs after the primary earthquake. fault (a fully interactive 3D model of this offset is provided in the
Although many landforms that are used to record slip over multiple Supplemental Materials). The figure illustrates both the digital
earthquake cycles are persistent in the landscape, most targets that elevation model from which morphological and displacement data
would uniquely record the offset during a single earthquake are can be extracted, and the fully textured model that preserves a
transient features, including the fault scarp itself, broken and offset natural visual depiction of the site. With simple scaling re-
trees, channel margins in active floodplains and, in the case of the quirements and rapid photograph collection, the additional work
2002 M7.9 Denali fault earthquake sequence, offset glacial cre- required for photo-based 3D reconstructions of individual coseis-
vasses (Fig. 4; Haeussler et al., 2004). Some of these features could mic offset measurements is nominal and will add to the robustness
be adequately documented with a post-earthquake ALS survey, but of the overall coseismic slip distribution dataset. Furthermore, as
others require site-by-site documentation because of the scale of observed following the 2014 M6.0 Napa, California, earthquake,
the offset. Traditionally, field measurements are collected with afterslip processes contributed to increasing surface displacements
standard geologic field equipment e tape measures and compasses, during the days following the earthquake (Brooks, 2014). The ease
but we propose that each coseismic offset be photographed for 3D of collecting data for photo-based 3D reconstruction of individual
reconstruction so as to archive the offset feature and to facilitate offsets could accommodate high spatial and temporal resolution
further analysis, such as automated slip-vector calculation (Gold monitoring of afterslip following future earthquakes.
et al., 2012). Because coseismic offsets are a relative displacement In regions of sparse vegetation, photo-based 3D reconstructions
measurement, georeferencing is not required but scale is critical. from aerial platforms can produce topographic surface models with
Scale is provided through introducing scale bars into the scene or comparable accuracy to ALS surveys (e.g., Fonstad et al., 2013;
physically measuring and recording the distance between features James and Robson, 2012; Johnson et al., 2014; Westoby et al.,
within the scene. The number of photographs required depends 2012). Although spatial coverage of ALS-derived topographic data
upon the nature of the offset feature and the magnitude of the is expanding, the availability is concentrated in relatively few
offset, although a minimum of 18 photos should be considered to countries and regions. Therefore, in regions where ALS-derived

Fig. 4. Examples of extremely short-lived geomorphic markers of coseismic fault displacement from the November 3rd, 2002, M7.9 Denali fault earthquake sequence in on the
Denali fault in south-central Alaska. Tremendous effort was invested in capturing these offsets, many of which disappeared by the following summer. Image collection for photo-
based 3D reconstruction techniques would not require more than a few extra minutes per site and preserve a richer record of the fault offset. (a) Patty Burns measures the vertical
separation of a fault scarp that is expressed as an offset glacier surface across the Susitna Glacier fault. (b) A 5.5 m offset of an active stream bank. (c) Measuring the displacement of
a tree that was broken and offset across the Denali fault. (d) Peter Haeussler attempts to measure the displacement of a glacial crevasse. The 5.5 m offset documented in (b)
immediately after the earthquake in November 2002 increased to 6.6 m by July 2003 (Haeussler et al., 2004). Photos (a), (b), and (d) are courtesy of the U.S. Geological Survey
(http://earthquake.usgs.gov/earthquakes/eqinthenews/2002/uslbbl/photos/pr071102/).
S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178 169

Fig. 5. Offset channel on the Carrizo Plain, California, from the 1857 Fort Tejon earthquake on the San Andreas fault. Photos were collected with a GoPro camera mounted on a long
pole and held overhead. This model is derived from 56 photos with (a) showing an oblique view of the resulting shaded relief surface model and (b) showing the identical oblique
view with the mosaicked image texture. This is channel offset ZA10543 from Zielke et al. (2010) who documented an offset of 6.6 m (þ0.5/1.0). Black arrows illustrate the offset of
the channel thalweg and point in the downstream direction. View is looking north. The 3D model these images are derived from is provided as a 3D PDF in the Supplementary
Material.

topographic data is not available, photo-based 3D reconstruction these photos are often taken at close range with wide angle camera
techniques can accommodate many of the innovative neotectonic lenses, surface irregularities on the margins of the photo are
analyses being performed with ALS-derived topography (e.g., Akçiz particularly problematic, potentially introducing positional errors
et al., 2010; Hilley et al., 2010; Nissen et al., 2012). Even where ALS- of several cm or more during manual rectification and obscuring
or TLS-derived topographic data already exist, photo-based 3D stratigraphic relationships. The resulting photomosaic then re-
reconstruction techniques can add to the temporal dimension of quires adjustment of tonal and lightness characteristics for
the topographic data by enabling more frequent surveys when adjoining photos to provide even colour and contrast across the full
repeat ALS or TLS surveys would be cost-prohibitive or logistically photomosaic.
difficult. Analyses of pre- and post-earthquake ALS topographic This complex and time-consuming approach to constructing 2D
data for synthetic earthquakes (Borsa and Minster, 2012) and actual photomosaics for paleoseismology can be largely circumvented
earthquake ruptures (Duffy et al., 2013; Nissen et al., 2012; Oskin through photo-based 3D reconstruction techniques. A properly
et al., 2012) demonstrate the ability to extract 3D displacements planned collection of photographs from a trench wall will facilitate
from high-resolution topography, and Krishnan et al. (2012) pre- the full 3D reconstruction of the topography of that surface, which
sent methods for registration of photo-based reconstruction sur- in turn accommodates the precise and automated orthor-
face models with ALS topography to enable similar analyses with ectification of the input photographs. For a planar trench wall, we
data derived from different methods. have achieved consistently high-quality results by taking photo-
graphs in a similar fashion as traditional photomosaics, but
4.2. Photo-based 3D reconstruction in paleoseismic investigations increasing photograph overlap so that ~3 times more photos are
collected (Fig. 6). In this case, rather than just photographing each
A fundamental product of any paleoseismic investigation is a measured grid cell, the image coverage is increased by collecting an
representation of the exposed stratigraphy and faulting e essen- intermediate photograph centered on the gridlines between each
tially producing a geologic map (the so-called “trench log”) of the grid cell e horizontally and vertically. This approach includes tak-
exposure. Early methods for producing the trench log relied upon ing additional photos of the trench margins recognizing that these
surveying or measurement from a reference grid to points along marginal photos will only partially cover portions of the trench
contacts and faults and transferring this information onto a sheet of wall. Most photographs should be taken orthogonal to the wall to
graph paper (McCalpin, 2009). To complete the logs in the field,
lines are visually interpolated on the log between measured points,
and characteristics of mapped units sketched in and described. To
increase the information contained and communicated by a trench
log, many paleoseismologists produce photomosaics for the expo-
sure, map directly onto these photos in the field, and create final
publication-quality logs. Using photographs as a base map captures
the rich tonal and textural information, but the traditional
approach to photomosaic production required time-intensive ad-
hoc rectification of individual photos. This rectification process
requires having a complete system of measured grid lines estab-
lished on the trench walls at a scale that allows individual photos to
capture ideally all four margins of a grid rectangle. The rectification
is performed by manually warping and distorting the image to
restore the reference grid lines within the photos to vertical and
horizontal, and then these individual photos are cropped and
assembled into the photomosaic. This manual rectification and
Fig. 6. Example of photomosaic production within Agisoft PhotoScan. Blue rectangles
mosaicking process is hindered by holes in the trench wall, clasts with black normal vectors (labelled with the image file name) represent the camera
and other features that protrude from the wall, and places where positions and orientations determined from the SfM process. The background image is
the reference grid is not flush with the trench wall surface. Because the orthorectified photomosaic that was exported to create Fig. 7.
170 S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178

ensure that complete wall coverage by photographs is attained and face. Nonetheless, in one instance, we were able to revisit photo-
to provide a systematic angular change between overlapping graphs collected in 2006 from a paleoseismic investigation on the
photographs. Additionally, a few photographs should be collected southern San Andreas fault near Coachella, California, that utilized
oblique to the trench wall to reduce possible error from broad-scale a 7 m deep trench with benched walls to expose a section of
warping of the reconstructed surface (e.g., James and Robson, alternating lacustrine and aeolian deposits spanning the past ~1000
2014). An example of a photomosaic produced by this process is years (Philibosian et al., 2009). Using just 15 photographs taken
shown in Fig. 7 (with the source 3D model provided in from the opposite side of the trench, we rapidly reconstructed the
Supplementary Materials) for a paleoseismic trench at a new geometry of a portion of the benched trench wall in 3D (Fig. 8a;
paleoseismic site on the Mojave section of the southern San Supplementary Material). Although mapping the stratigraphy has
Andreas fault near Elizabeth Lake, California. For this photo-based historically been done on a 2D representation of the trench wall,
reconstruction, we collected 191 photographs with a Tokina compiling this mapping onto the 3D reconstruction provides a
11e16 mm f/2.8 AT-X 116 Pro DX lens on a Nikon D7000 DSLR over more complete model for fault geometry and along-strike (cross-
the span of 29 min, covering one wall of the 17 m long, up to 4 m trench) variation.
deep, and ~1 m wide trench following the photo spacing illustrated In addition to the trench wall orthophotos, establishing the full
in Fig. 6. We imported this image collection into Agisoft PhotoScan 3D geometry of a trench facilitates robust structural and strati-
and used the masking tools to hide non-stationary objects and graphic analysis while avoiding the time-consuming process of
features that lie in the distant background outside of the trench surveying each fault and stratigraphic contact. Fig. 8 shows two
wall. The 3D reconstruction of these photographs took ~1 h total examples where we used a relatively small collection of photo-
using low resolution processing settings on a desktop computer graphs to reconstruct the geometry of an entire trench or complex
with a 2.8 GHz processor and 12 GB RAM. Lower resolution pro- trench wall (Fully interactive 3D models provided in
cessing may reduce the resolution of the 3D reconstruction of the Supplementary Material). A small reconnaissance trench (3 m long,
wall surface but does not impact the resolution of an output 1.5 m deep, and ~1 m wide) on the Denali fault, near Cantwell,
photomosaic. The output photomosaic resolution is controlled by Alaska, was fully captured by 13 photographs taken while standing
the quality of the input photographs and resolution parameters set on the ground surface outside the trench (Fig. 8b). The trench ge-
at the time of export. However, lower resolution processing may ometry was properly reconstructed in 3D from this small, highly
impact the accuracy of the photomosaic on the scale of several mm oblique, photograph collection. Additional photographs taken
if the resolution of the trench wall surface model is insufficient for orthogonal to the walls would improve resolution of the stratig-
proper orthorectification of input images. Some photo-based 3D raphy but are not critical for reconstructing trench geometry.
reconstruction software packages perform automatic colour and Although detailed surveying of contacts and gridlines are not
contrast matching across the mosaicked images to produce an necessary when photo-based 3D reconstructions are utilized, we
orthomosaic with consistent tone and colour spectrum. In our suggest that all paleoseismic studies should utilize high-precision
experience, we find this capability has the relatively minor disad- surveying instruments (e.g., total station or differential GPS) to
vantage of reducing overall contrast, and possibly clipping or flat- locate the 3D model in real world coordinates and provide precise
tening portions of the colour spectrum. This disadvantage is far relative positioning between successive excavations and for future
outweighed by the ability to manually enhance contrast, colour, examinations of the site.
saturation, etc. evenly across the entire photomosaic without An advantage for structural studies in paleoseismology vs. hard-
having to perform these adjustments individually for photos within rock structural geology is the ability in many paleoseismic studies
the photomosaic. to easily modify the target exposure. If faulting and/or stratigraphic
Producing a 2D photomosaic from more complex trench-wall relationships are not clear in an exposure, the trench wall can be cut
configurations simply requires strategic photograph collection to back to expose a new view with the hope that the relationships will
ensure capture of the full range of angles required to prevent oc- become clearer. Furthermore, this strategy can be implemented in a
clusion of portions of the trench wall. A particular example may be progressive fashion where the trench wall is cut back incremen-
paleoseismic trenches that are constructed of benched walls due to tally, with structure and stratigraphy documented on each fresh
their depth or unstable substrate, such that vertical faces are up to face for tracking changes in fault geometry and stratigraphic re-
~1.5 m tall and separated by a bench to the set-back higher wall lationships in the 3rd dimension. Unfortunately this destructive

Fig. 7. Orthorectified photomosaic created with photo-based 3D reconstruction techniques utilizing Agisoft Photoscan. No additional image adjustments were performed following
orthomosaic production except for rotation and downsampling. The even tone and contrast across the photomosaic along with the accurate spatial geometry improves the quality of
detailed mapping in the field and more completely preserves these characteristics for archiving and publication. This image is the east wall of one of the trenches at the Elizabeth
Lake paleoseismic site on the Mojave section of the southern San Andreas fault, California (view is looking to the southeast). The black box shows the area of photomosaic in Fig. 6.
The 3D model this image is derived from is provided as a 3D PDF in the Supplementary Material.
S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178 171

Fig. 8. Examples of geometric reconstructions from small collections of overview photographs. (a) oblique views of a section of an ~7 m tall, benched trench wall across the
southern San Andreas fault near Coachella, California. Blue rectangles and black vectors indicate the position and orientation of the photographs used to build this model. Although
these photos were not collected with photogrammetry in mind, there was sufficient overlap to reconstruct a detailed 3D model. Reed Burgette kneeling on the bottom bench for
scale. (b) A small (3 m wide, 1.5 m deep, 1 m wide) reconnaissance trench across the Denali fault near Cantwell, Alaska. As illustrated, this model was reconstructed from just 13
pictures taken from above the trench, but even with this highly oblique photography, the stratigraphy near the base of the trench resolves well. In particular, note how the
stratigraphy drapes over the boulder protruding from the trench wall. The 3D model these images are derived from is provided as a 3D PDF in the Supplementary Material.

technique obliterates the original stratigraphy and prevents the re- Conventional, high-resolution grid-mapping techniques suffer
examination of the stratigraphic record in the field. An important from time constraints and are either impractical to collect data at
positive outcome is that the photo-based 3D reconstruction pro- cm-scale resolutions over significant areas or take weeks to ach-
vides a solution for rapid and complete archival documentation of ieve. The results are also limited because the final product remains
each face created during a progressive excavation. After the face is a map interpretation that can only be verified by further visits to
cut, cleaned, and prepared, the face is photographed with either the study site. Photogrammetric datasets derived from UAV plat-
fixed reference points outside the zone of excavation or reference forms (Fig. 9) offer a cost-effective, ultra-high resolution alternative
points that are surveyed on each face. The faces are then recon- with a rapid acquisition time. They have the added advantage of
structed individually and compiled into a 3D model that records the providing access to vertical or unstable exposures, while delivering
removed stratigraphy. This 3D model enables the detailed inter- visual data in digital form that can be shared and reanalysed
pretation of the 3D geometry of faults and stratigraphic horizons, as without need to revisit the outcrop.
well as preserving the original tonal and textural information for Fig. 10 is a rendering and interpretation of coastal outcrop from
future investigations. A 3D archive of recent stratigraphy and Piccaninny Point on the northeast coast of Tasmania, Australia,
structure may prove to be a powerful tool for researchers of other which exposes a spectacular series of strike-slip faults. These faults
geologic subdisciplines or for paleoseismologists to re-interpret a comprise a damage zone of intersecting structures, crosscutting a
site using new insights that develop during later research. subvertical succession of metasedimentary sandstones and silt-
stones, belonging to the Mathinna Group (Banks, 1962; Gee and
5. Applications to structural geology Groves, 1971; Groves et al., 1977). We deployed an eight-rotor
Oktokopter (Fig. 9) with a small format digital camera (Canon
5.1. High-resolution mapping using UAV surveys 550D 15 Megapixel, DSLR with Canon EF-S 18e55 mm F/3.5-5.6 IS
lens). An onboard, navigation grade GPS receiver is integrated with
High-resolution fault, vein and fracture maps are routinely a Mikrokopter Flight Controller ME V2.0 and these permit an
created in structural geology to constrain, amongst other things, autonomous flight dictated by programmed waypoints. During this
the nucleation, growth, mechanics and scaling properties of frac- particular deployment, wind gusts reached speeds of 35 km/h,
tures (e.g. Shipton and Cowie, 2001; Wilson et al., 2009; Nixon requiring manual control of the UAV. Nevertheless, the strong wind
et al., 2011) and the permeability characteristics and sealing capa- conditions encountered demonstrated the robustness of the survey
bilities of fault systems (e.g., Willemse et al., 1997; Peacock et al., method. Approximately 140 outcrop photographs were collected
1998; Antonellini et al., 2008). Combinations of grid mapping, in- by the UAV. An area 100  100 m was covered in <5 min, at alti-
terpretations from overlapping outcrop photographs, aerial pho- tudes of 30e40 m, which produced high resolution imagery (1
tographs and detailed sketches are employed. More recently, digital pixel z 10 mm). Even lower altitudes of 15e20 m allow sub-cm
techniques such as terrestrial LiDAR scans are increasingly used, resolutions, although the advantage of this increase is countered
linked to satellite imagery and field observations (e.g., Pringle et al., by larger photographic datasets and associated increases in data-
2006; Wilson et al., 2009). processing time. In contrast, the best resolution available from
172 S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178

cloud containing many millions of points (Fig. 9bec). A real-world


coordinate system was derived for the output using the direct
georeferencing technique described in Turner et al. (2012). The
direct georeferencing technique links the timestamp of each
photograph to both the GPS position logged onboard at the time of
exposure and elevation as provided by the OktoKopter's barometric
altimeter (accuracies of ± 1 m during UAV flight), which derives
camera coordinates for each photograph. These coordinates are
then matched to the computed (Bundler) coordinates to solve for
the Helmert transformation parameters (3 translations, 3 rotations,
1 scale parameter). The derived transform is subsequently applied
to the sparse point cloud, resulting in thousands of real-world co-
ordinates linked to points in each image and allowing the images to
be orthorectified using a Delaunay triangulation. Turner et al.
(2012) report the absolute accuracy of the orthorectified images
produced from this method is typically < 400 mm. Using this direct
georeferencing technique, absolute accuracies are dominated by
the navigation grade GPS errors, estimation of the camera focal
length and interior/exterior orientation parameters and imprecise
synchronization between the GPS receiver and camera (Turner
et al., 2013). This series of steps is now largely implemented in
off-the-shelf software products such as Photoscan (Table 1).
Once georeferenced point cloud data are derived, obtaining
renderings is straightforward for direct use in structural geology,
such as textured wireframe models, orthorectified photomosaics
and DEMs. For example, the Piccaninny Point data were converted
to these datasets and rapidly digitized in a GIS environment
(Fig. 10), revealing a complex damage zone around an eroded fault
core (Buckley, 2013), which is the subject of ongoing research at The
University of Western Australia. The high-resolution nature of the
imagery, in combination with the intense layering of the outcrop,
allowed fault offset directions to be detected (Fig. 10d) and dis-
placementelength profiles to be calculated for each damage zone
structure. The pixel resolutions of 10 mm in the photomosaic and
20 mm in the DEM allowed for offsets greater than ~20e50 mm to
be identified with certainty over the 10,000 m2 outcrop area.
In cases where absolute location or improved accuracies are
required, real-world coordinates of Ground Control Points (GCPs)
can be established within the scene (James and Robson, 2012;
Turner et al., 2012). Typically, painted target markers are placed
in the area and surveyed using dual frequency differential GPS
Fig. 9. (a) Oktokopter UAV undergoing a test flight. A DSLR is suspended on a gimble
(horizontal accuracies of 20 mm and vertical accuracies of 40 mm)
and can be rotated in flight. (bec) SfM-derived point cloud of Piccaninny Point, viewed or total station techniques (<10 mm accuracies). For such an
from the southeast and above, respectively. The outcrop is 100 m long in this scene. A outcome, 10e15 markers placed throughout the area would be
wavecut rock platform is captured. Vegetation and a vehicle are present as noise. sufficient. When scale alone is required, only control distances
across the scene are required (James and Robson, 2012). Other
manned aircraft or satellites is typically 100e500 mm/pixel (e.g. important considerations to minimise error include using a high
Nixon et al., 2011). quality pre-calibrated camera (i.e. DSLR) with fixed optics, collect-
As with the paleoseismic studies, the workflow that followed ing slightly convergent imagery (Wackrow and Chandler, 2011;
acquisition of the photographs began with manual selection of the James and Robson, 2014), and surveying the outcrop using flight
most appropriate photos from the dataset. In this example, the lines from 2 orthogonal directions (P. Kovesi pers comm.). Recently,
images were processed using the Bundler software (Snavely, 2010). Turner et al. (2013) obtained large improvements to the absolute
Processing of the imagery was a semi-automated task, taking accuracy of the direct georeferencing technique with modifications
around 6 h to complete on a high-end desktop computer. Initially to the onboard GPS receiver, synchronization between the GPS
the SIFT algorithm (Lowe, 2004) is used to detect features across the receiver and camera and corrections between the GPS antenna and
images, and these features are then matched between overlapping camera position. Absolute accuracies of 100e200 mm were ach-
images. Bundler then used these matching features to complete a ieved (Turner et al., 2013). Similar improvements are only likely to
bundle adjustment and to align the images using an arbitrary co- accelerate in the next few years, and the need for GCPs may well be
ordinate system. An output file was generated that listed the eliminated in the near future at the point where direct georefer-
calculated position of each camera and each matched feature in the encing techniques achieve similar accuracies to GCP studies con-
Bundler arbitrary coordinate system. The result was a sparse point strained by differential GPS. Nonetheless, GCPs will continue to be
cloud in which the x, y, z position of each matched feature is listed useful where outcrop studies require ground-based verification or
along with its RGB colour from the original imagery. The point sub-cm accuracy (e.g. using total station surveys).
cloud was densified by use of the Patch-based Multview Stereo As a result of these developments, UAVs are likely to be a
software (PMVS2; Furukawa and Ponce, 2009) to produce a point commonly utilized tool by structural geologists in the near-future.
S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178 173

Fig. 10. (a) A selection of the overlapping photos captured by UAV, and their coverage of Piccaninny Point. (b) DEM derived from the point cloud shown in Fig. 9b, with an un-
derlying hillshade. Black dash e trace of large fault. (c) Orthorectified photomosaic for a portion of the outcrop. Pixel resolution 1 pixel ¼ 10 mm. (d) Structural interpretation of fault
and associated damage zone, showing dextral (red), sinistral (blue) and unidentified offset faults (grey) developed around a large fault (black dash), superimposed over the DEM.

A very large range of potential applications can be envisaged because they can fly at exceptionally low altitudes with generally
beyond the scope of this review. Nonetheless, it is worth providing higher quality cameras (relative to the size of the craft) mounted on
a brief note about the technology. UAVs are broadly subdivided into a stabilised platform. The Oktokopter employed in this case study is
fixed-wing types (e.g. model airplanes) and multi-rotor types such capable of covering approximately 20,000 m2 (2 ha) in a single
as used in the example above. For high-resolution studies, multi- flight with a 1.5 kg payload, and is particularly useful for vertical
rotor UAVs have significant advantages over fixed wing UAVs faces where imagery is required near to corners and surfaces at a
174 S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178

range of angles. On the other hand, fixed wing UAVs cover far larger
areas of ground in a smaller timeframe, and are less power hungry.
The main restrictions on the use of UAVs are from national regu-
latory frameworks, which vary widely from country to country. In
Australia for example, UAV use is governed by the Civil Aviation
Safety Authority (CASA; http://www.casa.gov.au/), which stipulates
that research use of UAVs requires controller's and operators cer-
tification to fly. This certification requires general aviation knowl-
edge in line with a pilot's licence. The UAV can fly no more than
120 m above ground unless special approval is provided, must
remain within line-of-sight, and must be flown over unpopulated
areas and outside controlled airspace.

5.2. Ground-based outcrop and open pit surveys

With the advent of SfM and its implementation in off-the-shelf


products (Table 1), photogrammetry is now an ideal tool for
ground-based structural studies because it generates digital 3D
models of natural outcrop, quarries/mine sites and hand specimens
(Fig. 11; Supplementary Material). Rock textures and fabrics can be
produced with high fidelity, using processing and visualization
tools such as Agisoft Photoscan™, which perform well in both the
ground-based studies discussed here and UAV-based studies
(Turner et al., 2013). Photo-based models also compare favourably
with laser scanning techniques when used to identify joints, dis-
continuities and orientations (e.g. Coggan et al., 2007), possibly
because photogrammetric data points inherently contain colour as
well as location and it is easier to derive information on surfaces
from multiple orientations using a hand-held DSLR rather than a
tripod-mounted scanner. It should be noted that laser-scanning
techniques are advancing as rapidly as photogrammetry so that a
number of these issues are now avoided (Hodgetts, 2013).
The vein array shown in Fig. 11 is a photo-based model of an en-
echelon sigmoidal vein array, which changes into planar en-
echelon veins on the alternate side of the hand specimen
(Fig. 11a). In this example, an ultra-high resolution model was
required and 100 photographs were collected under diffuse light,
using a 100 mm fixed focal length lens and Canon EOS-5 mark III
DSLR. Several different models were generated from the dataset
and it was found models with the best texture actually used less
than the full complement of 100 photographs. Fig. 11 was con-
structed from 51 photographs, which generated a densified point
cloud of 2,442,780 points and a wireframe with 493,195 elements.
The model was processed in ~4 h on a laptop with Intel i7 CPU and
8 Gb RAM.
The model faithfully reproduces the trajectories of individual
calcite vein fibres, sets of parallel, closely spaced microveins and
pressure solution seams, with only minor distortion present in the
orthorectified photomosaic of the front face (Fig. 11b). Photo-
grammetric datasets such as these are being used to refine our
existing kinematic and mechanical models for vein formation (e.g.
Beach, 1975; Olson and Pollard, 1991; Bons et al., 2012). As with the
UAV-based models of Piccaninny Point, hand specimens and out-
Fig. 11. (a) High fidelity textured photogrammetric model of a hand specimen from the
Cape Liptrap Formation, Wilsons Promontory, Victoria, Australia. The front and rear crops derived from ground-based studies can be georeferenced and
faces of the specimen show that 6 en-echelon sigmoidal veins link and develop into 3 converted to DEM and orthorectified photomosaics for mapping
planar en-echelon veins. The top and lower thirds of the hand specimen have been and analysis in a GIS environment. In addition, the open source 3D-
digitally removed. (b) Orthorectified photomosaic of the front face. Geological fabrics graphics package Blender (www.blender.org), allows true 3D
and geometries are reproduced in fine detail, with limited distortion in 2 spatially
restricted domains. (c) Photogrammetric model. The left-hand third of the model is
mapping to be carried out and projected onto the surface of the
textured, the other two-thirds reveal the underlying wireframe with such high reso- photo-based reconstruction (Fig. 11c). Spatial information
lution that individual triangular elements are not discernible. Pressure solution seams describing vein geometry and the orientation of features such as
(red), calcite veins (green) and calcite vein fibre trajectories (white) were mapped in vein surfaces or internal crystal fibres can be extracted. Given the
3D using the open-source Blender graphics environment. The 3D model these images
imagery, we have since been able to generate super high-resolution
are derived from is provided as a 3D PDF in the Supplementary Material.
models of individual veins and their fibres using overlapping im-
ages captured through bifocal microscopes.
S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178 175

For lower resolution models, the same sample was digitally Practical difficulties associated with these considerations are
reproduced using just 30 photographs, captured by a compact typically overcome by UAV-based surveys, which can provide su-
Canon S95 digital camera (i.e., not a DSLR). The photos were pro- perior coverage in a short period of time (<10 min). Nonetheless,
cessed in <1 h on a laptop, and even in this case, individual calcite from ground-based surveys, we have been able to generate 3D
fibres are visible. As such, the application of photogrammetry for geological maps, identify and track lithologies across inaccessible
pedagogical purposes is self-evident, because it is now being rela- sub-vertical faces and extract planimetric orthorectified photo-
tively easy to build a virtual library of useful hand specimens. mosaics with equivalent or higher resolution and less expense than
Furthermore, these models may be printed in 3D using inexpensive achieved by industry-standard aerial photographs (e.g., Fig. 11a).
web-based providers (e.g. http://www.shapeways.com/), which Following discretization of the wireframe meshes, mapping was
allows precious or fragile samples to be reproduced for lab-based completed in the interpolation modelling package Leapfrog Mining,
exercises. which is based on Fast Radial Basis Functions and has the potential
At larger scales, the same workflow is followed for ground- for true 3D mapping (sensu stricto McCaffrey et al., 2005), where
based digital mapping of open-pit minesites and large outcrops. photogrammetric datasets can be integrated with drill core logging,
Trials involving ground-based digital mapping of a >500 m long multi-element chemistry and geophysical data etc.
open pit required 150e250 photos with a 28 mm focal length lens,
and 500e700 photos with a 50 mm lens (Fig. 12). Larger focal 5.3. The future of photogrammetry and structural interpretation
lengths lead to higher resolution models but, as a general rule,
twice the number of photographs requires a four-fold increase in Photogrammetry and other digital techniques, such as photo-
processing time. Our preliminary trials have shown a number of realistic laser scanning (Hodgetts, 2013), offer a step-change in
additional parameters must be considered to achieve satisfactory the amount of data available from outcrop. These techniques are
results, and that model quality is influenced by three main part of an arsenal of digital approaches to field mapping that are
parameters: now available (McCaffrey et al., 2005) and may become particularly
powerful as we develop techniques to merge and automatically
(1) Lighting conditions: Reflective surfaces and strong contrasts analyse photogrammetric data with other potential field and
in light across a scene negatively affect point matching. remote sensing data (e.g. aeromagnetics, hyperspectral and radio-
Diffuse lighting conditions are preferable. metric data). In addition, with the advent of Virtual Globes, such as
(2) Duration of survey: Because the sun's azimuth continues to Google Earth, it becomes possible to visualize photogrammetric
change as a survey progresses, point matching between alongside other structural data (Blenkinsop, 2012) on a carto-
photographs becomes complicated by changes in shadow graphic representation of the Earth (De Paor and Whitmeyer, 2011).
length and surface albedo. It was found that model quality As demonstrated, photo-based 3D approaches permit the ca-
degrades significantly for durations >30 min. For long- pacity for mm-cm scale resolution, over many hundreds of metres,
duration surveys, this effect can be circumnavigated by if not more in the case of fixed wing UAVs. Nonetheless, digital
returning to the outcrop at approximately the same time the mapping of the data using manual approaches remains time
next day if similar weather conditions prevail. consuming. The fault-fracture interpretation of orthorectified im-
(3) Image network geometry: The capture of photographs from a ages extracted from Piccaninny Point (Fig. 10d) required ~2 weeks
limited number of poorly distributed locations (stations) can of manual digitizing of polylines and the building of a database of
lead to model distortions (e.g. Wackrow and Chandler, 2011; attributes (e.g. slip sense, offset etc). This raises the twin questions
James and Robson, 2014) and missing regions. The use of of how can we use such data more efficiently, and how can we
GCPs and convergent imagery is important to minimize any extract information not readily available to manual interpretation?
such distortions. Convergent imagery will also allow recon- Two promising approaches exist to aid rapid mapping and in-
struction of complex surfaces with a wide variety of face formation extraction. The first involves image analysis of the DEM
directions. Outcrops, mine sites and quarries are best and orthorectified photomosaic data (e.g. Stumpf et al., 2013;
reproduced if access is available all around (including inside) Vasuki et al., 2014). The second involves identification, mapping
and images are captured in an organised semi-continuous and classification of point cloud attributes using Artificial Intelli-
manner. gence approaches (Hodgetts, 2013). Recently, a semi-automatic

Fig. 12. (a) A high fidelity textured photogrammetric model of a legacy open pit mine from the Coolgardie greenstone domain, Western Australia. The pit is > 500 m long and this
model was constructed using 600 ground-based photos from a 50 mm focal length lens. Features on the scale of ~10 cm can readily be detected in the model and the spatial
resolution of orthorectified images extracted from the data compete with those obtained by the highest quality aerial photographs and can be obtained at any angle of projection.
(b) Oblique view of the same model.
176 S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178

method for the rapid mapping of discontinuities like faults was


developed using phase congruency and phase symmetry as edge
detection methods (Micklethwaite et al., 2012; Vasuki et al., 2013,
2014) and user interaction to guide the process. Fig. 13 shows
stages in the process and a comparison with a manually digitized
fault map. A user rapidly defines the broad location and length of
faults in the outcrop (Fig. 13a), then edges detected in the data are
used to construct the true geometry of the faults (Fig. 13bec). In the
example shown, the user-guided interpretation was completed in
10 min, while manual digitizing took approximately 7 h, repre-
senting a significant increase in efficiency when interpreting the
data. In a later stage (not shown), the detected fault traces were
combined with the point cloud data to extract orientation data
systematically along the faults (Vasuki et al., 2014) using the
RANSAC algorithm (Fischler and Bolles, 1981) to best-fit planes
through points lying along the fault. Simple triangulation or tensor
analysis approaches can also be used to identify surface orienta-
tions of faults, fractures or bedding surfaces (Feng et al., 2001;
Fern andez, 2005).
Secondly, as highlighted by Hodgetts (2013 and references
therein) analysis of the attributes of point cloud data can emphasise
patterns and textures not obvious at first inspection. Artificial In-
telligence approaches such as Neural Networks, Fuzzy Logic and
Evolutionary Algorithms allow the automatic classification of point
cloud data, aiding the identification of varying stratigraphy (van
Lanen et al., 2009; Rarity et al., 2013) or the extraction and
upscaling of fault and fracture populations (e.g. Gillespie et al.,
2010; Seers and Hodgetts, 2013), especially when combined with
field observations. The approach is implemented in Virtual Reality
Geological Studio (VRGS; developed in-house at the University of
Manchester), and has been applied mostly to laser scan data
(Hodgetts, 2013) but is directly applicable to photogrammetric
point clouds.

6. Discussion and conclusions

With a variety of software implementations of photo-based 3D


reconstruction, including open-source options (Table 1), these
techniques represent the ‘democratization’ of high-resolution 3D
geospatial data collection by making the collection of these data
available to anyone with a computer and a digital camera.

 Photo-based 3D reconstruction techniques have a broad spec-


trum of applications in structural geology and neotectonics due
to the ability to collect vast quantities of high-resolution 3D
geospatial data across multiple spatial scales. The limited
infrastructure requirements that increase portability and
decrease cost may allow photo-based 3D reconstruction tech-
niques to supersede other common high-resolution surface
modeling techniques in many research settings. Photo-based 3D
reconstruction provides an opportunity for archiving and
sharing of high fidelity imagery and 3D geometry from critical
rock exposures and sediments. It is made particularly powerful
because of its ability to easily collect data from inaccessible or
unsafe exposures, or to record a time series of data such as when
Fig. 13. (a) A sketch map of the approximate locations of faults in an orthorectified
paleoseismic trenches are successively cut back.
image, using a limited number of clicks (approximately 10 min interpretation). (b) A  Resolution and accuracy during 3D reconstruction is dependent
semi-automatic fault map constructed by matching the user defined approximate fault upon a number of parameters. Nonetheless, it is relatively
locations with automatically detected edges, to construct realistic fault geometry, straightforward to design a survey and processing routine that
segmentation and lengths. (c) Comparison between the semi-automatic fault map and
suits the resolution needs of the project. The relative precision
a manually digitized map derived in a standard GIS-environment. Dark blue e semi-
automatically identified discontinuities; red e manually digitized faults; light blue e ratio described by James and Robson (2012) is a useful guide
manually digitized joints; green e manually digitized extension fractures. The semi- when designing a data collection routine by illustrating the
automatically mapped faults have the same geometry and approximate lengths as expected order of measurement precision relative to the
the manually digitized counterparts, without false positives. Finer detail would have observation distance based upon observed capabilities of photo-
been possible with slightly longer user interaction.
based 3D reconstruction software. For example, with this ratio
S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178 177

generally exceeding 1:1000 (James and Robson, 2012), a study Barazzetti, L., Remondino, F., Scaioni, M., 2010. Automation in 3D Reconstruction:
Results on Different Kinds of Close-Range Blocks. IAPRS&SIS.
that requires mm-scale precision will require photographs to be
Beach, A., 1975. The geometry of en-echelon vein arrays. Tectonophysics 28,
collected within several meters of the target in addition to the 245e263.
requirements for well-distributed camera positions and ground Bevis, M., Hudnut, K., Sanchez, R., Toth, C., Grejner-Brzezinska, D., Kendrick, E.,
control. Caccamise, D., Raleigh, D., Zhou, H., Shan, S., Shindle, W., Yong, A., Harvey, J.,
Borsa, A., Ayoub, F., Shrestha, R., Carter, B., Sartori, M., Phillips, D., Coloma, F.,
 Future directions in structural geology and paleoseismology 2005. The B4 Project: Scanning the San Andreas and San Jacinto Fault Zones.
using photo-based 3D reconstruction derive from the rich AGU Fall Meet. Abstr. 34, 01.
dataset that includes complete tonal and textural information Birdseye, C.H., 1940. Stereoscopic Phototopographic mapping. Ann. Assoc. Am.
Geogr. 30, 1e24. http://dx.doi.org/10.1080/00045604009357193.
combined with a high-resolution 3D model. Novel image anal- Blenkinsop, T.G., 2012. Visualizing structural geology: from Excel to google earth.
ysis methods are becoming available to rapidly generate maps Comput. Geosci. 45, 52e56.
from the large geospatial datasets, without spending days to Bons, P.D., Elburg, M.A., Gomez-Rivas, E., 2012. A review of the formation of tectonic
veins and their microstructures. J. Struct. Geol. 43, 33e62.
weeks manually digitizing (e.g. Vasuki et al., 2014). Further- Borsa, A., Minster, J.-B., 2012. Rapid determination of Near-Fault earthquake
more, Artificial Intelligence-type algorithms provide new ways deformation using differential LiDAR. Bull. Seismol. Soc. Am. 102, 1335e1347.
to extract meaningful sedimentological and stratigraphic pa- http://dx.doi.org/10.1785/0120110159.
Brooks, B.A., 2014. Briefing on the M6.0 south Napa earthquake. In: Southern Cal-
rameters from exposures using the combined 3D geometry and
ifornia Earthquake Center Annual Meeting, September 6-10, 2014. Palm Springs,
image information (Hodgetts, 2013). California.
Finally, photo-based 3D models offer the ability to build virtual Buckley, D., 2013. An Analysis of Fault Damage at Piccaninny Point, North-East
Tasmania. Unpublished Hons thesis. University of Western Australia.
archives of geoscientific data with relatively low cost or exper-
Coggan, J.S., Wetherelt, A., Gwynn, X.P., Flynn, Z., 2007. Comparison of hand-
tise required. In the realm of the communication of scientific mapping with remote data capture systems for effective rock mass character-
results in Structural Geology and Paleoseismology, this ability isation. In: Proceedings of 11th Congress of the International Society for Rock
ought to lead to the development of electronic archives of crit- Mechanics - the Second Half Century of Rock Mechanics, vol. 1, pp. 201e205.
De Paor, D.G., Whitmeyer, S.J., 2011. Geological and geophysical modeling on virtual
ical outcrop or hand-specimen observations to accompany globes using KML, COLLADA, and Javascript. Comput. Geosci. 37, 100e110.
publications. Such a process would be analogous to electronic Dolan, J.F., Haravitch, B.D., 2014. How well do surface slip measurements track slip
lab-books common to the biosciences. Secondly, there are at depth in large strike-slip earthquakes? the importance of fault structural
maturity in controlling on-fault slip versus off-fault surface deformation. Earth
obvious pedagogical applications. Many students find difficulty Planet. Sci. Lett. 388, 38e47. http://dx.doi.org/10.1016/j.epsl.2013.11.043.
in visualizing the 3D nature of geological features when these Duffy, B., Quigley, M., Barrell, D.J.A., Dissen, R.V., Stahl, T., Leprince, S., McInnes, C.,
are shown as standard photographs. 3D models of outcrops and Bilderback, E., 2013. Fault kinematics and surface deformation across a releasing
bend during the 2010 MW 7.1 Darfield, New Zealand, earthquake revealed by
hand specimens, which can be manipulated individually by differential LiDAR and cadastral surveying. Geol. Soc. Am. Bull. 125, 420e431.
students, would help to overcome this hurdle in visualization http://dx.doi.org/10.1130/B30753.1.
and learning. Rather than replacing field trips, these interactive Eardley, A.J., 1942. Aerial Photographs: Their Use and Interpretation. Harper and
Brothers, New York.
models could allow students to revisit key exposures as new
Farenzena, M., Fusiello, A., Gherardi, R., 2009. Structure-and-motion pipeline on a
concepts are discussed in order to layer the concepts and rein- hierarchical cluster tree, In: 2009 IEEE 12th International Conference on
force the connections between the concepts and real world Computer Vision Workshops (ICCV Workshops). In: Presented at the 2009 IEEE
12th International Conference on Computer Vision Workshops (ICCV Work-
examples.
shops), pp. 1489e1496. http://dx.doi.org/10.1109/ICCVW.2009.5457435.
Favalli, M., Fornaciai, A., Isola, I., Tarquini, S., Nannipieri, L., 2012. Multiview 3D
reconstruction in geosciences. Comput. Geosci. 44, 168e176. http://dx.doi.org/
Acknowledgements 10.1016/j.cageo.2011.09.012.
Feng, Q., Sjo€gren, P., Stephansson, O., Jing, L., 2001. Measuring fracture orientation at
Cees Passchier is thanked for his ideas and encouragement to exposed rock faces by using a non-reflector total station. Eng. Geol. 59,
133e146.
submit this paper. David Hodgetts hosted and introduced SM to Fernandez, O., 2005. Obtaining a best fitting plane through 3D georeferenced data.
VRGS and its application for photogrammetric data. Tom Blenkin- J. Struct. Geol. 27, 855e858.
sop and an anonymous reviewer are thanked for reviews that Fischler, M.A., Bolles, R.C., 1981. Random sample consensus: a paradigm for model
fitting with applications to image analysis and automated cartography. Com-
contributed to the clarity of the paper. Southern California Earth- mun. ACM 24, 381e395.
quake Center award #13136 to SB supported the work at Elizabeth Fonstad, M.A., Dietrich, J.T., Courville, B.C., Jensen, J.L., Carbonneau, P.E., 2013.
Lake, CA, presented here. SM was supported by the Hammond- Topographic structure from motion: a new development in photogrammetric
measurement. Earth Surf. Process. Landf. 38, 421e430. http://dx.doi.org/
Nisbet Endowment during this work.
10.1002/esp.3366.
Furukawa, Y., Ponce, J., 2009. Accurate, dense, and robust multi-view stereopsis.
IEEE Trans. Pattern Anal. Mach. Intell. 32, 1362e1376.
Appendix A. Supplementary data Furukawa, Y., Curless, B., Seitz, S.M., Szeliski, R., 2010. Towards Internet-scale multi-
view stereo. In: 2010 IEEE Conference on Computer Vision and Pattern
Supplementary data related to this article can be found at http:// Recognition (CVPR), pp. 1434e1441. http://dx.doi.org/10.1109/
CVPR.2010.5539802. Presented at the 2010 IEEE Conference on Computer
dx.doi.org/10.1016/j.jsg.2014.10.007. Vision and Pattern Recognition (CVPR).
Gee, R.D., Groves, D.I., 1971. Structural features and mode of emplacement of
part of the Blue Tier Batholith in Northeast Tasmania. J. Geol. Soc. Aust. 18,
References 41e56.
Gillespie, P., Monsen, E., Maerten, L., Hunt, D., Thurmond, J., Tuck, D., 2010. Fractures
Abdullah, Q., Bethel, J., Hussain, M., Munjy, R., 2013. Photogrammetric project and in carbonates: from digital outcrops to mechanical models. In: Martinsen, O.J.,
mission planning. In: McGlone, J.C. (Ed.), Manual of Photogrammetry. American Pulham, A.J., Haughton, P., Sullivan, M.D. (Eds.), Outcrops Revitalized: Tools,
Society for Photogrammetry and Remote Sensing, pp. 1187e1220. Techniques and Applications. SEPM (Society for Sedimentary Geology).
Akçiz, S.O., Ludwig, L.G., Arrowsmith, J.R., Zielke, O., 2010. Century-long average Gold, P.O., Cowgill, E., Kreylos, O., Gold, R.D., 2012. A terrestrial lidar-based work-
time intervals between earthquake ruptures of the San Andreas fault in the flow for determining Threeedimensional slip vectors and associated un-
Carrizo Plain, California. Geology 38, 787e790. http://dx.doi.org/10.1130/ certainties. Geosphere 8, 431e442. http://dx.doi.org/10.1130/GES00714.1.
G30995.1. Granshaw, S.I., 1980. Bundle adjustment methods in engineering photogrammetry.
Antonellini, M., Tondi, E., Agosta, F., Aydin, A., Cello, G., 2008. Failure modes in deep- Photogramm. Rec. 10, 181e207.
water carbonates and the impact for fault development: Majella Mountain, Groves, D.I., Cocker, J.D., Jennings, D.J., 1977. The geology, geochemistry and min-
Central Apennines, Italy. Mar. Pet. Geol. 25, 1074e1096. eralisation of the Blue Tier Batholith. Geol. Surv. Tasman. Bull. 55, 7e116.
Arrowsmith, J.R., Zielke, O., 2009. Tectonic geomorphology of the San Andreas fault Haddad, D.E., Akçiz, S.O., Arrowsmith, J.R., Rhodes, D.D., Oldow, J.S., Zielke, O.,
zone from high resolution topography: an example from the cholame segment. Toke , N.A., Haddad, A.G., Mauer, J., Shilpakar, P., 2012. Applications of airborne
Geomorphology 113, 70e81. http://dx.doi.org/10.1016/j.geomorph.2009.01.002. and terrestrial laser scanning to paleoseismology. Geosphere 8, 771e786.
Banks, M.R., 1962. The Silurian and Devonian systems. J. Geol. Soc. Aust. 9, 177e188. http://dx.doi.org/10.1130/GES00701.1.
178 S.P. Bemis et al. / Journal of Structural Geology 69 (2014) 163e178

Haeussler, P.J., Schwartz, D.P., Dawson, T.E., Stenner, H.D., Lienkaemper, J.J., Peacock, D.C.P., Fisher, Q.J., Willemse, E.J.M., Aydin, A., 1998. The relationship be-
Sherrod, B., Cinti, F.R., Montone, P., Craw, P.A., Crone, A.J., Personius, S.F., 2004. tween faults and pressure solution seams in carbonate rocks and implications
Surface rupture and slip distribution of the denali and Totschunda faults in the for fluid flow. Geol. Soc. Lond. Special Publ. 147, 105e115.
3 november 2002 M 7.9 earthquake, Alaska. Bull. Seismol. Soc. Am. 94, Philibosian, B., Fumal, T.E., Weldon, R.J., Kendrick, K.J., Scharer, K.M., Bemis, S.P.,
S23eS52. http://dx.doi.org/10.1785/0120040626. Burgette, R.J., Wisely, B.A., 2009. Photomosaics and Logs of Trenches on the San
Hagan, T.O., 1980. A case for terrestrial photogrammetry in deep-mine rock struc- Andreas Fault Near Coachella. California (Open-File Report No. 2009-1039). U.S.
ture studies. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 17, 191e198. http:// Geological Survey.
dx.doi.org/10.1016/0148-9062(80)91085-2. Pillmore, C.L., 1964. Application of high-order stereoscopic plotting instruments to
Haugerud, R.A., Harding, D.J., Johnson, S.Y., Harless, J.L., Weaver, C.S., Sherrod, B.L., photogeologic studies. In: Procedures and Studies in Photogeology, Geological
2003. High-resolution lidar topography of the Puget Lowland, Washington - a Survey Bulletin 1043-B. U.S. Department of the Interior, Geological Survey,
bonanza for earth science. GSA Today 13, 4e10. Washington, D.C, pp. 22e34.
Hilley, G.E., DeLong, S., Prentice, C., Blisniuk, K., Arrowsmith, J., 2010. Morphologic Pringle, J.K., Howell, J.A., Hodgetts, D., Westerman, A.R., Hodgson, D.M., 2006. Vir-
dating of fault scarps using airborne laser swath mapping (ALSM) data. Geo- tual outcrop models of petroleum reservoir analogues: a review of the current
phys. Res. Lett. 37, L04301. http://dx.doi.org/10.1029/2009GL042044. state-of-the-art. First Break 24, 33e42.
Hodgetts, D., 2013. Laser scanning and digital outcrop geology in the petroleum Rarity, F., van Lanen, X.M.T., Hodgetts, D., Gawthorpe, R.L., Wilson, P., Fabuel-
industry: a review. Mar. Pet. Geol. 46, 335e354. http://dx.doi.org/10.1016/ Perez, I., Redfern, J., 2013. LiDAR-based Digital Outcrops for Sedimentological
j.marpetgeo.2013.02.014. Analysis: Workflows and Techniques. In: Geological Society of London Special
Hudnut, K.W., Borsa, A., Glennie, C., Minster, J.-B., 2002. High-resolution topography Publications 387. http://dx.doi.org/10.1144/SP387.5.
along surface rupture of the 16 October 1999 Hector mine, California, earth- Remondino, F., 2006. Detectors and descriptors for photogrammetric applications.
quake (Mw 7.1) from airborne laser swath mapping. Bull. Seismol. Soc. Am. 92, In: Fo €rstner, W., Steffen, R. (Eds.), International Archives of the Photogram-
1570e1576. http://dx.doi.org/10.1785/0120000934. metry, Remote Sensing and Spatial Information Sciences: Symposium of ISPRS
James, M.R., Robson, S., 2012. Straightforward reconstruction of 3D surfaces and Commission III Photogrammetric Computer Vision PCV ' 06, pp. 49e54. Bonn,
topography with a camera: accuracy and geoscience application. J. Geophys. Germany.
Res. 117, F03017. http://dx.doi.org/10.1029/2011JF002289. Seers, T., Hodgetts, D., 2013. Comparison of Digital Outcrop and Conventional Data
James, M.R., Robson, S., 2014. Mitigating systematic error in topographic models Collection Approaches for the Characterization of Naturally Fractured Reservoir
derived from UAV and ground-based image networks. Earth Surf. Process. Analogues. In: Geological Society of London Special Publication 374. http://
Landf. 39, 1413e1420. http://dx.doi.org/10.1002/esp.3609. dx.doi.org/10.1144/SP374.13.
Johnson, K., Nissen, E., Saripalli, S., Arrowsmith, J.R., McGarey, P., Scharer, K., Shipton, Z.K., Cowie, P.A., 2001. Damage zone and slip-surface evolution over um to
Williams, P., Blisniuk, K., 2014. Rapid mapping of ultrafine fault zone topog- km scales in high-porosity Navajo sandstone, Utah. J. Struct. Geol. 23,
raphy with structure from motion. Geosphere. http://dx.doi.org/10.1130/ 1825e1844.
GES01017.1. GES01017.1. Smith, M.J., Chandler, J., Rose, J., 2009. High spatial resolution data acquisition for
Krauss, K., 1993. Photogrammetry. In: Fundamentals and Standard Processes, vol. 1. the geosciences: kite aerial photography. Earth Surf. Process. Landforms 34,
Dümmlers. 155e161. http://dx.doi.org/10.1002/esp.1702.
Krishnan, A.K., Saripalli, S., Nissen, E., Arrowsmith, R., 2012. 3D change detection Snavely, N., 2010. Bundler: Structure from Motion for Unordered Image Collections.
using low cost aerial imagery. In: 2012 IEEE International Symposium on Safety, http://phototour.cs.washington.edu/bundler/.
Security, and Rescue Robotics (SSRR), pp. 1e6. http://dx.doi.org/10.1109/ Snavely, N., Seitz, S.M., Szeliski, R., 2006. Photo tourism: exploring photo collections
SSRR.2012.6523892. Presented at the 2012 IEEE International Symposium on in 3D. In: ACM SIGGRAPH 2006 Papers, SIGGRAPH ’06. ACM, New York, NY, USA,
Safety, Security, and Rescue Robotics (SSRR). pp. 835e846. http://dx.doi.org/10.1145/1179352.1141964.
Kuenen, P.H., 1950. Stereoscopic projection for demonstration in geology, geo- Snavely, N., Garg, R., Seitz, S.M., Szeliski, R., 2008a. Finding paths through the
morphology, and other natural sciences. J. Geol. 58, 49e54. http://dx.doi.org/ World's photos. In: ACM SIGGRAPH 2008 Papers, SIGGRAPH '08. ACM, New
10.1086/625694. York, NY, USA, pp. 15:1e15:11. http://dx.doi.org/10.1145/1399504.1360614.
Lowe, D.G., 2004. Distinctive image features from scale-invariant keypoints. Int. J. Snavely, N., Seitz, S.M., Szeliski, R., 2008b. Modeling the world from internet photo
Comput. Vis. 60, 91e110. http://dx.doi.org/10.1023/B: VISI.00000296 collections. Int. J. Comput. Vis. 80, 189e210. http://dx.doi.org/10.1007/s11263-
64.99615.94. 007-0107-3.
McCaffrey, K.J.W., Jones, R.R., Holdsworth, R.E., Wilson, R.W., Clegg, P., Imber, J., Stumpf, A., Malet, J.-P., Kerle, N., Niethammer, U., Rothmund, S., 2013. Image-based
Holliman, N., Trinks, I., 2005. Unlocking the spatial dimension: digital tech- mapping of surface fissures for the investigation of landslide dynamics. Geo-
nologies and the future of geosciences fieldwork. J. Geol. Soc. Lond. 162, morphology 186, 12e27. http://dx.doi.org/10.1016/j.geomorph.2012.12.010.
927e938. Turner, D., Lucieer, A., Watson, C., 2012. An automated technique for generating
McCalpin, J., 2009. Paleoseismology, second ed. In: International Geophysics Aca- georectified mosaics from ultra-high resolution unmanned aerial vehicle (UAV)
demic Press. imagery, based on structure from motion (SfM) Point clouds. Remote Sens. 4,
Micklethwaite, S., Turner, D., Vasuki, Y., Kovesi, P., Holden, E.-J., Lucieer, A., 2012. 1392e1410.
Mapping from an Armchair: rapid, high-resolution mapping using UAV and Turner, D., Lucieer, A., Wallace, L., 2013. Direct georeferencing of ultrahigh-
computer vision technology. Struct. Geol. Resour. 130e133. resolution UAV imagery. IEEE Transactions on Geoscience and Remote
Moreels, P., Perona, P., 2007. Evaluation of features detectors and descriptors based Sensing 52, 2738e2745. http://dx.doi.org/10.1109/TGRS.2013.2265295.
on 3D objects. Int. J. Comput. Vis. 73, 263e284. http://dx.doi.org/10.1007/ van Lanen, X.M.T., Hodgetts, D., Redfern, J., Fabuel-Perez, I., 2009. Applications of
s11263-006-9967-1. digital outcrop models: two fluvial case studies from the Triassic Wolfville Fm.,
Morelan, A.E., Stock, J.M., Hudnut, K.W., Akciz, S.O., 2010. Using photogrammetry to Canada and Oukaimeden Sandstone Fm., Morocco. Geol. J. 44, 742e760. http://
produce high-resolution DEMs of the El Mayor-Cucapah surface rupture. In: dx.doi.org/10.1002/gj.1196.
Southern California Earthquake Center Annual Meeting, September, 2010. Palm Vasuki, Y., Holden, E.-J., Kovesi, P., Micklethwaite, S., 2013. A geological structure
Springs, California. mapping tool using photogrammetric data. In: ASEG Extended Abstracts, pp. 1e4.
Niethammer, U., James, M.R., Rothmund, S., Travelletti, J., Joswig, M., 2012. UAV- Vasuki, Y., Holden, E.-J., Kovesi, P., Micklethwaite, S., 2014. Semi-automatic mapping
based remote sensing of the Super-Sauze landslide: evaluation and results. Eng. of geological Structures using UAV-based photogrammetric data: an image
Geol., Integration Technol. Landslide Monit. Quant. Hazard Assess. 128, 2e11. analysis approach. Comput. Geosci. 69, 22e32. http://dx.doi.org/10.1016/j.
http://dx.doi.org/10.1016/j.enggeo.2011.03.012. cageo.2014.04.012.
Nissen, E., Krishnan, A.K., Arrowsmith, J.R., Saripalli, S., 2012. Three-dimensional Wackrow, R., Chandler, J.H., 2011. Minimising systematic error surfaces in digital
surface displacements and rotations from differencing pre- and post- elevation models using oblique convergent imagery. Photogramm. Rec. 26,
earthquake LiDAR point clouds. Geophys. Res. Lett. 39, n/aen/a. http:// 16e31.
dx.doi.org/10.1029/2012GL052460. Westoby, M.J., Brasington, J., Glasser, N.F., Hambrey, M.J., Reynolds, J.M., 2012.
Nixon, C.W., Sanderson, D.J., Bull, J.M., 2011. Deformation within a strike-slip fault “Structure-from-Motion” photogrammetry: a low-cost, effective tool for geo-
network at Westward Ho!, Devon U.K.: domino vs conjugate faulting. J. Struct. science applications. Geomorphology 179, 300e314. http://dx.doi.org/10.1016/
Geol. 33, 833e843. j.geomorph.2012.08.021.
Olson, J.E., Pollard, D.D., 1991. The initiation and growth of en e chelon veins. Willemse, E.J.M., Peacock, D.C.P., Aydin, A., 1997. Nucleation and growth of strike-
J. Struct. Geol. 13, 595e608. slip faults in limestones from Somerset, U.K. J. Struct. Geol. 19, 1461e1477.
Oskin, M.E., Le, K., Strane, M.D., 2007. Quantifying fault-zone activity in arid envi- Wilson, P., Gawthorpe, R.L., Hodgetts, D., Rarity, F., Sharp, I.R., 2009. Geometry and
ronments with high-resolution topography. Geophys. Res. Lett. 34, L23S05. architecture of faults in a syn-rift normal fault array: the Nukhul half-graben,
http://dx.doi.org/10.1029/2007GL031295. Suez rift, Egypt. J. Struct. Geol. 31, 759e775.
Oskin, M.E., Arrowsmith, J.R., Corona, A.H., Elliott, A.J., Fletcher, J.M., Fielding, E.J., Zielke, O., Arrowsmith, J.R., Ludwig, L.G., Akçiz, S.O., 2010. Slip in the 1857 and
Gold, P.O., Garcia, J.J.G., Hudnut, K.W., Liu-Zeng, J., Teran, O.J., 2012. Near-field earlier large earthquakes along the carrizo Plain, San Andreas fault. Science 327,
deformation from the El MayoreCucapah earthquake revealed by differential 1119e1122. http://dx.doi.org/10.1126/science.1182781.
LIDAR. Science 335, 702e705. http://dx.doi.org/10.1126/science.1213778.

Potrebbero piacerti anche