Sei sulla pagina 1di 32

CHAPTER 5

Digital terrain modeling and glacier


topographic characterization
Duncan J. Quincey, Michael P. Bishop, Andreas Kääb, Etienne Berthier, Boris Flach,
Tobias Bolch, Manfred Buchroithner, Ulrich Kamp, Siri Jodha S. Khalsa, Thierry Toutin,
Umesh K. Haritashya, Adina Racoviteanu, John F. Shroder, and
Bruce H. Raup

ABSTRACT 5.1 INTRODUCTION

The Earth’s topography results from dynamic inter- The Earth’s topography is the result of dynamic
actions involving climate, tectonics, and surface interactions involving climate, tectonics, and sur-
processes. In this chapter our main interest is in face processes (Molnar and England 1990, Bishop
describing and illustrating how satellite-derived et al. 2003). Numerous feedback mechanisms oper-
DEMs (and other DEMs) can be used to derive ate over unique spatiotemporal scales that govern
information about glacier dynamical changes. hypsometry and terrain parameters. Similarly, the
Along with other data that document changes in topography governs a variety of physical param-
glacier area, these approaches can provide useful eters and processes including stress fields, erosion
measurements of, or constraints on glacier volume and mass wasting, precipitation, sediment deposi-
balance and—with a little more uncertainty related tion, surface and groundwater flow direction and
to the density of lost or gained volume—mass bal- ponding, surface energy budget, and glaciation in a
ance. Topics covered include: basics on DEM gen- complex topographic feedback loop (Roe et al.
eration using stereo image data (whether airborne 2002, Reiners et al. 2003, Arnold et al. 2006a).
or spaceborne), the use of ground control points Scientists have long recognized the significance of
and available software packages, postprocessing, topographic parameters for studying and modeling
and DEM dataset fusion; DEM uncertainties and process mechanics, mapping the landscape, and
errors, including random errors and biases; various detecting spatiotemporal change. Consequently,
glacier applications including derivation of relevant digital terrain modeling and the generation and
geomorphometric parameters and modeling of analysis of terrain surface representations,
topographic controls on radiation fields; and the commonly referred to as digital elevation models
important matters of glacier mapping, elevation (DEMs) and geomorphometry, have become an
change, and mass balance assessment. Altimetric important research topic for studying and solving
data are increasingly important in glacier studies, problems in hydrology, geomorphology, and
yet challenges remain with availability of high-qual- glaciology (Wilson and Gallant 2000).
ity data, the current lack of standardization for Glaciological research indicates that glaciers in
methods for acquiring, processing, and representing most areas around the world are receding (Barry
digital elevation data, and the identification and 2006, Gardner et al. 2013). Internationally, scien-
quantification of DEM error and uncertainty. tists are attempting to inventory, monitor, and
114 Digital terrain modeling and glacier topographic characterization

understand current ice mass distributions and vol- The purpose of this chapter is to address the topic
ume, glacier–climate interactions, glacier sensitivity of digital terrain modeling and geomorphometry
to climate change, as well as assess the impact of for glacier assessment and mapping. Specifically,
glacier change related to water resources, natural we consider numerous issues that must be effec-
hazards, and sea level fluctuations (Meier 1984, tively dealt with in order to estimate selected glacier
Dyurgerov and Meier 2000, Haeberli et al. 2000, parameters and delineate glacier boundaries. We
Kaser 2001, Arendt et al. 2002, Huggel et al. first provide background information on digital
2002, Meier et al. 2003, Bishop et al. 2004, Kargel terrain modeling to set the stage for DEM produc-
et al. 2005). With many high-altitude glacial lakes tion using satellite imagery and methodological
now forming in alpine areas (Huggel et al. 2002, approaches. We then provide examples of analytical
Gardelle et al. 2011), with the threat of devastating techniques that can be used to study glaciers.
landslides ever present and perhaps shifting (Kargel
et al. 2010), with human habitation of risky places
near glaciers rising (Kargel et al. 2011), and con- 5.2 BACKGROUND
struction of valuable infrastructure increasing in
alpine environments that never before had much Digital elevation models are digital representations
human presence (Kargel et al. 2012), accurate of the Earth’s relief and are fundamental to many
DEM data are crucial for hazard assessments. This applications. The majority of currently available
is especially true for glacier hazard-prone regions digital elevation datasets are the product of photo-
where field access is difficult due to severe terrain, grammetric software, which exploits the technique
political sensitivity, or financial constraints. of stereoscopy using overlapping image data,
Glaciological research is increasingly utilizing although increasing volumes of elevation data are
remote-sensing studies and topographic informa- also derived from radar interferometry and from
tion generated from airborne and satellite imaging, laser altimetry. Additional elevation datasets are
radar and LiDAR (light detection and ranging) available from digitized map sheet topography
systems, and GNSS (global navigation satellite sys- and, although usually spatially limited, ground sur-
tem) surveying. Furthermore, DEMs are required veys. As with many remote-sensing exercises, the
for image orthorectification and radiometric cali- generation of digital elevation data represents a
bration (Bishop et al. 2004), surface energy balance complex process involving image acquisition, com-
studies (Arnold et al. 2006a), debris-covered glacier putation and modeling, data management, and
mapping (Paul et al. 2004b), glacier ice volume analysis. The terrain itself is spatially complex,
loss and mass balance estimates (Berthier et al. and the need to represent it digitally and describe
2007), and equilibrium line altitude estimation it quantitatively has led to various mathematical
(Furbish and Andrews 1984, Leonard and Fountain approaches for different scale and data constraints.
2003). Examples include the analysis of systematic high
The mere availability of satellite imagery and or low-frequency periodicity in altitude using
DEMs, however, does not necessarily equate to Fourier analysis, and the identification of self-
accurate thematic information production and similar, spatially autocorrelated, and scale-
glacier parameter estimation (Raup et al. 2007). dependent topography using fractal geometry and
Consequently, researchers must account for a semivariograms.
variety of issues in digital terrain modeling and Digital elevation data are normally represented in
geomorphometric analysis. Digital terrain modeling one of four ways, including use of: (1) a regular
is a complex process that focuses on terrain surface grid; (2) a triangulated irregular network (TIN);
representation schemes, acquisition of source data, (3) contour lines; or (4) point profiles. Regularly
terrain descriptors and sampling strategies, inter- gridded DEMs comprise square pixels of constant
polation techniques and surface modeling, quality spatial resolution (the effective ground dimension)
control and accuracy assessment, data manage- or constant latitude/longitude resolution, with each
ment, and interpretation and applications (Wilson pixel assigned an elevation value, usually in meters
and Gallant 2000, Li et al. 2005, Fisher and Tate and with respect to some datum (sea level or a
2006). Accurate glacier assessment using topo- higher order representation of the geoid). They
graphic information is frequently an issue of are favored by geoscientists because they are
DEM quality and methodological approach (Raup directly comparable with remote-sensing imagery
et al. 2007). of equal spatial resolution, simple to analyze statis-
Background 115

tically, computationally easy to represent, and can include surface cover or not, and are therefore gen-
be saved in a range of formats (e.g., GeoTIFF, erally referred to simply as DEMs.
HDF). Conversely, they are poor at representing The use of elevation data derived from remotely
abrupt changes in elevation and heterogeneous sensed imagery has a number of advantages over
topography, particularly at coarse resolutions, can alternative sources. First, the range of sensors avail-
generate large file sizes at fine resolution, and may able for producing DEMs offers flexibility in scale.
have a large amount of data redundancy across flat Many medium-resolution (10–30 m) satellite sen-
areas. TINs represent topography using a mesh of sors (e.g., ASTER, ERS-1/2) are able to provide
adjacent, nonoverlapping triangles with each vertex stereoscopic imagery for DEMs of wide areal cover-
assigned an elevation value. The TIN model stores age, while local to catchment-scale DEMs can be
topological information allowing proximity and derived from an increasing number of fine resolu-
adjacency analyses across a range of spatial scales. tion (1–10 m) sensors (e.g., SPOT 5 HRS, Quick-
TINs usually require fewer data points than a DEM bird, GeoEye) or aerial imagery. Normally, the
grid and are particularly useful for accurately repre- scale of the topographic variability to be modeled
senting extreme topographic features (e.g., channels informs the selection of the source imagery (see
and ridges) as well as for surface modeling (e.g., the Section 5.3.1). Second, many glacierized areas of
calculation of slope, aspect, surface area and length, the world are inaccessible for logistical or political
etc.). Contour maps, some digitized from paper reasons, so for the inclusion of three-dimensional
topographic maps, provide more generalized repre- data in global databases such as GLIMS, remotely
sentations of the terrain, but they are not so com- sensed elevation information provides the only
monly used in glacial applications given the option. Third, as historical archives of these data
availability of contemporary DEMs for the entire are accumulating, the opportunity to conduct
globe; however, for the derivation of ice thickness multitemporal (and thus dynamical) studies
changes over multiple decades, such digitized maps becomes possible. Fourth, and finally, the increas-
can be crucial. Lastly, point profile data normally ing availability of no-cost data derived from
represent elevations collected by laser altimetry remotely sensed sources opens up the discipline of
(e.g., GLAS on board the ICESat), which illumi- cryospheric remote sensing to all interested parties,
nate a circular sampling area (footprint), the size which can only ultimately advance scientific under-
and spacing of which is generally dependent on the standing at a greater pace.
altitude of the sensor. These data therefore tend to In relation to this last point, however, it is impor-
provide sparse coverage when acquired by satellite, tant that standards and procedures for DEM
and repeat pass data are not necessarily spatially generation and analysis are available for scientists
coincident. worldwide, an issue that is discussed later in this
The terms DEM, DSM (digital surface model), chapter (p. 137). In brief, the utility of a DEM is
and DTM (digital terrain model), are often used dependent upon a certain level of expertise from the
interchangeably to describe any digital representa- operator to ensure the correct scale of representa-
tion of the Earth’s topography, but differ funda- tion, accurate characterization of terrain param-
mentally in their inclusion or exclusion of surface eters, and overall planimetric and vertical
cover in the data. DSMs can be thought of as accuracy. For example, a major difficulty in using
including all elevation data recorded regardless of digital photogrammetry to generate elevation data
surficial cover; thus they represent tree canopy over glacier surfaces is that accurate DEMs can
height across a forested area and building heights only be derived if the glacier surface shows sufficient
in the urban environment, for example. DTMs topographic and/or radiometric heterogeneity to
‘‘strip out’’ such surface features and attempt to correlate image pairs (Favey et al. 1999), which
represent only the solid Earth, which can be a chal- otherwise can result in grossly inaccurate elevation
lenging task (sometimes requiring large amounts of estimates over some large expanses of bare ice and/
interpolation) where dense ground cover(s) exist. or snow. Image saturation or low dynamic numbers
The term DEM is used more generally to describe in deeply shadowed areas also result in either null
an elevation dataset without specifying the eleva- values or erroneous elevations. Consequently, great
tion reference (e.g., solid Earth, surface cover, care is required to ensure that the data selected for
subsurface topography). Many satellite-derived ele- an application are appropriate, processing is carried
vation datasets are of insufficient vertical resolution out with a high level of expertise, and errors in any
to require an explicit declaration of whether they derived data are accurately reported, so that real
116 Digital terrain modeling and glacier topographic characterization

geophysical patterns and features can be differen- etries at the time of acquisition. Numerous investi-
tiated from image and processing artifacts. gators have reported on the generation of DEMs
from aerial photography and a multitude of
satellite-imaging systems including ASTER, IRS,
and SPOT for glaciological applications (Kääb
5.3 DIGITAL ELEVATION 2002, Berthier and Toutin 2008). Although satellite
MODEL GENERATION imagery is generally available, aerial photography is
more difficult to acquire for many glacierized
5.3.1 Source data regions of the world (e.g., Karakoram) because of
Topographic maps exist for every country, and political sensitivities and poor archiving, and some
these can be utilized to generate a DEM, provided air photos are expensive (e.g., Alaska), which can
the map is of sufficient quality for the intended limit historical analyses. However, the demilitariza-
application. In developing countries, issues of tion of Corona satellite imagery provides historical
cartographic scale, map coverage, quality, and data for many regions of the world at relatively fine
availability may be problematic. In some countries (<10 m) spatial resolution; Corona also provides
the maps are military secrets, or have been degraded stereo coverage facilitating photogrammetric
in quality given the scale, contour interval, or intro- analyses (Surazakov et al. 2007, Bolch et al. 2008).
duction of error (e.g., Soviet topographic 1:50,000 For cryospheric applications, and especially
map series of Afghanistan). If suitable maps exist useful for three-dimensional information, ASTER
DEMs can be generated by cartographic digitiza- is one of the most appropriate and accessible sen-
tion using techniques such as line-following digiti- sors available for several reasons. First, and most
zation or raster scanning. importantly, the coupled nadir and backward-
High-resolution aerial and satellite images are looking sensor systems in the near-infrared enables
also effective ways to generate and update topo- fine-resolution and routine along-track stereoscopic
graphic information. Digital elevation models can vision (Fig. 5.1). Second, the spatial resolution of
be derived using stereo (overlapping) pairs and these stereoscopic data (15 m) provides the perfect
photogrammetric techniques, which depend on balance between image detail and wide swath width
knowledge of the exact image and terrain geom- (60 km) that is required for regional-scale glacier

Figure 5.1. Imaging geometry of the ASTER sensor, on board the Terra satellite, launched in December 1999. The
satellite is in a 705 km, Sun-synchronous orbit that results in a 16-day revisit period for any given location on Earth.
The ASTER sensor is equipped with two bands in the near-infrared: one at nadir (3N) and one backward looking
(3B). It takes the sensor approximately one minute to cover the same area of ground with both sensors, thus yielding
a stereo pair of 60 km swath from which DEMs can be extracted (adapted from Hirano et al., 2003).
Digital elevation model generation 117

studies. Third, the adjustable sensor gain settings interact with surface materials to regulate the inten-
provide increased contrast over bare ice and snow, sity of the returning signal. A LiDAR system pro-
which is critical for the derivation of accurate duces data that may be characterized as a 3D point
elevation data using photogrammetric techniques. cloud, which must be processed to remove erron-
Fourth, the relatively short revisit period of the eous measurements relating to reflectance from
sensor (16 days for nadir viewing or 2 days for clouds, smoke, or birds, for example (Arnold et
off-nadir pointing) provides high-intensity coverage al. 2006b). Consequently, filtering, classification,
in cases of natural disaster emergencies or other and modeling are required to generate a DEM
special needs; in more routine cases, the short revisit from laser-based measurements. The high quality
time provides many chances to acquire a cloud-free of LiDAR-based DEMs permits improved geo-
image of an area of interest, even in notoriously morphometric characterization of the terrain and
cloudy areas. ASTER data are routinely used to glacier surfaces, which translates into improved
provide DEMs for the derivation of three- assessment and modeling of a variety of
dimensional glacier parameters for all glacierized processes.
regions, and are available at no cost to regional Finally, Global Navigation Satellite Systems
centers involved in the GLIMS project. However, (GNSS) can be used for direct measurement of
the actual performance of ASTER has resulted in the Earth’s surface topography, and are increas-
far less frequent data acquisitions and often sub- ingly replacing the use of traditional theodolites
optimal sensor gain settings than the observation and total stations. Earth scientists are routinely
plan calls for, thus reducing the system’s value from utilizing GPS receivers on a variety of platforms
what it could be. to collect point and profile measurements for ter-
Images acquired by synthetic aperture radar rain and glacier surface representation (Miller and
(SAR) contain information relating to both the Pelto 2003). Differential GPS (dGPS) refers to cal-
magnitude and phase of the backscatter signal, both culation of an accurate roving receiver position with
of which are useful for deriving DEMs. Radargram- respect to a reference station and can yield accura-
metry exploits the magnitude element of the return cies as fine as 10 mm when the differential phase of
signal. Similar to photogrammetry, radargram- multiple satellite signals is exploited. This method is
metry makes use of two spatially separated back- particularly relevant for the collection of ground
scatter images to form a stereo model of the terrain control data as input to DEM generation.
(Toutin et al. 2013), with the exact image geometry
being supplied by increasingly accurate orbital data
records (Scharroo and Visser 1998). Interferometric
5.3.2 Aerial and satellite
SAR (InSAR) makes use of the phase element of
image stereoscopy
radar return. The phase difference between two
independent return signals is very sensitive to topo- The process of deriving elevation data by photo-
graphic variations (as well as any surface displace- grammetric techniques is well established and
ments between image acquisitions) and can be widely documented (Kääb et al. 1997, Kääb and
‘‘unwrapped’’ to derive a digital terrain surface, Funk 1999). Deriving elevation data from stereo
often with very high accuracy. DEMs over glacier- photographs relies on recreating the geometry
ized terrain have been successfully derived using a between sensor and terrain at the exact time of
range of SAR image sources including ERS-1/2 image acquisition. The user supplies approximate
(Eldhuset et al. 2003), Radarsat (Peng et al. data relating to the aircraft location and the average
2005), and sensors on board fixed wing aircraft terrain height above sea level in addition to accurate
(Muskett et al. 2003). The preceding chapter by statistics on the camera geometry and image dimen-
Kääb et al. has a detailed treatment of various sions. The software applies least-squares regression
methodologies and glacier applications using radar to reduce errors in the approximations and arrives
data. at a stereo model that realistically represents the
Airborne laser scanning systems are an important real-life situation at the time of image acquisition.
operational tool for generating high-resolution The least-squares adjustment algorithm is used to
topographic information. Numerous researchers estimate the unknown parameters associated with a
have utilized and evaluated DEMs generated from solution while also minimizing error within the
LiDAR data for glaciological applications (Rees solution. This is regarded as one of the most accu-
and Arnold 2007). Pulses of electromagnetic energy rate techniques to:
118 Digital terrain modeling and glacier topographic characterization

Figure 5.2. A typical stereo photogrammetric model (or block), illustrating the monotemporal generation of height
information from overlapping images (imagery of time 1 only). Where multitemporal imagery is available (imagery of
time 1 and imagery of time 2), calculations of surface elevation change become possible. By processing multi-
temporal images in a single block, potential errors, particularly those relating to external ground control, can be
minimized.

. estimate or adjust the values associated with resultant orthoimages and elevation information to
exterior orientation (sensor positioning and be considered accurate when compared with the
view direction at time of image acquisition); actual terrain properties.
. estimate the X, Y, and Z coordinates associated Once the error is within thresholds considered
with tie points; satisfactory by the user, the DEM is extracted by
. estimate or adjust the values associated with inter- the generation of spatial rays from the first image
ior orientation (sensor geometry, including focal that intersect in three dimensions with the corre-
length); and sponding points on the second image (identified
. minimize and distribute data error—introduced by image-matching algorithms), thus giving a
by ground control point (GCP) coordinates, tie height value for each modeled pixel (Fig. 5.2). Ele-
point locations, and camera information— vation is determined by measuring shifts in the x-
through the network of observations direction (x parallax) of the rectified images. The
(list modified from Leica Helava 2004). exact process followed in DEM generation may
differ slightly depending on the software employed
The least-squares algorithm processes the data (see Section 5.3.4) but typically follows a series of
iteratively until the errors associated with the input key stages (Fig. 5.3).
data are minimized. These errors are measured in
terms of residuals (the degree to which the observa-
5.3.3 Ground control points
tions deviate from the functional model). Residual
values (the distance between calculated statistics Ground control data may be derived by traditional
and user input values) are summarized by the root surveying techniques (such as triangulation, tri-
mean square error (RMSE), which should ideally be lateration, traversing, and leveling), contemporary
less than twice the pixel size of the imagery for the survey techniques (such as dGPS measurement and
Digital elevation model generation 119

Figure 5.3. The major steps required for the extraction of DEMs from satellite imagery. Exact processing sequences
vary according to the software used, with some offering total user control over all sensor/model parameters and
processing algorithms (white box), and others offering no control over internal implementation (black box). In both
cases, users should be careful to understand associated uncertainties in derived elevation data.

laser ranging), and/or the identification of clearly post-derivation accuracy tests (Toutin 2002). On
defined features (e.g., mountain peaks and road the other hand, where GCPs are known to have
intersections) on large-scale map sheets. They low accuracy it is beneficial to limit the number used
may give information on horizontal positioning, in the adjustment to avoid errors propagating
height, or both. Collectively, ground control data through to the output DEM.
inform the stereo model of the spatial position and In addition to the quality of any ground control
orientation of a photograph or satellite image rela- data used in a stereo model, the theoretical accuracy
tive to the ground at the time of exposure (or scan- of a DEM is also dependent on the base-to-height
ning), and their quality is important to the accuracy ratio of the system and the quality of the image
of derived elevation data. As a minimum, the accu- matching used to refine the stereo model. If no
racy and precision of any ground control data ground control data are available, only relative
should be twice that of the expected DEM to avoid DEMs may be derived (where horizontal and ver-
adversely affecting the derived data. Research has tical values are only known relative to an arbitrary
shown that the accuracy, number, and distribution datum). Their accuracy is thus dependent only on
of ground control points used in the iterative least- sensor characteristics and the success of image
squares adjustment to refine the geometric model matching. Further parameters that can influence
can also impact DEM accuracy significantly the accuracy of any derived elevation data include
(Toutin 2008). While three GCPs are theoretically the quality of geometric and radiometric image cali-
sufficient to compute a stereo model from frame brations (which also impact matching success and
imagery (Gonçalves and Oliveira 2004), the mini- ground control point identification) and the quality
mum number for other sensors depends on the of sensor ephemeris and attitude data that is entered
sensor model employed and the number of into the stereo model. Check points, which are most
unknowns in it. A larger number than the mini- often surplus ground control data, can be used to
mally required is usually employed to ensure redun- quantify DEM errors but, similarly, the quality of
dancy in least-squares adjustment, to reduce the accuracy assessment is only as good as that of the
impact of map and plotting errors, and to facilitate checkpoints used.
120 Digital terrain modeling and glacier topographic characterization

from ASTER Level 1A and 1B stereo pairs with


5.3.4 Software packages
the aid of orbit/sensor modeling, quasi-epipolar
In addition to the software used by NASA and the image generation, cross-correlation matching, and
Japanese Aerospace Exploration Agency (JAXA) automatic detection of water bodies. An existing
for the operational generation of ASTER DEMs, DEM can be added to fill in low-correlation areas.
there are four university or private software The claimed accuracy by ENVI for a relative DEM
packages, as well as five commercial off-the-shelf without GCPs is better than 20 m and for an
(COTS) software modules for processing stereo absolute DEM with GCPs 30 m in planimetry
ASTER data to generate DEMs (Toutin 2008). and 15 m in elevation (all with 90% confidence
We now give a brief description of some of the main interval). These results were confirmed by end users
software packages available. generating DEMs from Level 1A stereo pairs both
The Geomatica 2 OrthoEngine SE of PCI Geo- with and without GCPs, and from Level 1B stereo
matics (www.pcigeomatics.com), adapted to ASTER pairs without GCPs.
stereo data, was developed in 1999 under USGS The Desktop Mapping System Softcopy 2 ,
contract with the collaboration of the Canada Cen- Version 5.0 is designed to run under the Microsoft 1
tre for Remote Sensing, Natural Resources Canada, Windows 1 XP Professional operating system and
for mathematical, 3D modeling, and algorithmic provides a complete two-dimensional and three-
aspects. It may be the most-used software for dimensional photogrammetric mapping capability
performing 3D sensor orientation and DEM and using scanned aerial photographs or digital
orthoimage generation. Both relative DEMs with- images recorded by airborne or satellite sensor sys-
out GCPs and absolute DEMs with GCPs can be tems. Experiments with stereo ASTER data over
generated while an existing DEM can be added in different mountainous study sites achieved RMS
the processing. The accuracies obtained in different errors of about 15–25 m (1), showing equivalent
research studies were similar to those obtained at results to those derived by USGS (see PCI).
USGS: horizontal and vertical accuracy of 15 m DMS Softcopy 5.0 retains the speed, functionality,
and 20 m (1) respectively, depending on GCPs, and ease of use of earlier versions of the DMS
study site, and relief. software, but has been streamlined to allow for
The LPS photogrammetric software suite from even faster, more efficient implementation of
ERDAS (http://www.erdas.com) includes ASTER- operations.
specific data import, radiometric correction, and SilcAst, produced by Sensor Information
sensor model functions. It uses a feature-based, Laboratory Corp. Japan (www.silc.co.jp) and exclu-
automatic terrain-adaptive, hierarchical image- sively developed for ASTER, is written in IDLR6.1
matching scheme for automatic DEM extraction and can be executed with IDL VM without an IDL
and can output grid, TIN, and point cloud data. license. The main functions are digital elevation
Of the software described here, LPS provides the extraction either from Level 1A and Level 1B
most comprehensive control over the DEM extrac- imagery, water body identification, orthorectifica-
tion and editing process. A number of automatic tion, and Level 1B data generation, but the software
spike and blunder detection algorithms aid the does not accept GCPs. ASTER 3D data products
detection of local extraction errors, and automatic are provided as Standard (fully automatic) with
blending and smoothing around the perimeter of potential miscalculated elevations, and High
area edits leads to seamless data continuation (semi-automatic) with interactively corrected mis-
through interpolated or replaced values. The most calculated elevations. In 2009, SILC with SilcAst
promising of these algorithms fuses independent produced a high-quality global DEM (G-DEM1;
datasets to achieve a complete DEM; in one see Section 5.3.6) from ASTER data acquired
approach, Crippen (2010 and unpublished results) within an 83 latitudinal limit. Covering all the land
used parallax–elevation cross-correlation as a on Earth, it is freely available to all users in 1  1
means of detecting artifacts in the ASTER DEM grid tiles in latitude and longitude. While the
so as to determine where replacement with other ASTER G-DEM1 contains 30 m grid spacing with
data (e.g., from SRTM) should be done. 7 m accuracy (1), recent scientific studies demon-
The AsterDTM module, developed by SulSoft, strated some high-frequency errors (appearing
the exclusive ENVI distributor in Brazil, was added within tile grids as pits, bumps, ‘‘mole runs’’, and
to ENVI as an add-on module in 2004 (http:// other residuals and artifacts), requiring down-
www.ittvis.com). The software can create DEMs sampling to 90 m spacing.
Digital elevation model generation 121

5.3.5 Postprocessing (interpolation and influence on DEM quality. For example, a strong
smoothing) weighting based on only few sample locations pro-
duces a rough surface in which the sample points
DEM generation often requires the interpolation of
themselves are heavily emphasized.
elevation values between available sample points,
particularly when contour maps or satellite imagery
are the source data. Consequently, various ap- Spline: this interpolation method applies a special
proaches for how to interpolate this calculated function that is defined step by step by piecewise
(not measured) elevation in the DEM production polynomials that are continuously differentiable.
process have been developed. The choice of Most common in DEM generation are low-order
approach depends mainly on relief within the cubic (r ¼ 3) and bicubic (r ¼ 2) splines. In general,
imagery, the type of information that will be the surface of the final DEM is smooth and even.
derived from the DEM, and the proposed applica- However, the main problem is the occurrence of
tion (Rasemann et al. 2004). Interpolation accuracy overshoots (i.e., unnatural ‘‘holes’’ and ‘‘spikes’’).
depends on interpolation algorithm type, surface To prevent such overshoots, a regularized spline-
characteristics, and the distance between sample with-tension algorithm is used (Mitasova and Mitas
points (Kubik and Botman 1976). There is a paucity 1993). This method uses a function that specifies
of quantitative evaluation of interpolation accuracy how fast the collection of sample points decreases
over glacierized terrain and, of the few studies that with increasing distance thereby attenuating the
do exist, only a handful have employed ground magnitude of overshoots.
data to make the assessment (Cogley and Jung-
Rothenhausler 2004). At present, uncertainties in Kriging: a geostatistical method, based on a
DEM accuracy do exist depending on the choice stochastic model, which predicts the elevation at a
of the interpolation algorithm employed; therefore, specific location and is particularly used when a
all qualitative and, in particular, quantitative results trend (here: the spatial correlation of elevation
from glaciological analyses that include DEMs values) exists. One disadvantage of this method is
must be interpreted with caution. that adjustment of the semivariogram model to the
Some of the interpolation methods that have existing data structure requires experience.
been used in glacier research include (but are not
limited to) Triangular Irregular Network (TIN;
Various studies have considered the relative
Gratton et al. 1990); Inverse Distance Weighted
accuracy and appropriateness of each of the above
(IDW; Etzelmüller and Björnsson 2000); spline
approaches. For example, Kamp et al. (2005) tested
(Mennis and Fountain 2001, Bolch et al. 2005);
three interpolation methods—IDW, spline, and
and kriging (Burrough and McDonnell 1998).
TIN—when generating a DEM from 1:50,000 topo-
graphical maps of Cerro Sillajhuay in the Andes of
Triangulation: since this method produces indi- Bolivia and Chile using Esri ArcGIS 3.2 software.
vidual triangles with constant exposition and aspect In this study, both IDW and spline methods pro-
rather than a continuous relief, the final DEM is not duced very good results in high relief areas, but a
differentiable. To ensure that information on cur- large number of artifacts in generally flat terrain.
vature can be extracted, the DEM has to be con- The authors concluded that both algorithms had
verted to a grid. Problems occur at sharp edges, problems handling larger distances between con-
deep valleys, and cliffs. Often, valleys and ridges tour lines. Instead, the most accurate DEM was
show a typical ‘‘terracing’’ effect. Such artifacts generated using the TIN method, although a first
can be eliminated by manually adding elevation raw DEM contained many artifacts—mainly over
information, by using structural information with flat terrain, caused by the triangles that are used by
elevation values (Heitzinger and Kager 1999, the TIN method for interpolation.
Rickenbacher 1998), or by subdividing the triangles Tests similar to those made by Kamp et al. (2005)
(Schneider 1998). were described by Racoviteanu et al. (2007) for the
Cordillera Ampato (southern Peru). In this study,
Inverse Distance Weighted: a relatively simple algo- the examination of RMSEz (vertical error) values
rithm that weights the sample point value according for DEMs derived from topographic data (1:50,000
to its distance from the pixel of interest. The weight- maps constructed from 1955 aerial photography)
ing and number of sample points have a strong revealed that no interpolation method performed
122 Digital terrain modeling and glacier topographic characterization

perfectly. While the IDW DEM and TIN DEM combine, or fuse, DEMs from different sources in
showed accuracies of only 21 m and 24 m, respec- order to obtain the most complete coverage of an
tively, the spline (here: TOPOGRID) DEM pro- area. As well as filling in any gaps in the respective
duced a vertical accuracy of 15 m, DEMs, such fusion can also help to identify any
outperforming the other available algorithms. severe vertical and horizontal errors in common
However, the authors noted that all three DEMs areas (Kääb et al. 2005). DEM-merging techniques
produced ‘‘terracing’’ effects, an artifact of prefer- range from replacement of data (Kääb 2005a) and
ential sampling along the contour lines, with points weighted fusion, resulting in smooth transitions
closer to the contour lines being interpolated using between DEMs (Weidmann 2004), to merging in
the same elevation values. The terracing effect was support of DEM processing itself (e.g., DEM
most severe using the IDW interpolator. Similar approximation in order to geometrically constrain
terracing artifacts were described by Etzelmüller stereo parallax matching or interferometric phase
and Björnsson (2000) using the same interpolator, unwrapping) (Honikel 2002). One successful
and by Mennis and Fountain (2001) using the approach involves identification of artifacts in a
spline-with-tension interpolator. Wilson and Gallant single-scene ASTER DEM and then infilling with
(2000) showed that such terracing artifacts affect a global dataset such as SRTM or the ASTER G-
calculations of topographic characteristics such as DEM. Artifact delineation is almost perfect (almost
slope, aspect, and profile curvature. no errors of omission or commission), so the infilled
DEMs constructed over largely featureless ter- product is usually of very high quality (Crippen
rain often fail to characterize subtle (or even 2010).
abrupt) changes in surface topography, which can The ASTER G-DEM1 is the result of the fusion
be rectified by the use of breaklines. Breaklines are of suitable individual ASTER-derived DEMs by
linear features that describe interruptions in surface averaging all elevations in the vertical stack. While
smoothness and are typically used to define chan- DEM fusion using this approach is helpful for
nels or ridges, for example, where there is a clear deriving a spatially more complete and accurate
change in surface topography. However, the devel- set of geomorphometric parameters, it should be
opment of real 3D breaklines is still a problem, and avoided or handled with care for DEM-based stu-
Bolch and Schröder (2001) actually concluded that dies of glacier thickness changes. The number of
the integration of simple breaklines does not sig- stereo images used to produce a given G-DEM1
nificantly improve the overall quality of the DEM. grid point elevation varies from one to several tens.
As a solution, in their DEM generation process Whereas the number of scenes used for each grid
Kamp et al. (2005) manually added some contour point is known globally, G-DEM1 ancillary data do
lines using elevation information from stereoscopic not support knowledge of when the images were
analysis of aerial photographs. Eventually, the TIN acquired. The mean date is about 2004/2005, but
DEM was converted into a grid-based DEM using for a given locale it could be earlier or later by a few
Esri ArcInfo 7.2 software, and two nearest neighbor years. Variance of the mean from 2004/2005 is less
filters helped smooth the final model, albeit at the than a year in most areas, but it can be much greater
expense of some topographic detail (e.g., the loss of for some places where few ASTER images have
some edges, frost cliffs, and gorges). The authors been acquired. G-DEM1 carries some of the arti-
concluded that the final DEM was of sufficient facts present in individual DEMs, and is generally
quality for macro and mesoscale geomorphological very noisy, often rendering it unsuitable for detailed
analysis (e.g., of rock glaciers); however, processing studies requiring high resolution, although for
the DEM from contour lines was relatively time studies of basin-scale hypsometry, G-DEM1 can
consuming. be convenient. A new G-DEM2, of similar con-
struction but lacking the severity of noise and arti-
facts of the earlier version, was released in October
2011. G-DEM2 benefits from 260,000 additional
ASTER scenes, and preliminary tests indicate
5.3.6 Data fusion
improved accuracy in both horizontal and vertical
While small gaps in DEM data can be filled by dimensions, and a substantial reduction in the num-
interpolation, large voids stemming from shadow- ber of voids (Tachikawa et al. 2011). Still remaining
ing, correlation failures, and other effects require a for the future is development of a new GDEM that
different approach. It is sometimes desirable to would allow better tracking of input images.
DEM error and uncertainty 123

5.4 DEM ERROR AND UNCERTAINTY spatial structure of error in DEMs remains poorly
understood (Carlisle 2005). Numerous researchers
5.4.1 Representation of DEM error have attempted to study and describe the nature of
and uncertainty spatial error and have found that patterns of error
can be anisotropic, scale dependent, spatially vari-
Given the number of stages involved in generating a
able, and spatially correlated, but most importantly
DEM, errors are inevitable in both the vertical and
DEM error is closely related to the complexity and
planimetric coordinates. In general, error can be
characteristics of the terrain (Kyriakidis et al. 1999).
classified as: (1) gross errors that are associated with
Work in alpine environments often reveals patterns
equipment failure; (2) systematic error that repre-
of error over mountain terrain and glacier surfaces
sents the results of a deterministic system that can
that are highly variable in spatial extent and loca-
be accounted for by using functional relationships;
tion, and are related to acquisition of source data,
and (3) random errors that occur within (or because
spatial interpolation, and terrain geometry.
of ) the multitude of operational tasks that are
involved in computation. In this context, error is
typically defined as systematic and/or random var- 5.4.2 Type and origin of errors
iations about a ‘‘true’’ reference value (Fisher and
Once a DEM has been generated by determining
Tate 2006). The magnitude of vertical error can be
the parallax between pixels matched by cross-
an important consideration, as changes in glacier
correlation there is always a need for postprocessing
surface altitude over shorter time frames may
due to mistakes in pixel matching, correlation fail-
become undetectable, depending upon the nature
ures, and, if present, the impact of clouds in the
of the source data. Consequently, the usefulness
scenes. Further, shadowing in regions of high relief,
of a DEM for estimating mass balance must be
low image contrast over clean snow and ice, and
carefully evaluated (Bamber and Payne 2004).
self-similar surface features on glaciers will lead to
A common descriptor of error is root mean
voids and artifacts in the resulting DEM. Some
square error (RMSE):
sP ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi software packages, such as SilcAst, will try to
P 2 remove such features automatically, with no user
i j ðzi; j  zi; j;r Þ
RMSE ¼ ð5:1Þ control over the process. Alternatively, ENVI, LPS,
n and PCI Orthoengine all provide a suite of post-
where zi; j is the elevation at location (i; j ) in the processing tools for manual editing. However,
DEM, and zi; j;r is a reference elevation at the same Eckert et al. (2005) found that in some cases the
location. The metric has been used when comparing postprocessing tools provided by software packages
elevations off-glacier in a variety of glaciological were not sufficient and further morphologically
studies, although it does not describe very well based interpolation algorithms were required to
the statistical distribution of vertical error (Fisher produce a suitable DEM. Such matching mistakes
and Tate 2006). It has, however, become a standard and voids are relatively easy to identify and correct
measure of map accuracy. Other researchers have for, but where inaccurate GCPs and poorly chosen
used additional error metrics such as mean error tie points are used, distortions, biases, and plani-
(ME) and standard deviation (S) (Cuartero et al. metric shifts may also be introduced into the
2004): derived DEM, which can be difficult to model
PP without the use of a reference elevation dataset.
ðzi; j  zi; j;r Þ
ME ¼ ð5:2Þ This is particularly problematic where multi-
n temporal DEMs are required for the measurement
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
PP of elevation changes (see Section 5.5.5).
½ðzi; j  zi; j;r Þ  ME 2
S¼ ð5:3Þ Two types of error affect the measurement of
n1
glacier elevation changes: horizontal (planimetric)
Systematic under or overestimation (biases) are and elevation (altimetric). Planimetric error refers
depicted as positive or negative ME values. to horizontal shift of often correct elevation values
Such global summary (single-value) metrics are to erroneous easting/northing locations. This can
quick to calculate and easy to report, but they fail to be either a systematic or a variable shift, and can
characterize the spatial pattern of error, which is be exhibited in one or more directions. Such shifts
critical to environmental applications. DEM errors are easily quantified by correlating corresponding
can vary spatially, and research indicates that the orthoimages and quantifying the orthophoto paral-
124 Digital terrain modeling and glacier topographic characterization

prior to analysis. Particularly relevant in this


respect is the widely used continuous DEM pro-
duced by the Shuttle Radar Topography Mission,
which was flown in February 2000 and provides a
topography covering continental areas from 60 N
to 56 S (Rabus et al. 2003) with a 1 and 3 arcsec
spatial resolution (about 30 and 90 m, respectively).
It has yielded significant advances in the monitoring
of glacier wastage in Alaska and neighboring
Canada (Muskett et al. 2003, Larsen et al. 2007,
Schiefer et al. 2007, Berthier et al. 2010), Patagonia
(Rignot et al. 2003, Rivera et al. 2005), and central
Asia (Surazakov and Aizen 2006). However, since
its release, significant biases in SRTM data have
Figure 5.4. Scatterplot of mean difference altitude been identified by a number of authors, and a clear
generated from a SRTM DEM (2000) and an examination of the bias still needs to be performed.
ASTER-derived DEM (2004). Mean values were com-
For example, Kääb (2005a) found a systematic
puted off-glacier over the Nanga Parbat Massif in
mean error of about 7 m for the Gruben area in
Pakistan. Variations in viewing geometry and/or space-
craft orbital parameters result in a nonlinear bias that Switzerland. Still in the Alps, Berthier et al. (2006)
significantly deviates from zero. Consequently, linear found a small mean error (<3 m) but detected a bias
and nonlinear systematic biases must be removed when that increased linearly with elevation. Such a bias
comparing DEMs generated from different sensor prod- related to elevation is still not fully understood but
ucts. has since been confirmed for Patagonia (Möller et
al. 2007), the Himalayas (Berthier et al. 2007), and
the Canadian province of British Columbia (Schie-
lax (Li et al. 1996, Georgopoulos and Skarlatos fer et al. 2007). The latter authors used a piecewise
2003), since there should be no relief displacement model to correct elevation bias (constant bias at low
in either image if the DEMs are accurate. Altimetric elevation and linear bias above a threshold alti-
error refers to systematic biases that may artificially tude). The regional pattern of the biases in SRTM
increase or decrease elevation in a given direction data is also complex. Surazakov and Aizen (2006)
across part of, or the whole, DEM, as well as more found a localized ice-free region with systematic
locally variable slope or aspect-related biases (Fig. error in the SRTM data ranging from 20 to 12
5.4). The correction of both types of error usually m on a 30 km spatial scale. Berthier and Toutin
requires some form of mathematical modeling. The (2008) found a 20  20 km region where SRTM
most promising approach has focused on the use data are systematically 5–10 m higher than SPOT5
of ordinary least-squares regression modeling to or ICESat elevation data. Note here that, although
create spatially nonstationary, spatially correlated sensor platform orbits are too widely spaced for
and heteroskedastic error surfaces from only a mountain glacier elevation change mapping, ICE-
small number of sample locations (Kyriakidis et Sat elevation profiles have been shown to be a good
al. 1999, Carlisle 2005). These error surfaces can source of data to detect and understand errors in
then be used either to create higher accuracy SRTM data for ice-free regions where no elevation
DEMs, or to inform analyses of error propagation changes are expected (Carabajal and Harding 2005,
in the computation of surface derivatives (Oksanen Berthier and Toutin 2008).
and Sarjakoski 2005). It is crucial that both plani-
metric and altimetric errors are identified, particu-
larly in the case of glacier monitoring as they can 5.5 GEOMORPHOMETRY
lead to erroneous volume changes (Schiefer et al.
2007) or even change the sign of the resulting mass Glacier geomorphometry refers to the three-
balance calculation (Berthier et al. 2006). dimensional topographic characteristics of a
Systematic errors are common in some of the glacier surface including its size, shape, hypsometry,
most widely available topographic datasets, making orientation, and position. Geomorphometric
it essential for users to make appropriate error analyses have historically pursued fine-resolution
assessments, even if they believe biases are removed digital elevation models to improve and refine land
Geomorphometry 125

surface parameters and objects (Drăgut et al. 2009). ducing such metrics using standard algorithms, with
However, with the increasing availability of detailed the quality of the derived geomorphometric data
DEMs it is becoming clear that the relationship being largely dependent on the accuracy of the
between DEM scale and the ability to extract geo- input elevation data. In an applied context, land
morphometric information is not straightforward. surface parameters are a useful means of comparing
Indeed, for many applications, increased DEM the topographic characteristics of glacierized areas
resolution only serves to introduce noise into the and, in addition, are increasingly being used in
analysis. Accordingly, a number of methods have combination with satellite imagery to derive more
been applied to account for scale-related problems robust approaches to accurate glacier delineation
through DEM generalization, although further (see Chapter 4, and Section 5.6 for further details).
work along these lines is needed if accurate topo- They are also often employed in process modeling,
graphic characterization of glacierized areas is to be with elevation providing useful information on
achieved (Drăgut et al. 2009). Despite this current gravity-driven processes. For example, land surface
imperfection, many quantitative analyses of land- parameters such as slope gradient can indicate the
scape and landscape evolution remain heavily gravitational potential available for geomorphic
dependent on the derivation of geomorphometric work (i.e., erosion), and slope aspect the direction
data in their methods. of the work carried out (Smith and Pain 2009).
It is important to note that geomorphometric
data may vary dramatically between adjacent sub-
5.5.1 Geomorphometric land
catchments and, less obviously, for the same glacier
surface parameters
or glacierized area through time. Changes in glacier
When deriving geomorphometric information from topography, for example, may be caused by a sus-
DEM data it should be noted that, for every order tained change in mass balance, or by a change in ice
of differentiation, the effective resolution of the dynamics (or both). Basic elevation data reflecting
product declines and noise increases. The first these changes may be derived by taking a topo-
derivative (based on calculations between two graphic profile along the flow line of the glacier
points) normally defines slope, whereas the second from multitemporal DEMs or by computing the
derivative (based on calculations between three altitude–distance function, which effectively com-
points) normally defines curvature. For every order pares every pixel representing the surface of the
of differentiation, small errors in the primary data- glacier to a terminus position by plotting the aver-
set become exaggerated. Thus, a DEM having a age altitude in a distance bin. More complex eleva-
30 m posting can report the first and second deriva- tion change data can be derived by hypsometric
tive values still with 30 m postings, but each grid cell analysis, essentially a frequency analysis, character-
of the derivatives includes information from neigh- izing the altitude–area distribution of a glacier
boring cells and so the effective resolution is surface. However, the generation of more complex
reduced for the derivatives, even though a value is data such as these is not always desirable; for
computed for each grid cell. Effectively this means example, small changes in surface elevation may
that derivative information for very small glaciers is be easily detected in simple multitemporal topo-
useless, especially if the derivatives, such as curva- graphic profiles, but not so in a multitemporal
ture, are used to define lateral moraines, computed hypsometric analysis where large volumes of
drainages, or other landforms or hydrological units. (potentially noisy) data can mask subtle surface
Thus, differential geomorphometric parameters changes. Whichever geomorphometric approach is
carry both great power to discriminate distinct adopted, it is essential that systematic biases in the
terrains and landforms, and great risk that noise DEM are removed prior to analysis and interpreta-
or systematic bias is transmitted into geomorpho- tion of the results (see Section 5.4).
metric mapping products.
Depending on the tolerance for noise or low
5.5.2 Scale-dependent analysis
spatial resolution, geomorphometric analysis can
emphasize nondifferential parameters (based on A range of interacting geomorphic processes shape
hypsometry), first-order differential analysis (slope, the glacial environment, leading to landforms that
slope aspect), or second-order differentials (curva- are rarely strictly spatially delimited (Schneevoigt
ture, tangential curvature) (Fig. 5.5). Most off-the- and Schrott 2006). Such landforms are diverse in
shelf image-processing software is capable of pro- size and shape, even within a single land cover
126 Digital terrain modeling and glacier topographic characterization

Figure 5.5. Geomorphometric analysis of the Batura Glacier, Hunza Valley, Pakistan: (a) SPOT HRVIR
panchromatic satellite imagery (acquired October 17, 2008), copyright CNES 2008/Distribution Spot Image.
(b) SPOT HRS-derived composite DEM (image dates 2002–2004), from which (c) slope, (d) aspect, (e) shaded
relief, and (f ) plan convexity, can all be calculated. Figure can also be viewed as Online Supplement 5.1.

classification, and microscale landforms may com- geomorphometric features (e.g., glacier extent, loca-
pose local-scale landforms, which may themselves tion and density of crevasses) based on their shape
compose catchment-scale landforms. In modeling and textural properties. One such method is to
such dynamic and diverse environments, linear measure surface roughness on multiple scales and
process-form relations and those that are spatially in multiple directions, by variogram analysis. The
restrictive are therefore unrepresentative. There is variogram is effectively an index of dissimilarity
thus increasing focus on the hierarchical organiza- between pairs of points that are separated by a
tion of topography in studies on mountain geo- predefined vector and within a specified search win-
morphology, and the subsequent adoption of dow. Multiple variograms can be used to character-
scale-dependent analyses. These can be simply ize terrain morphological features, such as debris
defined as methods that examine the variability of flow deposits, channels, and rockfall deposits, and
the three-dimensional spatial structure of morpho- to calibrate secondary morphological indices such
logical features on multiple levels. as anisotropy, periodicity, and the spatial vari-
With the increasing availability of DEMs from ability calculated at specific lags. These indices,
airborne and spaceborne platforms, automated while providing useful quantitative landscape
approaches have been developed to identify specific characterization data in their own right, can then
Geomorphometry 127

be used as input into further, perhaps more complex identifying site shelter and exposure, and to be
(e.g., pattern recognition) landscape classification applied as input to energy balance melt models,
techniques (van Asselen and Seijmonsbergen 2006). which are clearly fundamental to understanding
A further derivative of the variogram is the frac- the relationship between glacier behavior and
tal dimension (D), which provides a measure of climate.
whether topographic surfaces and spectral reflec- Net radiation (also known as net flux) is defined
tances of land cover units are self-similar over a as the balance between incoming and outgoing
range of spatial scales. Many landscape features energy and can be computed as:
and other environmental data have been shown
Qn ¼ Eð1  Þ þ L #s þ L #t þ L "
to exhibit self-similarity, at least statistically (Good-
child and Mark 1987). The limited number of where E is total surface irradiance,  is surface
studies that have evaluated the utility of fractals albedo, L #s is longwave sky radiation, L #t is long-
for these purposes indicate that the majority of wave radiation from the surrounding terrain, and
geomorphic surfaces exhibit scale-dependent fractal L " is emitted longwave radiation. It is regulated by
dimensions (i.e., limited to certain spatial ranges; numerous processes and can be highly variable
Bishop et al. 1998), with the range in scales often depending on atmospheric, topographic, and sur-
determined by the controlling geophysical processes face biophysical conditions. Indeed, in mountain-
that have shaped, or continue to shape, the land- ous areas radiative fluxes are particularly complex,
scape (Lathrop and Peterson 1992). Fractal dimen- fluctuating considerably in space and time because
sions have also been employed to improve of the effects of slope, aspect and the effective hor-
classification accuracy (de Jong and Burrough, izon, and the input of reflected and emitted radia-
1995) and to provide insight into operational-scale tion from surrounding slopes (Hock 2005). As such,
and process–structure relationships. This approach there is a movement towards the use of increasingly
appears to have great potential in this respect, fine temporal (hourly) and spatial (25 m) resolu-
although to investigate self-similarity in surface tion modeling, with the latter again dependent on
roughness characteristics over the full range of the availability of suitably fine-resolution and
spatial scales requires the use of high-quality accurate digital elevation data. Where such data
DEMs, which may not always be available in do exist, studies have shown that the melt rate of
glacierized areas. snow and ice and spatial and temporal patterns in
energy balance can be calculated accurately across a
glacier, albeit with dependence on accurate field and
5.5.3 Topographic radiation modeling
meteorological data (Brock et al. 2000).
Digital elevation models are fundamental to the
three-dimensional modeling of solar radiation var-
5.5.4 Altitude functions
iation across a landscape. Insolation modeling can
either be point specific or area based. Point-specific Most glacier parameters vary as a function of
modeling considers the geometry of surface altitude, and these relationships can give insight
orientation, the visible sky, and the effect of local into glacier dynamics and processes that simple
topography. Local topographical effects can be two-dimensional analyses would not reveal. A
determined using empirical relations, visual estima- glacier’s most fundamental altitude parameter is
tions, or, where field data are logistically possible, the equilibrium line altitude (ELA) because it
by upward-looking hemispherical photography. divides the glacier into accumulation and ablation
Such calculations can give locally very accurate areas (Braithwaite and Raper 2010). This can be
results, but are clearly limited in spatial coverage. simply estimated using multispectral satellite
Area-based insolation modeling relies solely on the imagery in combination with a DEM, by assuming
use of a DEM to calculate surface orientation and that the ELA is equal to the snow line position at
shadowing across a subcatchment, for example, and the end of the melting season (Bamber and Rivera
as a result has much greater application for under- 2007). Such information is fundamental to mass
standing large-scale landscape processes. Along balance calculations, which are useful indicators
these lines, a number of studies have made surface of climatic variations where no actual ELA data
irradiance and ablation gradient calculations to or better alternative proxies exist (Yadav et al.
identify locations most suitable for past and present 2004). In the absence of calculated mass balance
glaciation (Carr and Coleman 2007, Allen 1998) by data or transient end-of-season snow line data,
128 Digital terrain modeling and glacier topographic characterization

the topographic profile of a glacier can be used as a


relative proxy for the dynamic state of glaciers, with
those glaciers in rapid recession tending to exhibit
concave-up or flattened (planar) surface profiles in
the ablation zone, and those accumulating or main-
taining mass characterized by convex-up surface
profiles (transverse profiles especially) in the abla-
tion zone (Quincey et al. 2009).
The altitude–area profile (or the hypsometric
curve) for each glacier represents a distinct 3D
spatial signature that is the end result of thousands
or millions of years of erosion and reflects interac-
tion of the regional climate with subglacial bedrock
(Tangborn 1999). It has been used in many cases to
infer the state of geomorphic development within a Figure 5.6. Hypsometric curves for the Batura,
glacial system (Katsube and Oguchi 1999). Glacier Ghulkin, Ghulmit, and Pasu Glaciers located in the
erosion processes are determined most significantly Hunza Valley, Pakistan.
by glacier velocity, ice thickness, subsurface geol-
ogy, and basal thermal conditions. Differential
erosion therefore occurs with altitude as each of slope/altitude function) expresses glacier relief,
these parameters changes accordingly. Topographic and is thus also useful for determining erosion rates
analyses of the Nanga Parbat Massif in Pakistan and mass balance fluctuations.
revealed a spatially variable relief structure that The relationship between altitude and glacier
correlated to geomorphic events and dominant velocity is not straightforward. Glaciers with low
surface processes (Bishop et al. 2001). Glaciation relief tend to exhibit flow that is relatively uniform
was found to generate the greatest mesoscale relief across the altitudinal range, whereas those with
at high altitudes and warm-based glaciation was high relief (and are thus fed predominantly by snow
found to reduce relief at intermediate altitudes. and avalanche material) tend to have very high flow
More recently, numerical modeling has shown that rates at altitude and much reduced flow towards
mountain heights can be explained by a combina- their termini. Altitude–velocity profiles can be use-
tion of glacier erosion above the snow line and ful for proxy assessments of mass balance (Luck-
isostatic uplift caused by erosional unloading, with man et al. 2007), for identifying areas susceptible to
maximum elevations being constrained to an alti- glacial lake development (Quincey et al. 2007), and
tude window just above the snow line (Egholm et al. for determining predominant glacier flow regimes
2009). (Copland et al. 2009). In the case of the latter,
More generally, glacier hypsometry determines generally flat transverse velocity profiles with a
how responsive glacier-wide mass balance is to rapid reduction in flow at the margins indicate
climatic perturbations (Furbish and Andrews ‘‘block flow’’, suggesting down-glacier movement
1984). For example, a rise in the snow line of of ice en masse by basal sliding, whereas convex
200 m in a given region (in response to local atmo- velocity profiles that display greatest displacements
spheric warming, for example) would have a much towards the centerline of the glacier indicate a
greater impact on a low-gradient valley glacier than flow regime dominated by ice deformation. Large
a steeply sloping one, as in the former case a seasonal variability in flow can further highlight the
greater proportion of the overall ice mass would importance of meltwater availability for a given
become subject to net annual mass loss. Thus, as glacier, and may even help to identify different flow
an example, if there were a rise in local atmospheric mechanisms that exist at different altitudes within a
temperature in the Hunza Valley of the Karakoram, single glacial system (Fig. 5.7; Quincey et al. 2009).
the mass balance of the Batura Glacier would be
affected to a greater extent than that of the
5.5.5 Glacier elevation changes and
Ghulkin Glacier, and the mass balances of the Pasu
mass balance calculations
and Ghulmit Glaciers would change somewhere
between those two end members (Fig. 5.6). Simi- Stereo photogrammetry has been employed to cal-
larly, glacier clinometry (characterization of the culate glacier surface elevation changes for many
Geomorphometry 129

Figure 5.7. Selected seasonal and annual centerline velocity profiles for the Baltoro Glacier, Pakistan Karakoram.
Date format is year, month, day. Note varying flow rates across summer and winter seasons, and varying patterns
with altitude. In particular, note the similarity of flow profiles in the lowermost 400 m of the glacier, and contrasting
profiles with greater altitude, which may be related to the availability of meltwater at the glacier bed.

years, using images acquired by both airborne and Satellite-derived DEMs are less accurate than those
satellite sensors. The most reliable source of DEMs computed from aerial photographs so caution must
for volume change estimates is still aerial photos. be taken when interpreting the signal and compre-
Given their very fine resolution, if precise GCPs are hensive error analysis is mandatory.
available, DEMs with RMS accuracies the size of Changes in surface elevation can be measured
approximately one pixel (few tens of centimeters) most accurately by processing all imagery within
can be obtained (Kääb 2002). Consequently, aerial a single block or, put more simply, using a single
DEMs are still frequently employed to study vol- mathematical model (Kääb 2000; Fig. 5.2). The
ume changes at the scale of a single glacier and quality of derived surface elevation data using this
detect systematic errors in cumulative mass balance technique is mainly dependent on the interior
estimates obtained from the traditional (pit and orientation of the stereo model (Baltsavias 1999)
stake) glaciological method (Østrem and Haaken- rather than the accuracy of any ground control
sen 1999). However, aerial photographs cover a used, which makes the technique particularly
limited surface on the ground (generally one glacier) appropriate for assessing remote terrain. Factors
and it is difficult to fly a plane in remote and high- such as scanned resolution of the images, accuracy
relief areas. Consequently, they cannot be regarded of fiducial mark locations, and accuracy of camera
as a means of obtaining regular coverage of glacier calibration parameters are particularly important
topography at the regional scale. Given the large in such analyses. Model GCPs introduce a second-
footprint of satellite images (typically 60  60 km), order error that is small (of the order of a few
it is tempting to apply the geodetic method to space- percent) in relation to final vertical displacements,
borne DEMs. The principle is the same as for aerial which is consistent between the two DEMs, provid-
photography: two sequential glacier surface models ing they are bundle-adjusted as a single image
are derived from pairs of stereoscopic images and block. Therefore, the technique can, in theory, be
subtracted to yield a map of elevation changes. used without any ground control at all, while still
130 Digital terrain modeling and glacier topographic characterization

producing accurate vertical displacement results The accumulation zone is the most critical part of
from the relative image orientation alone (Kääb the glacier when mapping elevation changes. If
2002). DEMs are derived from optical images or aerial
A major source of error in vertical differences photographs, the high reflectance of the year-round
between multitemporal DEMs is their three- snow-covered accumulation zone may lead to lim-
dimensional co-registration. In cases where no com- ited features in the photographs/images or, in the
mon GCPs and tie points have been applied, severe worst case, saturation of the sensor (Kääb 2002). In
biases might exist in particular between space- the case of old maps derived from aerial photo-
derived DEMs. Errors in elevation differences from graphs, errors may have been introduced by inter-
such a lack of precise co-registration might easily polation of contours (Schiefer et al. 2007). For
exceed real elevation changes, and should therefore satellite images, featureless terrain can lead to data
be investigated as part of every DEM-differencing gaps or erroneous pits and spikes in DEMs of the
work (cf. Fig. 5.4). The most efficient method to accumulation zone. Enhanced elevation change
investigate and possibly correct DEM biases is errors may also affect the accumulation zone
the cosine method that relates elevation differences because the ice-free regions, which are used for
found for stable terrain to the combination of ter- detecting and modelling any elevation biases, are
rain slope and aspect (Kääb 2005a, Nuth and Kääb located at lower altitude. Only a few nunataks
2011). may be visible to check the presence of relative
While a 10–20 m elevation error may be accept- elevation biases at higher altitude, but, in practice,
able for the orthorectification of images, it is not even this is difficult because they are generally steep
for glacier elevation change mapping. Typically, a and small features. Consequently, any elevation
glaciologist aims at measuring the rate of change in biases need to be extrapolated from the lowest to
surface elevation within 0.5 m/yr (this is the upper the highest elevations. Other reasons for increased
bound for the error). Larger errors are acceptable uncertainties in the accumulation zone relate to
only when massive (tens of meters), but unusual difficulty in choosing density values to convert
elevation changes are monitored (Stearns and volume-to-mass changes and, when DEMs derived
Hamilton 2007). A glaciologically relevant accuracy from SAR images are used, the necessity to take
can be obtained by (1) increasing the time period into account penetration of the radar signal
between the two datasets and/or (2) improving the (depending on its frequency) into the snow/firn
quality and limiting the biases of the DEMs. The (Rignot et al. 2001). Consequently, good practice
first strategy is generally explored by comparing is to compute separate estimates of glacier volume
recent spaceborne DEMs (ASTER, SRTM, SPOT) changes for the ablation and accumulation zones,
with maps from the 1950s, 1960s, or 1970s. Apply- even if only a rough estimate of the position of the
ing such historic topographic datasets with more equilibrium line altitude (ELA) is available. Differ-
recently derived DEMs may permit determination ent uncertainties can thus be attributed to each of
of glacier elevation changes across several time- the two regions.
steps. For example, Kargel et al. in this book’s A number of strategies exist to actually compute
Figure 15.4F show a technique that captures three glacier volume changes from a set of glacier eleva-
sequential DEM elevation-differencing analyses tion changes (Kääb 2008). (1) If the two (or more)
(over a 40-year total period) into one RGB compo- DEMs to be differenced are complete, the summa-
site image. Each color channel (RGB) represents tion of vertical differences gives the volume differ-
the earliest, middle, and the latest elevation-differ- ence. Since this is often not the case (see above),
encing analysis, respectively; the composited image either (2) the hypsographic method can be used
provides immediate visual perception of the accel- where mean thickness change for each elevation
eration or deceleration of elevation changes coin- bin is multiplied by its area and all bins summed,
cident with the color-scaled time period. In this or (3) polynomial variation of elevation changes
way, spatial, dynamic, and multitemporal informa- with height can be fitted to measured elevation
tion on elevation changes are uniquely viewed differences and the elevation differences for each
within one image. The second strategy relies on glacier point then computed according to its height
higher resolution satellite images with stereo cap- using the polynomial parameters. In cases where
abilities such as SPOT5 HRS or ALOS PRISM. one of the multitemporal elevation datasets is not
Even in the latter case a minimal image time span a grid, or elevation information is distributed
of about 3 to 5 years is required. unevenly, such as in contour lines or altimetry
Glacier mapping 131

footprints, calculations must be made carefully. For isostatic rebound or tectonic movement) although
example, it might be advantageous in terms of mini- in reality this is normally small compared with
mizing error propagation not to interpolate point glacier surface elevation change. The ultimate result
elevation data to a DEM and then difference two is mass balance quantification expressed in meters
datasets, but rather to first calculate differences water equivalent (MWE) over the imaged time
between the elevation points and the DEM, and period.
then to interpolate elevation differences across the A number of studies have used this approach to
wider glacier surface. In a number of cases, eleva- assess the mass balance of alpine glaciers, using
tion differences will show less spatial variability combinations of historical topographic maps (Riv-
than elevation itself, and interpolation routines will era et al. 2007) and DEMs derived from SPOT
thus produce less serious artifacts. imagery (Berthier et al. 2007), SRTM (Racoviteanu
Data from nonoptical sensors can also provide et al. 2007), ASTER (Khalsa et al. 2004), Corona
accurate measurement of glacier surface eleva- (Bolch et al. 2008), and laser altimetry (Sauber et al.
tions and, with multitemporal datasets, elevation 2005). Very few have compared remotely sensed
changes. Spaceborne radar altimeters (on board mass balance calculations with those derived by
ERS-1, ERS-2, or Envisat, for example) are well field investigation, but those that have (e.g., Hagg
suited to flat and large ice sheets or ice caps but et al. 2004, Rabatel et al. 2005, Zemp et al. 2010)
are not suitable to survey mountain glaciers have found good agreement between dataset
targeted by GLIMS because of their kilometric results. As noted above, because of the limited accu-
footprints. The smaller (60 m) footprint of the racy of elevation data derived by remote-sensing
Geoscience Laser Altimeter System (GLAS) on methods, many studies have strived to obtain mass
board ICESat makes it a potential source of data balance estimates derived over longer, preferentially
for large mountain glaciers and ice caps (Zwally et decadal time-scales, to ensure the change in
al. 2002). However, the large spacing of the orbits, observed data exceeds the uncertainty in the
the fact that they are not repetitive, and the diffi- approach.
culty of removing the influence of clouds limit the
usefulness of ICESat data. Still, ICESat data can
be used to evaluate the quality of other DEMs
(Berthier and Toutin 2008) or to estimate glacier 5.6 GLACIER MAPPING
elevation changes when compared with other
multitemporal elevation data (Sauber et al. 2005, Snow and ice exhibit distinct spectral properties
Surazakov and Aizen 2006, Kääb 2008). Repetitive that make accurate delineation of debris-free
surveys by airborne laser altimetry have been shown glaciers possible using a variety of remote-sensing
to be extremely accurate. However, central line sur- techniques (e.g., thresholded ratio images), but the
face profiling may lead to errors due to the difficulty automatic classification of debris-covered ice is
of extrapolating to the whole glacier surface more complex, because of the inherent difficulty
(Arendt et al. 2006). of spectrally distinguishing supraglacial sediment
It is clear then that DEMs derived from remote and landslide material from that present in the
sensing can provide glacier elevation change data surrounding terrain (Paul et al. 2004b). For this
on a range of temporal and spatial scales. While reason, DEMs are widely used for terrain visualiza-
useful in their own right, elevation change data tions and to provide additional information on
are also a fundamental component of glacier mass landscape morphometry, land surface context,
balance calculations. If two DEMs cover the whole and spatial topology, all of which facilitate a more
glacier, their difference (i.e., volume change) can be accurate and reliable classification result for debris-
converted to mass balance assuming a constant covered glaciers.
density of 900 kg m 3 in the ablation area, although At the most basic level, DEM data can be used to
the true density value to use in the accumulation generate terrain visualizations to aid the manual
zone is still the topic of ongoing research. Most delineation of glacier boundaries (Fig. 5.8). Glacier
authors use the same density as in the ablation zone, termini are often difficult to identify using plani-
as using a density different from 900 kg m 3 implies metric imagery alone because, particularly when
a change in the rate of densification. Ideally the heavily debris-covered, they appear contiguous
calculation should include quantification describing with the proglacial zone, but small elevation varia-
any shift in the underlying ground (e.g., due to tions visible in three-dimensional visualizations can
132 Digital terrain modeling and glacier topographic characterization

et al. 2004b, Bolch et al. 2007). However, specific


slope thresholds used to delineate glacier facies vary
from region to region, with ice masses in the Euro-
pean Alps generally characterized by short, steep
tongues relative to the long (>10 km) flat tongues
of parts of the Himalaya, for example. Further, this
approach depends on slope angle being calculated
across a sufficiently broad spatial kernel to avoid
highlighting local slope variability (e.g., natural
peaks and troughs across the glacier surface), while
retaining sufficient detail so that small-scale features
such as inner moraine flanks are not excluded from
the analysis.
Additional first-order information that can be
calculated from DEM data include landscape plan
and profile curvatures, which can be used to define
both glacial and nonglacial landform elements
(Bolch and Kamp 2006; see Fig. 5.5c–f ). By cluster-
ing surfaces with similar curvature characteristics it
is possible to differentiate between glacier surfaces
and valley bottoms (low convexity), ridges and
lateral moraines (high convexity), medial moraines
(moderate convexity), and transitions between
glacier margins and lateral moraines (high concav-
ity). A similar but more complex approach is to use
hierarchical object-oriented modeling, which
employs various DEM-derived terrain object prop-
Figure 5.8. Simple 3D visualization (SPOT) HRVIR erties such as slope angle, slope azimuth, curvature,
panchromatic imagery acquired October 17, 2008, and relief to identify locally contiguous portions of
overlain on SPOT HRS-derived DEM from imagery the landscape, which are iteratively aggregated to
spanning 2002–2004) of the Batura Glacier, Pakistan, form higher order landform objects at smaller and
which could be used to aid glacier boundary or catch- smaller scales (Bishop et al. 2001). Glacier bound-
ment headwall delineation, where they are unclear
aries can thus be delineated at a landform level, and
using planimetric data in isolation, for example. For
scale, the glacier is approximately 2 km in width across
objects and the relationships between objects can
the main trunk. Copyright CNES 2008/Distribution be used for further landscape analyses, such as
Spot Image. Figure can also be viewed as Online Sup- studying the role of surface processes (e.g., ero-
plement 5.2. sion/uplift) in topographic evolution. Morpho-
metric approaches such as these show great
potential for glacier mapping across the world,
sometimes remove this ambiguity. Similarly, the but are heavily dependent on the input of high-
exact location of a catchment divide may be quality DEM data for accurate results.
unclear, especially where fresh snowfall masks the A range of satellite sensors already have stereo-
headwall ridge; three-dimensional information is scopic imaging capability and some recent missions
particularly useful in this respect. For the glacier (e.g., Cartosat-1, ALOS PRISM, Terra-SAR-X,
itself, DEM data are useful for identifying areas and the TanDEM-X-mission) have significantly
of low slope angle (see Fig. 5.5b), which are char- increased the potential for extracting DEM data
acteristic of debris-covered glacier tongues in most at very high resolution for use in glacier mapping.
parts of the world. When combined with comple- For many remote high-mountain regions of the
mentary data such as land cover classifications, and world, however, gaining access to high-quality
with results from neighborhood and change detec- DEMs can be problematic, with ASTER and
tion analyses, it has been shown that slope angle SRTM data often providing the best (or only) avail-
data can be used as the primary morphological able sources. Errors common in ASTER-derived
characteristic to delineate glacierized terrain (Paul DEMs of steep, high-mountain relief can sometimes
Glacier mapping 133

produce artifacts in classification results, but may cation, let R denote the pixel domain of a set of
still provide a quality of output that exceeds classi- co-registered data layers (e.g., satellite imagery,
fication methods not employing a DEM of any sort elevation and geomorphometric parameters). Com-
(Paul et al. 2004b). DEM resolution can be an bining all these layers into vector space results in a
equally important consideration, with aerial feature vector per pixel. Similar vectors can be
photography DEMs (such as those provided by derived to describe a shading field (to handle
the Swiss Topo Survey) outperforming satellite sen- across-image slope and exposition variability) and
sor–derived data in almost all cases (Bolch and a segmentation field, such that:
Kamp, 2006), particularly in the delineation of
small ice masses. In areas where accurate and x : R ! R 000
~ vector field of features (observable)
fine-resolution DEMs are not available, manual ~: R ! R
 000
shading field (unknown)
digitization of small samples may still therefore
be necessary to achieve a satisfactory classification s:R!K segmentation field (unknown)
result. Therefore, attempts have been made to intro-
duce additional information into the classification where K denotes the set of labels used for segmen-
process, such as metrics relating to geomorph- tation. In the simplest case there are only two poss-
ometry, texture, and context (Bishop and Shroder ible labels—one for a debris-covered glacier and
2000, Bishop et al. 2001, Taschner and Ranzi 2002, another for ‘‘background’’, but more are usually
Paul et al. 2004b, Ranzi et al. 2004, Buchroithner employed to gain a full description of the environ-
and Bolch 2007, Bolch et al 2008). Sophisticated ment. The statistical model relates the three quan-
classification approaches use both first and sec- tities (above) by a probability distribution:
ond-order topographic derivatives to segment land-
pð~ ~; sÞ ¼ pð~
x;  xj ÞpðsÞ
~; sÞpð~
scape units accordingly, and make use of statistical,
artificial intelligence and hierarchical–structural where it is assumed that shading and segmentation
methods. are a priori independent.
If the parameters of the normal distribution are
5.6.1 Pattern recognition already known, then the recognition task simply
requires the following: given the observation field
Pattern recognition techniques perform two key x we have to estimate the best shading field 
~ ~ and
tasks: description and classification. Given an then estimate the best segmentation field s. Routine
object (or set of objects—pixels in a satellite image, algorithms such as the Gibbs sampler (Geman et al.
for example), pattern recognition systems first seek 1990) and the Expectation Maximization algorithm
to describe those object(s), often in three dimen- (Schlesinger 1968, Dempster et al. 1977) are nor-
sions, and subsequently classify the objects based mally employed to make the estimations.
on those descriptions. The rigor with which the If the parameters of multivariate normal distri-
prior task is performed largely governs the quality butions are unknown (first factor in the equation
of the latter. The description and classification of above) they can be estimated by maximum likeli-
input data in this way can be handled by statistical hood learning given an expert’s segmentation of the
or structural methods or, more frequently, a com- scene. In this case the Expectation Maximization
bination of both (the hybrid approach). Statistical algorithm is again applied but with partial super-
methods make use of decision theory concepts to vision. The expert’s segmentation is relaxed by
segment the landscape based on their quantitative defining only three regions: background, fore-
features, whereas structural methods make use of ground, and a zone between, leaving the system
syntactic grammars to segment the landscape based to learn the optimal boundary. The resulting itera-
on the arrangement of their morphological features tive learning scheme comprises repetition of the
(see Section 5.6.3). Either way, DEM data are a following subtasks:
fundamental input to the classification procedure.
Statistical pattern recognition techniques are
largely dependent on the choice of model param- 1. Given the actual parameters of the normal dis-
eters, the quality of the expert knowledge used in tributions and the actual estimation of the shad-
the prior part of the model, and the resolution of ing field, sample the segmentations and estimate
the image and DEM data. For example, and with the marginal probabilities pr ðsr ¼ k j  xÞ for
~; ~
specific reference to debris-covered glacier classifi- each pixel r.
134 Digital terrain modeling and glacier topographic characterization

Figure 5.9. (a) Optical satellite image of the classification area (ASTER image acquired October 23, 2003);
(b) relaxed expert segmentation (see text); (c) first component of estimated shading; and (d) final segmentation
obtained after learning. Black, green and red encode background labels, blue and violet encode foreground labels,
yellow encodes the boundary label, and cyan encodes lakes. Figure can also be viewed as Online Supplement 5.3.

2. Use these marginals to improve the parameters range of spectral values) and then follow one of two
of the normal distributions (i.e., the covariance paths based on the result. Such ‘‘expert systems’’
matrices and mean vectors for all segments). systematically arrive at a single classification for
3. Re-estimate the shading field  ~ based on the each pixel by ruling out alternatives based on input
current marginal probabilities and parameters data values. This mimics the decision-making pro-
of the normal distributions. cesses employed by a field geomorphologist, for
example; indeed it is important to note that such
When the optimal boundary is reached, the itera- systems can also exploit three-dimensional contex-
tion is stopped and the pixels are assigned their tual and shape information in the same way a human
segment label (Fig. 5.9). can. DEMs are therefore a fundamental component
of a comprehensively designed (and truly three-
dimensional) expert system.
Artificial neural networks (ANNs) aim to process
5.6.2 Artificial intelligence techniques information in a similar way to the human brain, by
Artificial intelligence techniques (often referred to using a large number of highly interconnected pro-
as knowledge-based systems) essentially comprise a cessing elements (neurons) working in parallel to
set of logical rules that provide an analog for the solve a specific problem. The key difference between
reasoning used in anthropogenic decision making. ANNs and other artificial intelligence techniques
These can be as simple as a series of IF–THEN– is that ANNs learn by example and cannot be
ELSE-type statements that test input data (e.g., a designed to perform a specific task in the first
pixel) for certain characteristics (e.g., to be within a instance. ANNs are particularly useful for handling
Discussion 135

diverse and nonparametric datasets and for model- to this approach, if the spatial entities at the lower
ling nonlinear relationships, making them a particu- end of the hierarchical structure are to be accurately
larly attractive method for mapping glacierized quantified and error propagation through the sys-
terrain. Nevertheless, the value of the output classi- tem is to be avoided. This approach has been
fication is only as good as the quality of the input successfully employed to accurately delineate the
data; for ANN techniques to be truly robust they ablation zone of the heavily debris-covered Raikot
must learn to recognize patterns associated with Glacier in the Nanga Parbat Himalaya, for example
geomorphometric land surface parameters, such (Bishop et al. 2001). Terrain object properties were
as texture, context, and spatial topology, which found to be diagnostic of glacier processes repre-
can be derived (at least partly) from first-order senting glacierization, thereby differentiating glacier
and second-order DEM analyses (Bishop et al. topography from other surfaces governed by differ-
2001, 2004). ent surface processes.
Fuzzy classification algorithms are able to incor- Object-oriented mapping can be used as a classi-
porate inaccurate sensor measurements, vague class fication technique in its own right, but it has been
descriptions, and imprecise modeling into the anal- shown to be particularly effective when used in
ysis (Binaghi et al. 1997, Benz et al. 2003), and can combination with ANNs (see above) to classify
result in accurate portrayals of subpixel mixtures of alpine glacial environments. In this specific context,
classes. Fuzzy classification procedures assign the procedure may follow five key stages (Raup et
values to each pixel representing the pixel’s degree al. 2007):
or probability of membership in each class, rather
than clustering similar pixels together into a 1. Classification of land cover using spectral data
thematic set. The fuzzy approach is analogous to and topography.
human perception, in which there can be uncer- 2. Spatial analysis of imagery to generate geo-
tainty even between experts analyzing the same en- metric, shape, and topological information.
vironment. Fuzzy sets are a classification tool in 3. Generation of geomorphometric parameters
themselves; they can be used alone or incorporated from DEMs.
into higher level artificial intelligence techniques 4. Fusion of data into an object-oriented param-
such as expert systems; either way it is assumed that eterization scheme.
there is a priori expert knowledge about the en- 5. Classification of supraglacial features and
vironment under investigation to be able to design glaciers by neural networks.
fuzzy membership functions. In a similar way to
expert systems and ANNs, the addition of a topo- This hybrid approach appears to reduce error and
graphic dataset into the classification is fundamen- increase consistency in results when compared with
tal to the accurate identification of glacial other approaches. Indeed, it is true of all the pre-
landforms through shape and contextual analyses viously described techniques for glacier mapping
(Schneevoigt et al. 2008). that they may yield useful terrain classification in
their own right, but they have even greater potential
when they are integrated together. Fundamental to
5.6.3 Object-oriented mapping
their success is the provision of a three-dimensional
Object-oriented mapping describes the segmenta- analysis, however. Topological, contextual, shape,
tion and partitioning of the landscape into discrete and geomorphometric analyses are limited without
spatial entities based upon their geomorphometric the use of a DEM, and numerous studies have
characteristics (e.g., slope angle, slope azimuth, cur- demonstrated the need for such three-dimensional
vature and relief; Bishop et al., 2001). These entities properties if fully automated glacier mapping is to
represent the lowermost level of a hierarchical land- be realized (Bishop et al. 2004, Paul et al. 2004b,
scape structure, which in total describes the sup- Bolch and Kamp 2006).
position of surface processes that comprise the
topography of a glacierized landscape (Bishop
and Shroder 2000). Therefore, by combining spatial 5.7 DISCUSSION
entities at the lowermost level, specific terrain feat-
ures can be identified; by combining terrain features It is clear that digital elevation data derived from
at the second level, landforms can be identified, and stereoscopic airborne and satellite imagery are fun-
so on. High-quality DEMs are clearly fundamental damental to many essential cryospheric applica-
136 Digital terrain modeling and glacier topographic characterization

tions, and in particular to deriving glacier statistics example, the geodetic method has been proven a
for inclusion in global databases. Square-gridded useful tool to assess glacier surface lowering and,
(raster) DEMs are most widely employed by subsequently, to produce estimates of mass balance.
researchers because they provide more realistic ter- This work, however, has yet to be systematically
rain representations than data interpolated from conducted in many regions of the world. Further,
topographic maps and are easier to derive and to where independent studies have been carried out
handle than DEMs derived by radar interferometry using such approaches, they are not always directly
or airborne laser scanning. However, they are far comparable because of methodological incon-
from perfect, particularly for detecting the abrupt sistencies, or inconsistency in the reporting of data
changes in topography that characterize glacial error.
environments, and the challenge of producing Although remote-sensing sensor technology
fine-resolution DEM data with low data storage has advanced significantly in recent years, the
requirements remains an issue. Further, the genera- challenges of representing topography in high-
tion of reliable elevation data at any resolution finer mountain high-relief regions, and accurately quan-
than that offered by the SRTM (90 m) remains tifying the error within those topographic data,
problematic for many glacierized areas of the remain significant. For example, a range of studies
world. This is partly because of persistent cloud have shown that for many glacierized regions of the
cover, particularly in mountain regions, and a lack world (mostly those of moderate relief ) the topog-
of ground control data, particularly in areas of raphy can be characterized by ASTER DEMs with
political sensitivity. Additionally, SRTM data are an accuracy of 15–30 m (68% confidence level)
static and do not allow for multitemporal analyses. after some significant postprocessing, but only to
While the recently released G-DEM2 may represent 60 m (68% confidence) in areas with steep rock
a possible breakthrough in terms of spatial detail headwalls and large low-contrast accumulation
(30 m pixel size) and accuracy, elevation values areas (Kääb 2002, Toutin 2008). The issue is com-
represent the combined period of ASTER opera- plicated further in high-relief mountain areas by
tions (2000–2011), so offer no better temporal reso- slope–aspect error dependence (Bolch 2004), which
lution than the SRTM DEM. tends to result in better quality elevation data being
Indeed, it is this variability in data quality that derived on slopes facing south (in the northern
remains the major limiting factor in global glacier hemisphere) because of advantageous illumination
mapping and characterization. It has consequences and sensor–terrain viewing angle. Illumination
for the type, number, and quality of glacier param- variations can be compensated for (to a certain
eters that can be derived for any given area. degree) using topographic normalization methods,
Furthermore, it influences the effectiveness of dif- but this itself requires a DEM, which is not always
ferent methods and approaches, and makes the available at this processing step. In terms of
standardization of procedures a challenging task. error quantification, some standardization of
Ideally, elevation data at resolutions and accuracies methods is required. Inconsistencies currently exist
comparable with those offered by airborne laser between studies in the manner in which errors
altimetry are required for all glacierized regions are reported: some studies quote errors to two
of the world. Unfortunately, without a wide-swath standard deviations whereas some only quote the
spaceborne altimeter, we are many years (if not RMSE. A limited number do not even quantify the
decades) from being able to achieve such a dataset. expected error (or uncertainty) in the presented data
Even then, it will raise numerous issues regarding and, of those that do, the accuracy of data can
high-density point clouds and preprocessing. sometimes be highly dependent on the number
Nevertheless, there remains great potential in the and positioning of chosen check points (Toutin
use of various approaches, such as the integration 2008). It is often difficult, therefore, for an indepen-
of multiple multiscale elevation datasets for deriv- dent researcher to replicate results, or to establish
ing multiscale information about a single glacier- the exact error analysis that has taken place and
ized catchment, and the use of LiDAR data as how reliable it is.
accurate ground control for photogrammetric Where reliable topographic data do exist they can
DEM extraction, among others. While novel be used very effectively for mapping and change
approaches to deriving glacier parameters have detection studies. First-order geomorphometric
been successful, some are yet to be effectively and parameters have been successfully employed to
comprehensively employed on a regional basis. For enhance glacier boundary delineation, calculate sur-
Discussion 137

face lowering rates, and inform numerical models As a guide, the resolution of the derived DEM can
(e.g., surface uplift/erosion). This latter effort, provide a practical indication of the scale of infor-
specifically the integration of remote sensing and mation content; analyses of processes or features
modeling for the analysis of landform–process that occur on a finer scale than this should be made
coupling, remains in its early stages, along with with caution. For example, relatively small-scale
several other interesting possibilities that are bene- features (e.g., moraine ridges) can be entirely missed
fiting from better and more widely available digital by medium-resolution datasets, yet may change an
elevation data. For example, space-based methods interpretation of a process or set of features with
now exist capable of generating a mass balance their presence. Conversely, with a fine-resolution
index to improve ELA estimation, and glacier flow DEM, analyses made on a coarser scale may be
and erosion modeling are being enhanced by hampered by noise, and indeed a coarser resolution
improved and accurate boundary conditions. With DEM may be more appropriate for use. With con-
such ventures come further challenges, of course, stant computational developments, methods for
such as accounting for spatial and temporal fluctua- representing fine-scale shape and structure are con-
tions that were previously beyond modeling resolu- tinually improving and so is the incorporation of
tion; many glacier simulations have until recently terrain structure into considerations of spatial scale.
neglected to account for daily and seasonal abla- A classic example of this is the development of
tion, ice velocity, and erosion fluctuations, for scale-dependent classification methods, which are
example (MacGregor et al. 2000). increasingly considering terrain hierarchies with
The accurate modeling of glacierized terrain is the realization that landforms are often not
essential because it plays such a fundamental role spatially delimited and that one landform may par-
in the modulation of atmospheric, Earth surface, tially comprise another.
and glaciological processes (as well as being shaped Consequently, research on glacier mapping and
by these same processes). The glacial environment, the extraction of glacier parameters indicates that
and in particular the glacier surface, is extremely the integration of topographic information on a
labile; this is especially so at present with major variety of spatial scales is important in order to
adjustments occurring due to climatic forcing. On produce reliable results. However, methodologies
a catchment scale, topography governs sediment significantly vary, and some problems remain for
transport and ice fluxes, collectively influencing high-mountain areas, particularly in change detec-
glacier erosion. On a more local scale, glacier sur- tion studies. For example, surface features must be
face topography can influence debris cover distri- maintained across the imaged period for accurate
butions, debris depth, meltwater routing, and elevation datasets to be derived. This depends on
supraglacial ponding. Conversely, processes and there being sufficient optical contrast between the
parameters provide feedback that affects glacier feature and its surroundings as well as the feature
topography—thick debris covers can insulate being large enough to be resolved given the sensor
glacier ice, for example, and meltwater ponds ther- resolution; and it is also dependent on there being
mally erode ice at their margins and their bases. At no (or minimal) modification of the feature shape,
the most local level glacial features such as ogives, size, and contrast during the change, or the masking
moraines, seracs, and crevasse fields define the of the feature by snow cover, for example. Further,
topography of glacier surfaces. It is therefore clear the magnitude of the change(s) to be detected may
that there is complex two-way interaction between not exceed the uncertainty in the method employed.
topography and surface processes across a range of Errors in satellite-derived DEMs of high-relief
spatial and temporal scales that needs to be better terrain are particularly problematic; consequently
understood to inform ideas relating to geomorpho- it is unlikely that researchers will be able to reliably
logical landscape evolution. detect surface lowering on a timescale of less than a
The interaction between topography and most decade. The solution is to either increase the image
geophysical processes occurs over a range of spatial resolution (and thus accuracy) or to extend the
scales, which cannot currently be truly represented observation period (thus increasing the magnitude
within DEM data. Therefore, while the resolution of change). The latter solution is becoming increas-
of the generated DEM is limited by the resolution ingly possible as data archives extend with time and
of input source data, scientific interpretations new sources of historical data (e.g., Corona; Bolch
should also be mindful to restrict analyses to the et al. 2008) emerge. The former issue is aided by the
natural scale of the terrain-dependent application. launch of new and improved sensors.
138 Digital terrain modeling and glacier topographic characterization

Remote sensing is a constantly advancing disci- appropriate for the extraction of elevation informa-
pline, so the launch of a new and/or improved tion, but ASTER remains the most widely used data
sensor is often imminent and the promise of more source because of its stereoscopic capability, wide
accurate or more detailed imagery is never far away. spectral range, medium-to-fine spatial resolution
Some of the most promising developments for the and, importantly, low cost. The extraction of eleva-
generation of new elevation data come from non- tion information from ASTER data performs well
optical sources. For example, the recently successful in areas with low relief and gently sloping topog-
launch of TanDEM-X, the interferometric partner raphy, but errors can be large in areas with steep
of TerraSAR-X, promises to provide a global DEM rock headwalls and in areas with low contrast (e.g.,
to HRTI-3 specifications (Weber et al. 2006; i.e., 12 fresh snow cover). A range of postprocessing tools
m spatial resolution, <2 m relative vertical accuracy can be used to identify and reduce such errors
and <10 m absolute vertical accuracy). These before any secondary analyses are undertaken.
specifications can only currently be achieved by Digital elevation models from nonoptical sources
aerial photogrammetry and very fine–resolution (e.g., LiDAR, InSAR) are becoming increasingly
optical satellite imagery such as WorldView, but common and can often improve on, or be fused
with associated costs. Spaceborne laser altimeters with, existing elevation datasets to enhance cryo-
(e.g., GLAS, on board ICESat-1) provide compar- spheric studies. Such sources have tremendous
able accuracy, but with very limited swath width potential for the extraction of three-dimensional
and low point density, and there is no real prospect data over glacierized terrain in the future, and in
of routine collection of such data again until the particular for increasing the temporal resolution of
proposed launch of ICESat-2 in 2015. change detection analyses, which is currently of the
If topographic and spectral data are to be truly order of a decade or more. However, they may
integrated into surface process modeling, and also bring new challenges in terms of handling data
results compared between independent studies, volume (particularly in the case of LiDAR point
there is a requirement for a representational frame- clouds) and in upscaling acquired information,
work that scientists can work towards. Currently, where such detail is not required. For this reason,
scientists employ a range of analytical tools, algo- the development and publication of standard meth-
rithms, processing approaches, and software for the ods (or protocols) for acquiring, processing, and
generation, manipulation, and interpretation of representing digital elevation data will become
topographic data. Methods are highly empirical, increasingly important. In the short term, chal-
thus the type and quality of the derived data are lenges remain in identifying and quantifying alti-
dependent, to a large extent, on the analyst. Con- metric errors, particularly when comparing DEMs
sequently, replication of existing results can be from different sources, and in simply gaining access
difficult (even more so the application of one pub- to reliable data for some of the most politically
lished technique to a new area). Standardization sensitive regions of the world.
and protocols for information extraction and inte-
gration are therefore required if data quality, result
accuracy, and the validity of comparing measure- 5.9 ACKNOWLEDGMENTS
ments across different studies (and study areas) are
to be assured. Quincey was funded by a Research Council U.K.
Fellowship; Bishop was funded by the National
Aeronautics and Space Administration under the
5.8 CONCLUSIONS NASA OES-02 program (Award NNG04GL84G).
ASTER data courtesy of NASA/GSFC/METI/
Digital elevation models are fundamental to the Japan Space Systems, the U.S./Japan ASTER
extraction of three-dimensional glacier parameters Science Team, and the GLIMS project.
for inclusion in global databases (e.g., WGI,
GLIMS, GlobGlacier) as well as for calculating
mass balance data, automatically delineating 5.10 REFERENCES
glacier boundaries, modeling surface energy bal-
ance and radiation fluxes, characterizing geomor- Allen, T.R. (1998) Topographic context of glaciers and
phometry, and analyzing altitudinal-dependent perennial snowfields, Glacier National Park, Montana.
processes. A range of sensors exists that offers data Geomorphology, 21, 207–216.
References 139

Arendt, A., Echelmeyer, K., Harrison, W.D., Lingle, G., multisource remote sensing images. IEEE Transactions
and Valentine, V. (2002) Rapid wastage of Alaska on Geoscience and Remote Sensing, 35(2), 326–340.
glaciers and their contribution to rising sea level. Bishop, M.P. and Shroder, J.F., Jr. (2000) Remote
Science, 297, 382–386. sensing and geomorphometric assessment of topo-
Arendt, A.A., Echelmeyer, K., Harrison, W., Lingle, C., graphic complexity and erosion dynamics in the Nanga
Zirnheld, S., Valentine, V., Ritchie, B. and Drucken- Parbat massif. Geological Society, London, Special
miller, M. (2006) Updated estimates of glacier volume Publications, 170, 181–200.
changes in the western Chugach Mountains, Alaska, Bishop, M.P., Shroder, J.F., Jr., Hickman, B.L., and
and a comparison of regional extrapolation methods. Copland, L. (1998) Scale dependent analysis of satellite
Journal of Geophysical Research, 111, F03019. imagery for characterization of glacier surfaces in the
Arnold, N.S., Rees, W.G., Hodson, A.J., and Kohler, J. Karakoram Himalaya. Geomorphology, 21, 217–232.
(2006a) Topographic controls on the surface energy Bishop, M.P., Bonk, R., Kamp, U., and Shroder, J.F., Jr.
balance of a high Arctic valley glacier. Journal of Geo- (2001) Terrain analysis and data modelling for alpine
physical Research, 111, F02011. glacier mapping. Polar Geography, 25(3) 182–201.
Arnold, N.S., Rees, W.G., Devereux, B.J., and Amable, Bishop, M.P., Shroder, J.F., Jr., and Colby, J.D. (2003)
G.S. (2006b) Evaluating the potential of high- Remote sensing and geomorphometry for studying
resolution airborne LiDAR data in glaciology. Inter- relief production in high mountains. Geomorphology,
national Journal of Remote Sensing, 27(6), 1233–1251. 55, 345–361.
Baltsavias, E.P. (1999) A comparison between photo- Bishop, M.P., Barry, R.G., Bush, A.B.G., Copland, L.,
grammetry and laser scanning. ISPRS Journal of Dwyer, J.L., Fountain, A.G., Haeberli, W., Hall,
Photogrammetry and Remote Sensing, 54, 83–94. D.K., Kääb, A., Kargel, J.S. et al. (2004) Global
Bamber, J.L., and Payne, A.J. (2004) Mass Balance of the land-ice measurements from space (GLIMS): Remote
Cryosphere: Observations and Modelling of Contempor- sensing and GIS investigations of the Earth’s cryo-
ary and Future Changes. Cambridge University Press, sphere. Geocarto International, 19(2), 57–84.
Cambridge, U.K. Bolch, T. (2004) Using ASTER and SRTM DEMs for
studying glaciers and rockglaciers in northern Tien
Bamber, J.L., and Rivera, A. (2007) A review of remote
Shan. Paper presented at Proceedings of Theoretical
sensing methods for glacier mass balance determina-
and Applied Problems of Geography on a Boundary of
tion. Global and Planetary Change, 59, 138–148.
Centuries, Almaty, Kazakhstan, June 8–9, 2004, Part I,
Barry, R.G. (2006) The status of research on glaciers and
pp. 254–258.
global glacier recession: A review. Progress in Physical
Bolch, T., and Kamp, U. (2006) Glacier mapping in high
Geography, 30(3), 285–306.
mountains using DEMs, Landsat and ASTER data.
Benz, U.C., Hofmann, P., Willhauck, G., Lingenfelder,
Paper presented at Proceedings Eighth International
I., and Heynen, M. (2003) Multi-resolution, object-
Symposium on High Mountain Remote Sensing Carto-
oriented fuzzy analysis of remote sensing data for
graphy, March 20–27, 2005, La Paz, Bolivia (Grazer
GIS-ready information. ISPRS Journal of Photogram-
Schriften der Geographie und Raumforschung, 41),
metry and Remote Sensing, 58(3/4), 239–258. Institut für Geographie und Raumforschung, Univer-
Berthier, E., and Toutin, T. (2008) SPOT5-HRS digital sität Graz, Austria, pp. 13–24.
elevation models and the monitoring of glacier eleva- Bolch, T., and Schröder, H. (2001) Geomorphologische
tion changes in North-West Canada and South-East Kartierung und Diversitätsbestimmung der Periglazial-
Alaska. Remote Sensing of Environment, 112, 2443– formen am Cerro Sillajhuay (Chile/Bolivien) (Erlanger
2454. Geographische Arbeiten, Sonderband 28), Institut für
Berthier, E., Arnaud, Y., Vincent, C., and Remy, F. Geographie, Universität Erlangen, Germany, 141 pp.
(2006) Biases of SRTM in high-mountain areas: Impli- [in German].
cations for the monitoring of glacier volume changes. Bolch, T., Kamp, U., and Olsenholler, J. (2005) Using
Geophysical Research Letters, 33, L08502. ASTER and SRTM DEMs for studying geomorphol-
Berthier, E., Arnaud, Y., Rajesh, K., Sarfaraz, A., ogy and glaciers in high mountain areas. In: M. Oluic
Wagnon, P., and Chevallier, P. (2007) Remote sensing (Ed.), New Strategies for European Remote Sensing:
estimates of glacier mass balances in the Himachal Proceedings 24th Annual Symposium EARSeL, May
Pradesh (Western Himalaya, India). Remote Sensing 25–27, 2004, Dubrovnik, Croatia, Millpress, Rotter-
of Environment, 108(3), 327–338. dam, The Netherlands, pp. 119–127.
Berthier, E., Schiefer, E., Clarke, G.K., Menounos, B., Bolch, T., Buchroithner, M., Kunert, A., and Kamp, U.
and Rémy, F. (2010) Contribution of Alaskan glaciers (2007) Automated delineation of debris-covered
to sea-level rise derived from satellite imagery. Nature glaciers based on ASTER data. Paper presented at
Geosci., 3, 92–95. Geoinformation in Europe: Proceedings of 27th
Binaghi, E., Madella, P., Montesano, M.G., and EARSeL Symposium, June 4–7, 2007, Bolzano, Italy,
Rampini, A. (1997) Fuzzy contextual classification of pp. 403–410.
140 Digital terrain modeling and glacier topographic characterization

Bolch, T., Buchroithner, M., Pieczonka, T., and Kunert, in geomorphometry. Paper presented at Proceedings of
A. (2008) Planimetric and volumetric glacier changes in Geomorphometry 2009, Zurich, Switzerland, August
the Khumbu Himal, Nepal, since 1962 using Corona, 31–September 2, pp. 133–139.
Landsat TM and ASTER data. Journal of Glaciology, Dyurgerov, M.B., and Meier, M.F. (2000) Twentieth
54(187), 592–600. century climate change: Evidence from small glaciers.
Braithwaite, R.J. and Raper, S.C.B. (2010) Estimating Proceedings of the National Academy of Science, 97(4),
equilibrium-line altitude (ELA) from glacier inventory 1406–1411.
data. Annals of Glaciology, 50(53), 127–132. Eckert, S., Kellenberger, T., and Itten, K. (2005) Accu-
Brock, B., Willis, I.C., Sharp, M.J., and Arnold, N.S. racy assessment of automatically derived digital eleva-
(2000) Modelling seasonal and spatial variations in tion models from ASTER data in mountainous terrain.
the surface energy balance of Haut Glacier d’Arolla, International Journal of Remote Sensing, 26(9), 1943–
Switzerland. Annals of Glaciology, 31, 53–62. 1957.
Buchroithner, M., and Bolch, T. (2007) An automated Egholm, D.L., Nielsen, S.B., Pedersen V.K., and
method to delineate the ice extension of the debris- Lesemann, J.-E. (2009) Glacial effects limiting moun-
covered glaciers at Mt. Everest based on ASTER tain height. Nature, 460, 884–887.
imagery. Paper presented at Proceedings Ninth Inter- Eldhuset, K., Andersen, P.H., Hauge, S., Isaksson, E.,
national Symposium on High Mountain Remote Sensing and Weydahl, D.J. (2003) ERS tandem InSAR proces-
Cartography (Grazer Schriften der Geographie und sing for DEM generation, glacier motion estimation
Raumforschung, 43), Institut für Geographie und and coherence analysis on Svalbard. International
Raumforschung, Universität Graz, Austria, pp. 71–78. Journal of Remote Sensing, 24(7), 1415–1437.
Burrough, P.A., and McDonnell, R.A. (1998) Principles Etzelmüller, B., and Björnsson, H. (2000) Map analysis
of Geographic Information Systems. Oxford University techniques for glaciological applications. International
Press, Oxford, U.K. Journal of Geographical Information Science, 14, 567–
Carabajal, C.C., and Harding, D.J. (2005) ICESat valida- 581.
tion of SRTM C-band digital elevation models. Geo- Favey, E., Geiger, A., Gudmundsson, G.H., and Wehr,
physical Research Letters, 32, L22S01. A. (1999) Evaluating the potential of an airborne laser-
Carlisle, B.H. (2005) Modelling the spatial distribution of scanning system for measuring volume changes of
DEM error. Transactions in GIS, 9(4), 521–540. glaciers. Geografiska Annaler, 81A, 555–561.
Carr, S., and Coleman, C. (2007) An improved technique Fisher, P.F., and Tate, N.J. (2006) Causes and conse-
for the reconstruction of former glacier mass-balance quences of error in digital elevation models. Progress
and dynamics. Geomorphology, 92, 76–90. in Physical Geography, 30(4), 467–489.
Cogley, J.G., and Jung-Rothenhausler, F. (2004). Uncer- Flach, B., Kask, E., Schlesinger, D., and Skulish, A.
tainty in digital elevation models of Axel Heiber (2002) Unifying registration and segmentation for
Island, Arctic Canada. Arctic, Antarctic and Alpine multi-sensor images. In: L. Van Gool (Ed.), Pattern
Research, 36, 249–260. Recognition (Lecture Notes in Computer Science,
Copland, L., Pope, S., Bishop, M.P., Shroder, J.F., Jr., 2449), Springer-Verlag, Berlin, pp. 190–197.
Clendon, P., Bush, A., Kamp, U., Seong, Y.B., and Furbish, D.J., and Andrews, J.T. (1984) The use of hyp-
Owen, L.A. (2009) Glacier velocities across the central sometry to indicate long-term stability and response of
Karakoram. Annals of Glaciology, 50(52) 41–49. valley glaciers to changes in mass transfer. Journal of
Crippen, R.E. (2010) Global topographical exploration Glaciology, 30(105), 199–211.
and analysis with the SRTM and ASTER elevation Gardelle, J., Arnaud, Y., and Berthier, E. (2011) Con-
models. Geological Society Special Publications, 345, trasted evolution of glacial lakes along the Hindu Kush
5–15. Himalaya mountain range between 1990 and 2009.
Cuartero, A., Felicı́simo, A.M., and Ariza, F.J. (2004) Global Planet. Change, 75, 47–55.
Accuracy of DEM generation from TERRA-ASTER Gardner, A.S., Moholdt, G., Cogley, J.G., Wouters, B.,
stereo data. International Archives of Photogrammetry Arendt, A.A., Wahr, J., Berthier, E., Hock, R., Pfeffer,
and Remote Sensing, 35(B2), 559–563. W.T., Kaser, G. et al. (2013) A reconciled estimate of
de Jong, S.M., and Burrough, P.A. (1995) A fractal glacier contribuions to sea-level rise: 2003 to 2009.
approach to the classification of Mediterranean vegeta- Science, 340, 852–857.
tion types in remotely sensed images. Photogrammetric Geman, D., Geman, S., Graffigne, C., and Dong, P.
Engineering and Remote Sensing, 61, 1041–1053. (1990) Boundary detection by constrained optimiza-
Dempster, A.P., Laird, N.M., and Rubin, D.B. (1977) tion. IEEE Transactions on Pattern Analysis and
Maximum likelihood from incomplete data via the Machine Intelligence, 12(7), 609–628.
EM algorithm. Journal of the Royal Statistical Society, Georgopoulos, A., and Skarlatos, D. (2003) A novel
Series B (Methodological), 39(1), 1–38. method for automating the checking and correction
Drăgut L., Eisank, C., Strasser, T., and Blaschke, T. of digital elevation models using orthophotographs.
(2009) A comparison of methods to incorporate scale The Photogrammetric Record, 18(102), 156–163.
References 141

Gonçalves, J.A., and Oliveira, A.M. (2004) Accuracy test study on Edgeøya, Eastern Svalbard. IEEE Trans-
analysis of DEMs derived from ASTER imagery. actions on Geoscience and Remote Sensing, 46(10),
International Archives of Photogrammetry and Remote 2823–2830.
Sensing, 35, 168–172. Kääb, A., and Funk, M. (1999) Modelling mass balance
Goodchild, M.F., and Mark, D.M. (1987) The fractal using photogrammetric and geophysical data: A pilot
nature of geographic phenomena. Annals of the Asso- study at Griesgletscher, Swiss Alps. Journal of Glaciol-
ciation of American Geographers, 77(2), 265–278. ogy, 45, 575–583.
Gratton, D.J., Howarth, P.J., and Marceau, D.J. (1990) Kääb, A., Haeberli, W., and Gudmundsson, G.H. (1997)
Combining DEM parameters with Landsat MSS and Analysing the creep of mountain permafrost using high
TM imagery in a GIS for mountain glacier character- precision aerial photogrammetry: 25 years of monitor-
ization. IEEE Transactions on Geosciences and Remote ing Gruben rock glacier, Swiss Alps. Permafrost and
Sensing, 28, 766–769. Periglacial Processes, 8, 409–426.
Haeberli, W., Barry, R., and Cihlar, J. (2000) Glacier Kääb, A., Huggel, C., Fischer, L., Guex, S., Paul, F.,
monitoring within the Global Climate Observing Sys- Roer, I., Salzmann, N., Schlaefli, S., Schmutz, K.,
tem. Annals of Glaciology, 31, 241–246. Schneider, D. et al. (2005) Remote sensing of glacier-
Hagg, W., Braun, L., Uvarov, V.N., and Makarevich, and permafrost related hazards in high mountains: An
K.G. (2004) A comparison of three methods of mass overview. Natural Hazards and Earth System Science,
balance determination in the Tuyuksu Glacier Region, 5(4), 527–554.
Tien Shan. Journal of Glaciology, 50(171), 505–510. Kamp, U., Bolch, T., and Olsenholler, J. (2005) Geomor-
Heitzinger, D. and Kager, H. (1999) Hochwertige Ge- phometry of Cerro Sillajhuay (Andes, Chile/Bolivia):
ländemodelle aus Höhenlinien durch wissensbasierte Comparison of digital elevation models (DEMs) from
Klassifikation von Problemgebieten. Photogrammetrie ASTER remote sensing data and contour maps. Geo-
Fernerkundung Geoinformation, 1, 29–40 [in German]. carto International, 20, 23–34.
Hirano, A., Welch, R., and Lang, H. (2003) Mapping Kargel, J.S., Abrams, M.J., Bishop, M.P., Bush, A.,
from ASTER stereo image data: DEM validation Hamilton, G., Jiskoot, H., Kääb, A., Kieffer, H.H.,
and accuracy assessment. ISPRS Journal of Photo- Lee, E.M., Paul, F. et al. (2005) Multispectral imaging
grammetry and Remote Sensing, 57(5/6), 356–370. contributions to global land ice measurements from
Hock, R. (2005) Glacier melt: A review of processes and space. Remote Sensing of the Environment, 99, 187–219.
their modeling. Progress in Physical Geography, 29(3), Kargel, J.S., Leonard, G., Crippen, R.E., Delaney, K.B.,
362–391. Evans, S.G., and Schneider J. (2010) Satellite monitor-
Honikel, M. (2002) Fusion of spaceborne stereo-optical ing of Pakistan’s rockslide-dammed Lake Gojal. EOS
and interferometric SAR data for digital terrain model Trans. Am. Geophys. Union, 91(43), doi: 10.1029/
generation. Mitteilungen des Institutes für Geodäsie und 2010EO430002.
Photogrammetrie, 76. Kargel, J., Furfaro, R., Kaser, G., Leonard, G., Fink, W.,
Huggel, C., Kääb, A., Haeberli, W., Teysseire, P., and Huggel, C., Kääb, A., Raup, B., Reynolds, J., Wolfe,
Paul, F. (2002) Remote sensing based assessment of D. et al. (2011) ASTER imaging and analysis of glacier
hazards from glacier lake outbursts: A case study in hazards. In: B. Ramachandran, C.O. Justice, and M.J.
the Swiss Alps. Canadian Geotechnical Journal, 39(2), Abrams (Eds.), Land Remote Sensing and Global En-
316–330. vironmental Change: NASA’s Earth Observing System
Kääb, A. (2000) Photogrammetric reconstruction of and the Science of Terra and Aqua, Springer-Verlag,
glacier mass balance using a kinematic ice-flow model: New York, pp. 325–373.
A 20 year time series on Grubengletscher, Swiss Alps. Kargel., J.S., Alho, P., Buytaert, W., Celleri, R., Cogley,
Annals of Glaciology, 31, 45–52. J.G., Dussaillant, A., Zambrano, G., Haeberli, W.,
Kääb, A. (2002) Monitoring high-mountain terrain Harrison, S., Leonard, G. et al. (2012) Glaciers in
deformation from repeated air- and spaceborne optical Patagonia: Controversy and prospects. EOS Trans.
data: Examples using digital aerial imagery and Am. Geophys. Union, 93, 212–213.
ASTER data. ISPRS Journal of Photogrammetry and Kaser, G. (2001) Glacier–climate interaction at low lati-
Remote Sensing, 57, 39–52. tudes. Journal of Glaciology, 47, 195–204.
Kääb, A. (2005a) Remote Sensing of Mountain Glaciers Katsube, K., and Oguchi, T. (1999) Altitudinal changes in
and Permafrost Creep (Schriftenreihe Physische Geo- slope angle and profile curvature in the Japan Alps: A
graphie, 48), University of Zurich, Switzerland, 266 pp. hypothesis regarding a characteristic slope angle. Geo-
Kääb, A. (2005b) Combination of SRTM3 and repeat graphic Review of Japan, 72B, 63–72.
ASTER data for deriving alpine glacier flow velocities Khalsa, S.J.S., Dyurgerov, M.B., Khromova, T., Raup,
in the Bhutan Himalaya. Remote Sensing of Environ- B.H., and Barry, R.G. (2004). Space-based mapping of
ment, 94(4), 463–474. glacier changes using ASTER and GIS tools. IEEE
Kääb A. (2008) Glacier volume changes using ASTER Transactions on Geoscience and Remote Sensing,
satellite stereo and ICESat GLAS laser altimetry: A 42(10), 2177–2183.
142 Digital terrain modeling and glacier topographic characterization

Kubik, K., and Botman, A.G. (1976) Interpolation accu- Molnar, P., and England, P. (1990) Late Cenozoic uplift
racy for topographic and geological surfaces. ITC of mountain ranges and global climate change:
Journal, 2, 236–274. Chicken or egg? Nature, 346, 29–34.
Kyriakidis, P.C., Shortridge, A.M., and Goodchild, M.F. Muskett, R.R., Lingle, C.S., Tangborn, W.V., and
(1999) Geostatistics for conflation and accuracy assess- Rabus, B.T. (2003) Multi-decadal elevation changes
ment of digital elevation models. International Journal on Bagley Ice Valley and Malaspina Glacier, Alaska.
of Geographical Information Science, 13, 677–707. Geophysical Research Letters, 30(16), 1857.
Larsen, C.F., Motyka, R.J., Arendt, A.A., Echelmeyer, Nuth, C., and Kääb, A. (2011) Co-registration and bias
K.A., and Geissler, P.E. (2007) Glacier changes in corrections of satellite elevation data sets for quantify-
southeast Alaska and northwest British Columbia ing glacier thickness change. The Cryosphere, 5, 271–
and contribution to sea level rise. Journal of Geophys- 290.
ical Research, 112, F01007. Oksanen, J., and Sarjakoski, T. (2005) Error propagation
of DEM-based surface derivatives. Computers and
Lathrop, R.G., and Peterson, D.L. (1992) Identifying
Geosciences, 31, 1015–1027.
structural self-similarity in mountainous landscapes.
Østrem, G., and Haakensen, N. (1999) Map comparison
Landscape Ecology, 6(4), 233–238.
or traditional mass-balance measurements: Which
Leica Helava (2004) ERDAS Imagine Version 8.7 and
method is better? Geografiska Annaler, 81A(4), 703–
Leica Photogrammetry Suite (LPS) User Guide, Leica
711.
Helava, London.
Paul, F., Kääb, A, Maisch, M., Kellenberger, T., and
Leonard, K.C. and Fountain, A.G. (2003) Map-based Haeberli, W. (2004a) Rapid disintegration of Alpine
methods for estimating glacier equilibrium line alti- glaciers observed with satellite data. Geophysical
tudes. Journal of Glaciology, 49(166), 329–336. Research Letters, 31, L21402.
Li, D., Wang, S., and Li, R. (1996) Automatic quality Paul, F., Huggel, C., and Kääb, A. (2004b) Combining
diagnosis in DTM generation by digital image match- satellite multispectral image data and a digital eleva-
ing techniques. Geomatica, 50(1), 65–73. tion model for mapping debris-covered glaciers.
Li, Z., Zhu, Q., and Gold, C. (2005) Digital Terrain Remote Sensing of Environment, 89, 510–518.
Modeling: Principles and Methodology, CRC Press, Peng, X., Wang, J., and Zhang, Q. (2005) Deriving terrain
Boca Raton, FL. and textural information from stereo RADARSAT
Luckman, A., Quincey, D.J., and Bevan, S. (2007) The data for mountainous land cover mapping. Inter-
potential of satellite radar interferometry and feature national Journal of Remote Sensing, 26(22), 5029–5049.
tracking for monitoring flow rates of Himalayan Quincey, D.J., Richardson, S.D., Luckman, A., Lucas,
glaciers. Remote Sensing of Environment, 111, 172–181. R.M., Reynolds, J.M., Hambrey, M.J., and Glasser,
MacGregor, K.R., Anderson, R.S., Anderson, S.P., and N.F. (2007) Early recognition of glacial lake hazards in
Waddington, E.D. (2000) Numerical simulations of the Himalaya using remote sensing datasets. Global
glacial-valley longitudinal profile evolution. Geology, and Planetary Change, 56, 137–152.
28, 1031–1034. Quincey, D.J., Luckman, A., and Benn, D. (2009) Quan-
Meier, M.F. (1984) Contribution of small glaciers to tification of Everest-region glacier velocities between
global sea level. Science, 226, 1418–1421. 1992 and 2002, using satellite radar interferometry
and feature tracking. Journal of Glaciology, 55, 596–
Meier, M.F., Dyurgerov, M.B., and McCabe, G.J. (2003)
606.
The health of glaciers: Recent changes in glacier
Rabatel, A., Dedieu, J.P., and Vincent, C. (2005) Using
regime. Climatic Change, 59, 123–135.
remote-sensing data to determine equilibrium-line alti-
Mennis, J.L., and Fountain, A.G. (2001) A spatio-
tude and mass-balance time series: Validation on three
temporal GIS database for monitoring alpine glacier
French glaciers, 1994–2002. Journal of Glaciology,
change. Photogrammetric Engineering and Remote
51(175), 539–546.
Sensing, 67, 967–975.
Rabus, B., Eineder, M., Roth, A., and Bamler, R. (2003)
Miller, M.M., and Pelto, M.S. (2003) Mass balance meas- The shuttle radar topography mission: A new class of
urements on the Lemon Creek Glacier, Juneau Icefield, digital elevation models acquired by spaceborne radar.
Alaska 1953–1998. Geografiska Annaler: Series A, ISPRS Journal of Photogrammetry and Remote
Physical Geography, 81(4), 671–681. Sensing, 57, 241–262.
Mitasova, H., and Mitas, L. (1993) Interpolation by reg- Racoviteanu, A.E., Manley, W.F., Arnaud, Y., and Wil-
ularized spline with tension, I: Theory and Implemen- liams, M.W. (2007) Evaluating digital elevation models
tation. Mathematical Geology, 25, 641–655. for glaciologic applications: An example from Nevado
Möller, M., Schneider, C., and Kilian, R. (2007) Glacier Coropuna, Peruvian Andes. Global and Planetary
change and climate forcing in recent decades at Gran Change, 59, 110–125.
Campo Nevado, southernmost Patagonia. Annals of Ranzi, R., Grossi, G., Iacovelli, L., and Taschner, S.
Glaciology, 46, 136–144. (2004) Use of multispectral ASTER images for map-
References 143

ping debris-covered glaciers within the GLIMS Pro- Schlesinger, M. (1968) The interaction of learning and
ject. Paper presented at IEEE International Geoscience self-organization in pattern recognition. Cybernetics
and Remote Sensing Symposium, Vol. II, pp. 1144– and Systems Analysis, 4(2), 66–71.
1147. Schneevoigt, N.J. and Schrott, L. (2006) Linking geo-
Rasemann, S., Schmidt, J., Schrott, L., and Dikau, R. morphic systems theory and remote sensing: A con-
(2004) Geomorphometry in mountain terrain. In: M. ceptual approach to Alpine landform detection
Bishop and J.F. Shroder (Eds.), Geographic Informa- (Reintal, Bavarian Alps, Germany). Geographica Hel-
tion Science and Mountain Geomorpology, Springer- vetica, 3, 181–190.
Verlag, Berlin, pp. 101–137. Schneevoigt, N.J., van der Linden, S., Thamm, H.-P., and
Raup, B., Kääb, A., Kargel, J.S., Bishop, M.P., Hamil- Schrott, L. (2008) Detecting Alpine landforms from
ton, G., Lee, E.M., Paul, F., Rau, F., Soltesz, D., remotely sensed imagery: A pilot study in the Bavarian
Khalsa, S.J. et al. (2007) Remote sensing and GIS Alps. Geomorphology, 93, 104–119.
technology in the Global Land Ice Measurements from Schneider, K. (1998) Geomorphologisch plausible
Space (GLIMS) Project. Computers and Geosciences, Rekonstruktion der digitalen Repräsentation von
33, 104–125. Geländeoberflächen aus Höhendaten. Unpublished
Rees, W.G., and Arnold, N.S. (2007) Mass balance PhD thesis, University of Zurich, Switzerland, 226
and dynamics of a valley glacier measured by high- pp. [in German].
resolution LiDAR. Polar Record, 43, 311–319. Smith, M.J., and Pain, C.F. (2009) Applications of
Reiners, P.W., Ehlers, T.A., Mitchell, S.G., and Mont- remote sensing in geomorphology. Progress in Physical
gomery, D.R. (2003) Coupled spatial variations in pre- Geography, 33(4), 568–582.
cipitation and long-term erosion rates across the Stearns, L.A., and Hamilton, G.S. (2007) Rapid volume
Washington Cascades. Nature, 426, 645–647. loss from two East Greenland outlet glaciers quantified
Rickenbacher, M. (1998) Die digitale Modellierung des using repeat stereo satellite imagery. Geophysical
Hochgebirges im DGM25 des Bundesamtes fuer Research Letters, 34, L05503.
Landestopographie. Wiener Schriften zur Geographie Surazakov, A.B., and Aizen, V.B. (2006) Estimating vol-
und Kartographie, 11, 49–55 [in German]. ume change of mountain glaciers using SRTM and
Rignot, E., Echelmeyer, K., and Krabill, W. (2001) Pene- map-based topographic data. IEEE Transactions on
tration depth of interferometric synthetic-aperture Geoscience and Remote Sensing, 44(10), 2991–2995.
radar signals in snow and ice. Geophysical Research Surazakov, A.B., Aizen, V.B., Aizen, E.M., and Nikitin,
Letters, 28(18), 3501–3504. S.A. (2007) Glacier changes in the Siberian Altai
Rignot, E., Rivera, A., and Casassa, G. (2003) Contribu- Mountains, Ob river basin, (1952–2006) estimated with
tion of the Patagonia Icefields of South America to Sea high resolution imagery. Environmental Research Let-
Level Rise. Science, 302(5644), 434–437. ters, 2, 045017.
Rivera, A., Bown, F., Casassa, G., Acuña, C., and Tachikawa, T., Kaku, M., Iwasaki, A., Gesch, D.,
Clavero, J. (2005) Glacier shrinkage and negative mass Oimoen, M., Zhang, Z., Danielson, J., Krieger, T.,
balance in the Chilean Lake District (40S). Hydrologi- Curtis, B., Haase, J. et al. (2011) ASTER Global Digi-
cal Sciences Journal, 50(6), 963–974. tal Elevation Model Version 2: Summary of Validation
Rivera, A., Benham, T., Casassa, G., Bamber, J., and Results. Available at http://www.ersdac.or.jp/GDEM/
Dowdeswell, J.A. (2007) Ice elevation and areal ver2Validation/Summary_GDEM2_validation_report_
changes of glaciers from the Northern Patagonia Ice- final.pdf
field, Chile. Global and Planetary Change, 59(1/4), 126– Tangborn, W. (1999) A mass balance model that uses
137. low-altitude meteorological observations and the
Roe, G.H., Montgomery, D.R., and Hallet, B. (2002) area-altitude distribution of a glacier. Geografiska
Effects of orographic precipitation variations on the Annaler, Series A, Physical Geography, 81(4), 753–765.
concavity of steady-state river profiles. Geology, 30, Taschner, S., and Ranzi, R. (2002) Comparing the oppor-
143–146. tunities of Landsat TM and Aster data for monitoring
Sauber, J., Molnia, B., Carabajal, C., Luthcke, S., and a debris covered glacier in the Italian Alps within the
Muskett, R. (2005) Ice elevations and surface change GLIMS project. Paper presented at IEEE International
on the Malaspina Glacier, Alaska. Geophysical Geoscience and Remote Sensing Symposium, Vol. II,
Research Letters, 32, L23S01. pp. 1044–1046.
Scharroo, R., and Visser, P. N. A. M. (1998) Precise orbit Toutin, T. (2002) Three-dimensional topographic map-
determination and gravity field improvement for the ping with ASTER stereo data in rugged topography.
ERS satellites. Journal of Geophysical Research, 103, IEEE Transactions on GeoScience and Remote Sensing,
8113–8127. 40(10), 2241–2247.
Schiefer, E., Menounos, B., and Wheate, R. (2007) Toutin, T. (2008) ASTER DEMs for geomatic and geo-
Recent volume loss of British Columbian glaciers, scientific applications: A review. International Journal
Canada. Geophysical Research Letters, 34, L16503. of Remote Sensing, 29(7), 1855–1875.
144 Digital terrain modeling and glacier topographic characterization

Toutin, Th., Blondel, E., Clavet, D., and Schmitt, C.V. in the Pamir Tadjikistan. Unpublished thesis, FHBB
(2013) Stereo radargrammetry with Radarsat-2 in the Muttenz, Basle [in German].
Canadian Arctic. IEEE Transactions on Geoscience and Wilson, J.P., and Gallant, J.C. (Eds.) (2000) Terrain Anal-
Remote Sensing, 51(5), 2601–2609, doi: 10.1109/ ysis: Principles and Applications. John Wiley and Sons,
TGRS.2012.2211605. New York.
van Asselen, S., and Seijmonsbergen, A.C. (2006) Expert- Yadav, R.R., Park, W.K., Singh, J., and Dubey, B. (2004)
driven semi-automated geomorphological mapping for Do the western Himalayas defy global warming? Geo-
a mountainous area using a laser DTM. Geomorphol-
physical Research Letters, 31, L17201.
ogy, 78(3/4), 309–320.
Zemp, M., Jansson, P., Holmlund, P., Gärtner-Roer, I.,
Weber, M., Herrmann, J., Hajnsek, I., and Moreira, A.
(2006) TerraSAR-X and TanDEM-X: Global mapping Koblet, T., Thee, P., and Haeberli, W. (2010) Reana-
in 3D using radar. Paper presented at Proceedings of lysis of multitemporal aerial images of Storglaciären,
the Second International Workshop on ‘‘The Future of Sweden (1959–99), Part 2: Comparison of glaciological
Remote Sensing,’’ Antwerp, Belgium (ISPRS Intercom- and volumetric mass balances. Cryosphere, 4, 345–357.
mission Working Group I/V Autonomous Navigation, Zwally, H.J., Schutz, B., Abdalati, W., Abshire, J.,
36–1/W44), International Society for Photogrammetry Bentley, C., Brenner, A., Bufton, J., Dezio, J.,
and Remote Sensing, Hanover, Germany. Hancock, D., Harding, D. et al. (2002) ICESat’s
Weidmann, Y. (2004) Combination of ASTER- and laser measurements of polar ice, atmosphere, ocean,
SRTM-DEMs for the assessment of natural hazards and land. Journal of Geodynamics, 34, 405–445.

Potrebbero piacerti anche