Anda di halaman 1dari 31

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/316723307

GIS Applications in Geomorphology

Chapter · December 2017


DOI: 10.1016/B978-0-12-409548-9.10029-6

CITATIONS READS

0 4,704

4 authors:

Jan-Christoph Otto Günther Prasicek


University of Salzburg University of Lausanne
106 PUBLICATIONS   540 CITATIONS    29 PUBLICATIONS   81 CITATIONS   

SEE PROFILE SEE PROFILE

Jan Henrik Blöthe Lothar Schrott


University of Bonn University of Bonn
33 PUBLICATIONS   136 CITATIONS    140 PUBLICATIONS   1,951 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

PermArg - Rock glacier permafrost in the Central Andes of Argentina: regional distribution – ice content – hydrological significance. Project start: 11/2015 (funded by
DFG) View project

GeoHype - Geomorphological and hydrological significance of permafrost in the alps View project

All content following this page was uploaded by Jan-Christoph Otto on 08 June 2017.

The user has requested enhancement of the downloaded file.


GIS Applications in Geomorphology☆
Jan-Christoph Otto and Günther Prasicek, University of Salzburg, Salzburg, Austria
Jan Blöthe and Lothar Schrott, University of Bonn, Bonn, Germany
ã 2017 Elsevier Inc. All rights reserved.

Introduction 1
Land Surface Parameters and Geomorphological Indices 3
Data Sources 3
Digital Terrain Models 3
Optical Imagery 6
Optical satellite imagery 6
Unmanned aerial vehicles and structure from motion 7
Other Data Sources 8
Digital Geomorphological Mapping 9
Map Creation 9
Automated Land Surface Classification 9
General land surface classification 9
Specific land surface classification 11
Application of Various Geomorphological Indices for Process and Landform Analysis—Case Study Obersulzbach
valley, Eastern Alps, Austria 11
Hillslopes and Gravitational Processes 11
Glacier Environments 13
Periglacial Environments 16
Fluvial Environments 18
Sediment Flux and Erosion in Mountain Areas 21
Conclusions 25
References 27

Introduction

Modern geomorphological research is inextricably linked with geospatial technology and geographic information systems (GIS).
Driven by rapid technological advances of remote sensing, geodesy, photogrammetry, computer science, and GIS, the application
of analysis tools using digital information on the land surface revolutionized quantitative geomorphological research (Bishop,
2013). In the last three decades, GIS has increasingly influenced various fields of geomorphology. GIS are designed to facilitate
spatial investigations, for example, through geostatistical analyses or the mathematical description of surfaces and are hence
inherently linked to methodology and concepts in geomorphology. GIS tools support and enable many upfront research fields in
geomorphology from the quantitative description of landforms to process modeling, the investigation of form–process interrela-
tions and linkages to climate and environmental conditions, or the assessment of sediment flux. Furthermore, process and form
modeling, statistical analysis and regionalization of field data as well as graphical visualization and map creation are key features of
GIS applied in geomorphology. A starting point for GIS studies commonly is the digital elevation model (DEM) supplemented
with image data of various types (see Section “Data sources”). However, GIS tools also allow linking remotely sensed information
with field data, for example, land surface features, process rates, or subsurface information, recorded with geopositioning systems.
The roots of the first geomorphographic relief analyses can be identified in early studies of Penck (1894). His pioneering ideas of
landforms led to the establishment of taxonomical structures which have been used in many subsequent studies (e.g., Ahnert,
1970; Kugler, 1975; Evans, 1972). A new era in the application of GIS in geomorphological studies, however, started almost 100
years later, in the 1990s. Classic papers by Dikau et al. (1991), Moore et al. (1991), Pike and Dikau (1995), or Wilson and Gallant
(2000) focused on digitally derived landform classifications and general geomorphometrical advances using DEMs, respectively.
First applications of GIS to traditional geomorphological topics such as landslides, soil erosion, and mountain permafrost
distribution were successful on regional or local scales (Chairat and Delleur, 1993; Deroo et al., 1989; Dikau and Jäger, 1995;
Eash, 1994; Jäger, 1997; Keller, 1992; vanWesten and Terlien, 1996; Koethe and Lehmeier, 1993).
Since the late 1990s, we observe an increasing use of GIS in geomorphological studies (see Fig. 1). This development is strongly
related to advances in computer science, remote sensing and photogrammetric techniques, as well as shallow geophysics (Bishop,
2013). In particular, the availability of global digital terrain datasets has boosted applications and research in GIS for land surface
and process analysis. On a global scale, DEMs with resolutions between 1 and 30 m are now available for the entire terrestrial


Supplementary data associated with this article can be found, in the online version, at doi: http://dx.doi.org/10.1016/B978-0-12-409548-9.10029-6.

Comprehensive Geographic Information Systems http://dx.doi.org/10.1016/B978-0-12-409548-9.10029-6 1


2 GIS Applications in Geomorphology

TanDEM-X (1 m)
ASTER GDEM (30 m)

Prerelease of
SRTM DEM (30 m)
Release of
Release of
60

Release of SRTM DEM (90 m)


50

40
Number of papers

Release of GTOPO30 (1km)


30

20

10

0
1990 1995 2000 2005 2010 2015
Year
Fig. 1 Total annual number of papers explicitly including “GIS” in the title, abstract or key words, published in four international journals of
geomorphology from 1989 to 2016 (data source: Webofknowledge). Release dates of global DEM data sets have been included as benchmarks of data
availability. Data from 1989–2009 are taken from Oguchi, T. and Wasklewicz, T. A. (2011). Geographic information systems in geomorphology. In:
Gregory, K. J. and Goudie, A. S. (eds.) The SAGE handbook of geomorphology. London: SAGE.

landmass. In addition, laserscanning (LIDAR: light detection and ranging) and structure from motion (SFM) techniques both
ground- and air-based provide high-resolution DEMs (<1 m) on local and regional scales. Additionally, numerous GIS software
tools, both commercially and open source, are available today, opening unlimited opportunities for scientists. As a consequence,
the use of GIS tools for geomorphological analyses became increasingly popular. Comprehensive reviews on basic elements of
remote sensing techniques and application of GIS in geomorphological research are provided by Bishop (2013) and Oguchi and
Wasklewicz (2011).
Applications of GIS in geomorphology span from pure visualization approaches, landform classification, land surface and
hydrological analysis, process and erosion modeling, topographic change detection to hazard susceptibility modeling. While many
applications focusing on land surface analysis, change detection, or hazard modeling are performed within the specific GIS
software, some approaches use statistical software (e.g., R software package), or special modeling software (e.g., Matlab, IDL, a.
o.) to perform geospatial analysis. For example, modeling of erosional processes and landform evolution often demand require-
ments that exceed capabilities of GIS software and are produced using other resources (e.g., Chen et al., 2014; Coulthard, 2001;
Tucker and Hancock, 2010; also refer to https://csdms.colorado.edu/for a list of available models).
While GIS software became more powerful and even provided advanced graphical tools, a simultaneous increase in geomor-
phological mapping cannot be observed. This is somewhat surprising because the overlay of different geomorphological, litho-,
and pedological information is one of the most important tools in GIS applications and improves the applicability of maps (Otto
and Smith, 2013). However, geomorphological mapping and GIS became a self-evident combination and geomorphological
symbol sets are designed for specific purposes and frequently used (Gustavsson et al., 2006; Otto and Dikau, 2004; Schoeneich,
1993). Moreover, geomorphological maps are now serving as an intermediate product for quantitative sediment budget analyses.
For this, GIS-based modeling of landforms is combined with subsurface information such as soil or regolith thickness, which is
derived from geophysical surveys. The gained knowledge on spatial distribution of sediment storage types plays an important role
in quantitative sediment budget studies (Otto et al., 2009; Schrott et al., 2003b; Theler et al., 2008).
Many useful GIS modeling approaches have been developed in the field of natural hazards. Rockfalls, landslides, floods,
avalanches, or soil erosion share inherent characteristics of hazards such as magnitude or spatial extension and depend strongly on
slope angle, aspect, or other parameters which can be ideally integrated and displayed in GIS environments (e.g., Gruber and
Mergili, 2013; Gruber and Bartelt, 2007; Lan et al., 2007; vanWesten and Terlien, 1996; Wilford et al., 2004; Wichmann and Becht,
2006). Hazard assessment using GIS often combines geomorphometric analysis with geostatistical analysis of related parameters
GIS Applications in Geomorphology 3

to generate models of spatial susceptibility (Carrara and Guzzetti, 1995). Comprehensive reviews concerning methodological
aspects and GIS-based hazard assessments can be found in Guzzetti et al. (1999), Huabin et al. (2005), and van Westen
et al. (2008).
This article gives an overview to various GIS applications in geomorphology. We introduce basic principles of parameters and
indices used for landform and process analysis and briefly highlight typical and innovative data sources and references to
geomorphological mapping. Instead of reviewing the vast amount of GIS applications in the literature we visualize GIS capabilities
for geomorphology by applying a selection of tools and indices in a case study area in alpine terrain (Obersulzbach Valley, Austria,
European Alps). The selection touches various fields of geomorphology, however is far from being complete. Applications in the
following fields are presented:
(i) Hillslope and gravitational processes
(ii) Glacial processes
(iii) Periglacial processes
(iv) Fluvial processes
(v) Sediment flux and erosion in mountain areas

The results of the case study can also be accessed online in a WebGIS application linked on the publisher’s website (see
supplementary information).

Land Surface Parameters and Geomorphological Indices

Quantitative analysis of the land surface is defined by the term geomorphometry, a highly active research field within geomor-
phology (Hengl and Reuter, 2009; Wilson and Bishop, 2013). Its focus is on the quantification of land surface parameters (LSPs)
and the detection of objects from digital elevation data. In turn geomorphometry as a research area builds a theoretical foundation
and serves as a bridge between GIS and geomorphology (Dikau, 1996). Geomorphometric analysis can be separated into general
and specific approaches (Evans, 1972; Goudie, 1990). The main distinction between the two approaches is the continuous or
discontinuous character of the object in focus. General approaches analyze the continuous land surface without addressing specific
landforms or boundaries. Specific geomorphometry aims to identify and describe discrete landforms and their morphological
characteristic. One focus of specific approaches is the extraction of these forms from a continuous surface (see what follows), an
issue that is at the research frontier of geomorphometry (Evans, 2012).
LSPs are geometrical or statistical attributes of a land surface that can be derived directly from a DEM. They can be quantified
locally, or involve a regional analysis approach (Olaya, 2009). While local parameters are quantified for a single location in relation
to its immediate surrounding cells, regional parameters include relations to more distant cells. The most common basic LSPs are
altitude, aspect, slope, and curvature and represent examples for local parameters. Regional parameters include aspects of flow over
the surface, for example, utilized in the modeling of hydrological conditions, calculation of viewsheds, or solar radiation (Gruber
and Peckham, 2009; Böhner and Antonić, 2009). Examples for hydrological LSPs are flow direction, flow accumulation, and
drainage networks architecture. Relating LSPs to the three fundamental concepts in geomorphology, (i) form, (ii) process, and
(iii) material (Gregory and Lewin, 2014), we could identify curvature and slope as principal descriptors of form, altitude, slope, and
contributing drainage area as influential factors of process activity (in case of fluvial and gravitational processes) and surface
roughness as an indicator of surface material characteristics (Otto et al., 2012).
Based on these basic parameters numerous topographic or geomorphological indices have been developed to study geomor-
phological form and process configurations (Table 1). Geomorphological indices are combinations of the primary attributes that
describe or characterize the spatial variability of specific processes or landforms occurring in the landscape and can be used for
landform and process analysis or landscape comparison (Pike and Wilson, 1971; Wilson and Gallant, 2000). These indices are
applied in erosional process modeling, hydrological modeling, or digital soil mapping, to name just a few (Marthews et al., 2015;
Moore et al., 1991). Many geomorphologic indices have been formulated before the rise of GIS resulting from classical geomor-
phological works from the early days of quantitative geomorphology (e.g., Leopold et al., 1964; Strahler, 1952, 1957; Bagnold,
1960). GIS tools, however, facilitate the quantification of these parameters and in combination with DEMs enable rapid
application of these indices on large areas. It must be acknowledged, however, that some indices are connected to a distinct spatial
scale and their application makes sense on large scales only. They serve, for example, for comparing drainage basin characteristics or
landform assemblages (e.g., drainage density, hypsometry, elevation relief ratio, a.o.). Other indices can be applied on several
scales, for example, terrain or surface roughness.

Data Sources
Digital Terrain Models
An abundance of gridded elevation data has been produced from different sources within the last decades. Extent and resolution of
such datasets have been growing with computing power and digital storage capacities. Modern geomorphological studies typically
4 GIS Applications in Geomorphology

Table 1 Frequently used indices in GIS applications in geomorphology (collected from various sources, see references in table, online resource:
http://gis4geomorphology.com/)

Index Description Formula, units

Channel sinuosity The Sinuosity index (SI) describes the ratio of the sinuous length SI ¼ Lc/Lv
(measured down the centerline of the channel) to the straight- where:
line distance of a reach. The sinuous length is divided by the SI ¼ channel sinuosity
straight-line distance A sinuosity of 1 describes a completely Lc ¼ length of channel
straight channel. Ratios around 1.5 refer to sinuous channels, Lv ¼ length of valley
while channels with higher ratios are considered meandering
channels (Burbank and Anderson, 2011)
Drainage density Drainage density (DD) of a catchment is the total line length of the DD ¼ sum(L)/A
stream network divided by catchment area. High density values where:
potentially reveal maturity of the channel system, rapid surface DD ¼ drainage density
runoff and low infiltration rates, or thin vegetation cover L ¼ length of channel
(Horton, 1932) A ¼ basin area
Hypsometry, Hypsometry is a measure of the relationship between elevation HI ¼ (Emean  Emin)/(Emax  Emin)
hypsometric index, and area in a catchment. Catchment hypsometry may reveal where:
hypsometric curves local flood response and erosional maturity HI ¼ hypsometric integral
The hypsometric integral (HI) expresses the elevation/relief ratio E ¼ elevation
and is often used as an estimate of the erosional development of
a catchment. Strahler (1952) described HI values of <0.30 as
“tectonically stable” or “mature” basins whereas HI values
>0.60 indicate “actively uplifting” or “young” basins. Interme-
diate or straight hypsometric curves (HI 0.50) suggest a
relatively stable landscape. Please note that complex interplay
of climatic, tectonic factors, as well as sedimentation and rock
resistance may produce similarly shaped curves. The HI may
hence be somewhat ambiguous and needs to be evaluated
carefully
Hypsometric curves plot normalized relief against the catch-
ment’s normalized cumulative area. The curve’s shape may
reveal the dominant geomorphic processes in the catchment
(diffusive/fluvial). Convexity indicates a larger portion of the
catchment’s area (volume of rock and soil) in the higher ele-
vated areas of the catchment where diffusive hillslope pro-
cesses dominate. Concavity implies a larger portion of the
catchment’s area at lower elevation and more channelized,
linear, fluvial or alluvial processes.
Hypsometric curves and integrals where first introduced by
Strahler (1952) and are commonly calculated for catchments.
If the HI is calculated using a regular kernel (e.g., a 3  3 cell
window) it is referred to as the elevation relief ratio (ERR)
Mountain front Mountain front sinuosity is a classic index of tectonic activity, Smf ¼ Lmf/Ls
sinuosity based on the notion that straight mountain fronts tend to lie where:
along active faults (Burbank and Anderson, 2011) Smf ¼ mountain front sinuosity
Lmf ¼ sinuous length measured along a path at the
break of mountain slope and alluvial fan
Ls ¼ length of the mountain front segment
(straight line)
Relief ratio Catchment relief (km) divided by catchment length (km) [km/km]
Stream frequency Stream frequency (F) counts all stream segments per unit area of F ¼ N/A
a catchment to describe the stream network’s texture, strongly where:
governed by bedrock and surficial material properties (strength, F ¼ stream frequency (total number of channels
fracture density, infiltration, mass wasting tendencies) per unit area)
N ¼ number of channels of all stream orders
A ¼ catchment area
Terrain or surface (A) Relative topographic position (also: topographic position (A) (“DEMsmooth”  “DEMmin”)/(“DEMmax”  “DEMmin”)
roughness index) estimates terrain ruggedness and serves as an index for where:
(ruggedness) local elevation. Topographic position of each pixel is a relative DEMsmooth ¼ smoothed elevation raster
metric based on its local neighborhood (10  10 pixels)
Used to identify landscape patterns corresponding to DEMmin ¼ minimum elevation raster
environmental (geomorphic, vegetational, etc.) factors. DEM max ¼ maximum elevation raster
Applicable to bathymetric data (B) (“DEMmean”  “DEM”)/“DEMrange”
(B) Standard deviation of elevation is a statistic measure of where:
topographic roughness DEMmean ¼ mean elevation raster
GIS Applications in Geomorphology 5

Table 1 (Continued)

Index Description Formula, units

(C) Slope variability calculates the slope relief based on a slope DEMrange ¼ raster containing range of elevation values
raster and its local cell neighborhood (e.g., >100 m) DEM (original elevation raster)
(D) Basin-scale Ruggedness (Rb) compares the relief of (C) SV ¼ “Smax”  “Smin”
catchments using streamlines and basin boundary polygons where:
(E) Standard deviation of residual topography compares ratio of SV ¼ slope variability
surface height and averaged surface on a local cell neighborhood Smax ¼ maximum slope value raster
(Grohmann et al., 2011) Smin ¼ minimum slope value raster
(F) Standard deviation of slope (Smith, 2014) (D) Rb ¼ A/DD
where:
Rb ¼ catchment ruggedness index
A ¼ area of the catchment
DD ¼ drainage density (calculated from segmented
streamline)
(see also: Drainage density)
STD (DEM  DEM average)
Averaged surface can be low pass filtered, for example
STD (Slope)
Valley width-to-height Valley width-to-height ratio (Vf) compares erosional patterns Vf ¼2Vfw/[(Eld  Esc) + (Erd  Esc)]
ratio between catchments (one Vf value per catchment) based on where:
values derived from a DEM or aerial photos along a single Erd ¼ elevation of the river-right valley divide (ridgeline)
cross-section per catchment. Vf was originally used to Eld ¼ elevation of the river-left valley divide (ridgeline)
distinguish V-shaped valleys (low Vf values, often close to 0) Esc ¼ elevation of the valley floor (canyon)
from U-shaped valleys (higher Vf values) (Burbank and Vfw ¼ width of valley floor
Anderson, 2011)
Glaciality index The glaciality index (GI) measures the concavity of a valley flank y ¼ axb
based on the idea that glacial valleys have a parabolic cross- b ¼ power-law exponent
section. It is defined as the exponent of a power-law fitted to the
valley flank, after Svensson (1959) and was calculated auto-
matically by Prasicek et al. (2015). A GI of 1 represents a fluvial
valley and progressively higher exponents indicate progres-
sively more U-shaped valleys
Steepness index The steepness index (ks) is the factor of a power-law describing @H
ks ¼ Ay
(stream power) the drainage area—channel slope relation after Flint (1974). @L
Drainage area can be combined with channel slope to derive the ks ¼ steepness index
stream power or steepness index, a simple metric for the ability A ¼ drainage area
of a stream to incise into bedrock (Flint, 1974) y ¼ concavity index
It should be emphasized that channel slope represents elevation ¼ channel slope
change over flow path length and hence differs from the topo-
graphic gradient as calculated in most GIS. The exponent y
depicts the shape of the power-law that describes the relation
between channel slope and A. This relation can only be
described by a single power-law (i.e., uniform stream power
regardless of location) if (i) influencing factors such as climate
and rock type are homogenous and (ii) the topography is steady
over time (Whipple and Tucker, 1999). If y is known and fixed,
ks becomes ksn, the normalized steepness index, and can be
used to identify a change in stream power and hence deviations
from the spatial and temporal constraints mentioned
previously. y generally ranges between 0.25 and 0.7 (Tucker
and Whipple, 2002; Whipple, 2004; Whipple et al., 2013), but
most studies use a reference value of 0.5 (Hack, 1957) or 0.45
(Whipple et al., 2013)
Gradient index The gradient index (SL) (Hack, 1973) allows the identification of @H
SL ¼ L
breaks in channel geometry based on the assumption that @L
channel length L is related to channel slope (Hack, 1957) SL ¼ gradient index@H=@L ¼ channel slope
L ¼ channel length from the divide
 
Topographic Wetness A parameter describing the tendency of a location to accumulate A
TWI ¼ ln
index (TWI) water. TWI was developed within the hydrological runoff model tanb
TOPMODEL (Beven and Kirkby, 1979) and applied in studies on where:
soil moisture, soil chemistry or species distribution analysis A ¼ upslope area draining through a certain point per
(Marthews et al., 2015) unit contour
length
b ¼ local slope

(Continued)
6 GIS Applications in Geomorphology

Table 1 (Continued)

Index Description Formula, units


 
Connectivity index (IC) The connectivity index (IC) focuses on the influence of Dup
IC ¼ log 10
topography on sediment flux. It is intended to represent the Ddn
pffiffiffi
linkage between different parts of the catchment and aims, in S
Dup: upslope component Dup ¼ W  A
particular, at evaluating the potential connection between Ddn: downslope component
hillslopes and features of interest like channels, sinks and X di
sediment storage landforms. (Cavalli et al., 2013) Ddn ¼
i
Wi Si
W ¼ is a weighting factor (based on roughness)
S ¼ average slope gradient of the upslope/downslope
contributing area
A ¼ contributing area
d ¼ length of the per cell flow path in steepest
downslope direction

employ DEMs with a resolution between 1 and 90 m. Low-resolution DEMs (cell size  30 m) are widely used for large-scale
analyses of landscape evolution. The extensive, often global coverage of low-resolution DEMs allows large-scale analyses and
comparisons between study areas worldwide. Medium to high resolution datasets (cell size < 30 and  1 m) are typically national
grids with a more limited extent and are a good choice for regional modeling of different LSPs. Submeter resolutions are mostly
produced by individual campaigns and spatially limited to single catchments or landscape patches. Such data sets are inevitable for
detailed analyses of weathering processes, soil erosion, and rock wall retreat.
Acquisition techniques vary and comprise active (radar, LiDAR) and passive (optical) remote sensing. While terrestrial and
airborne LiDAR dominated the acquisition of high-resolution elevation models over the last two decades, photogrammetry has
experienced a renaissance due to affordable drone technology and very high-resolution DEMs from SFM techniques.
The most widely used global DTMs with a resolution < 90 m are the elevation data of the shuttle radar topography mission
(SRTM) (Farr et al., 2007), available from 60 degree north to 60 degree south with a resolution of 1 arc second (approximately 30 m
at the equator) and the ASTER GDEM (Gesch et al., 2012), a global DSM derived from satellite imagery via stereo photogrammetry
and available from 83 degree north to 83 degree south with a resolution of 1 arc second. Furthermore, HydroSHEDS (Hydrological
data and maps based on Shuttle Elevation Derivatives at multiple Scales) (Lehner et al., 2008) are derived from different SRTM
versions and provide hydrographic information (e.g., imprinted river networks) along with elevation data at a resolution of 3 arc
seconds. The radar interferometry-based WorldDEM (http://www.intelligence-airbusds.com/worlddem/) developed by the
TanDEM-X Mission by the German Aerospace Center was just completed in October 2016 and for the first time provides a global
terrain model with a resolution of 12 m (Zink et al., 2006, 2011).

Optical Imagery
Optical satellite imagery
Geomorphologists have been using remotely sensed imagery since it became available during the first half of the 20th century. Carl
Troll was one of the first physical geographers who systematically used aerial imagery for the emerging field of geomorphology
(Lautensach, 1959).
While conventional aerial photography is still widely applied for local studies, satellite remote sensing has become a useful tool
when looking at larger areas. With the start of the first Landsat satellite in 1972, conventional earth observation from space entered
a new era. It was now possible to survey large areas continuously from space, revisiting the same locality in only 18 days’ time. Since
then, the lower earth orbit (160–2000 km above ground) has become packed with satellites from different agencies. In addition to
the large fleet of satellites launched by the US National Aeronautic and Space Administration (NASA) since the 1970s, several other
national space agencies and private companies launched their own earth observation missions, for example, the SPOT 1-7 satellites
(launched between 1986 and 2014) by CNES (France), the IRS family of satellites (launched between 1988 and 1996) operated by
ISRO (India), the World-View 1–4 satellites (Digital Globe) (launched between 2007 and 2016), and Sentinel 1–3 satellites by ESA
(launched between 2015 and 2017), to name only a few.
With the ongoing development of new sensors (and the launch of new satellites), spatial ground resolution of optical imagery
has been enhanced very quickly. While the first Landsat satellites had a maximum ground resolution of 60 m (resampled), the
WorldView-3 satellite’s panchromatic channel is collecting images with a ground resolution of 31 cm, a level of detail that only
aerial surveys could achieve before. Services like GoogleEarth or BingMaps make these high-resolution images available for visual
interpretation and comparison of different time steps.
For geomorphological applications, the temporal resolution of satellite images might be at least as important as the ever-increasing
spatial resolution of newly launched missions. For the detection of geomorphological changes, both long time series of satellite
imagery and short revisit times of the same locality are of vital importance. Increasing computational power but also political
GIS Applications in Geomorphology 7

2500
Keyword search in Web of Science

600
Landsat (n = 626)
Remote sensing (n = 2111)

2000
Absolute number of publications

Absolute number of publications


Photogrammetry (n = 324)

500
Structure from motion (n = 217)
400

1500
Landsat archive opened
300

1000
200

500
100
0

0
1980 1990 2000 2010
Year of publication
Fig. 2 Web of Science search results for different keywords. Results show absolute number of publications that have the keyword listed as topic.
Search results were refined by Web of Science categories: Geosciences multidisciplinary, geography physical, environmental sciences and geology. Data
from webofknowledge.com.

decisions during the past decade have boosted the use of satellite remote sensing products in geosciences (Wulder and Coops, 2014).
Since a change in data policy in 2008, millions of images acquired by the Landsat family of satellites between 1972 and today have
become available at no cost. The Landsat 8 satellite, launched in 2013 to ensure data continuity with nearly the same specifications as
before, collects several hundred images each day, revisiting the same location every 16 days. The free distribution of these images via
the USGS EarthExplorer (https://earthexplorer.usgs.gov) has spurred the use of satellite products in geomorphological research,
enabling the detection of geomorphological change from space over more than four decades. Shortly after 2008, geoscientific studies
making use of freely available Landsat imagery sharply increased (Fig. 2). Landsat imagery is not the only high-resolution imagery
available at no cost. In 2016, data recorded by the Advanced Spaceborne Thermal Emission and Reflection Radiometer (ASTER), a
sensor onboard the Terra satellite (sensor resolution: 15–90 m), became freely available as well (https://asterweb.jpl.nasa.gov/). Since
2014, the Sentinel satellite imagery from the European Space Agency (ESA) has been distributed via the ESA Science Hub (https://
scihub.copernicus.eu) at no cost. Launched in 2015, the Sentinel-2A satellite already collects optical imagery at 10 m ground
resolution with a revisit interval of 10 days at the equator, delivering comparable ground resolution and spectral specifications as
Landsat and ASTER imagery. Together with the Sentinel-2B satellite (launched in early 2017; phased at 180 degree with Sentinel-2A),
the Sentinel-2 mission will deliver optical imagery with a revisit time of 5 days at the equator (2–3 days in mid-latitudes). Alongside
with the remote sensing products, ESA offers a stand-alone desktop Sentinel-2 Toolbox that features a set of processing and
visualization tools for Sentinel-2 imagery, but also for other ESA and third-party remote sensing data.
The potential for geomorphological research to gain knowledge through the enormous amount of remote sensing products is
huge. One of the most prominent applications in geomorphology is the detection of glacial changes from space (Kääb et al., 2016;
Paul et al., 2015), but multitemporal satellite imagery has also been applied to landslide studies (Scaioni et al., 2014; Stumpf et al.,
2017), fluvial geomorphology (Legleiter and Fonstad, 2012; Rowland et al., 2016), and coastal erosion (Hara et al., 2015; Li and
Damen, 2010). It is beyond the scope of this article to give a full overview of potential applications; the following example shall
merely highlight the potential of satellite imagery to (a) pinpoint geomorphological change in remote regions and (b) to determine
rates of certain processes from space.
Despite potential distortion by intensive cloud cover, the high spatio-temporal resolution of Sentinel-2A makes it possible to
detect geomorphological change in remote areas. In early 2016, a rock avalanche detached from the northeastern flank of the Cerro
Alto San Juan, situated on the border between Argentina and Chile (Fig. 3). The mass detached at around 5200 m and dropped
onto the large glacier descending from the massif. The rock avalanche deposit covers an area of roughly 1 km2 (>2000 m length;
500 m width). Repeated Sentinel-2A imagery narrows the detachment down to the time span between 22 January 2016 and
04 February 2016. Comparison with an image obtained on 26 January 2017 shows how the debris is transported 50–100 m on top
of the moving glacier.

Unmanned aerial vehicles and structure from motion


Regarding the ground resolution, conventional aerial photography still outperforms satellite imagery by far. The major downside of
conventional aerial photography, however, is the cost-intensive data acquisition from small planes or helicopters. Technological
8 GIS Applications in Geomorphology

69°48‘W 69°47W 69°48‘W 69°47W

(B) (C)

33°27‘S

0 1000 2000 m

(A)
33°26‘S

Panel B & C
33°28‘S

Cerro Alto San Juan

0 1000 2000 m

69°49‘W 69°47‘W 69°45‘W


Fig. 3 Observation of geomorphological processes in remote regions: Multitemporal Sentinel 2A optical imagery (10 m ground resolution) of a rock
avalanche at the Cerro Alto San Juan massif on the border of Argentina and Chile. (A) Image acquired on 26 January 2017 showing the rock avalanche in
very clear conditions. (B) Image acquired on 22 January 2016, shortly before the rock avalanche detached. (C) Image acquired on 04 February 2016
showing the freshly deposited material on the glacier. Image courtesy of ESA.

advances during the past decade revitalized the use of aerial photography in geosciences. Unmanned aerial vehicles (UAVs), also
referred to as drones or multicopters, are available at low cost and make data acquisition cheap and relatively easy. Especially in
remote areas and in steep and rugged terrain, where shadowing effects limit the use of conventional aerial photography, the
acquisition of high-resolution aerial photography from UAV platforms offers new possibilities. Furthermore, repeated surveys
make geomorphological change detection also feasible on small scales that remote sensing data from satellites cannot resolve in
comparable detail.
Along with the advances in data acquisition, newly developed image processing software and algorithms offer new opportu-
nities for geoscientific research. Digital photogrammetry, grounding on the same principles as classical photogrammetry, is a
powerful method for extracting digital topography from overlapping images (Baltsavias et al., 2001; Keutterling and Thomas, 2006;
Lane et al., 2000). In the last couple of years, SFM, a new low-budget photogrammetric method to create DEMs from overlapping
images, has caught the attention of the geoscience community (Fig. 1). As opposed to classical photogrammetry, picture geometry,
orientation, and position of the cameras toward the object of interest are not a prerequisite. SFM software calculates these
parameters by matching common features in a set of overlapping digital images (Snavely et al., 2008). For a comprehensive
summary of SFM tools and their applications in geosciences, see Westoby et al. (2012).

Other Data Sources


GIS offers numerous possibilities to combine data sets of different sources for analysis and visualization. Besides DEM and image
data, all kinds of field and lab data can be imported into GIS for analysis. Data on surface material characteristics produced by
GIS Applications in Geomorphology 9

sampling, coring, or near-surface geophysics, mapping data, as well as measurements of process rates and lab data, for example,
sediment analysis or dating information, can be combined with digital land surface data. A prerequisite for the combination of
field/lab and digital data is correct geopositioning. Global positioning systems (GPSs) have become a standard requisite for field
work. Modern GPS receivers combined with correction signals transferred via mobile communications deliver high-accuracy
positioning data. Additionally, low-resolution positioning is available in every smartphone, which appears almost ubiquitous
today. Great potential lies in the combination of surface and subsurface information. Especially with the commonly applied
techniques of near-surface geophysics, landforms can be studied in three dimensions. The most frequently used techniques in
geomorphology are ground penetrating radar, seismic methods, resistivity and EM methods, and gravity methods (Kruse, 2013;
Schrott and Sass, 2008). Since most geophysical systems have a distinct data format and many methods deliver data along a survey
line, transfer of subsurface information often requires a conversion of, for example, depth information into point or line data,
before further processing in GIS software is possible.

Digital Geomorphological Mapping


Map Creation
Geomorphological mapping is a fundamental tool for geomorphologists. Geomorphological maps are highly complex thematic
maps that contain various different layers of geomorphological information. Digital map creation requires good map design and
sophisticated graphical tools to produce a readable and understandable map. The developments of graphical and analytical
functions in GIS software provide numerous valuable tools that facilitate map creation and distribution. Compared to analog
methods of map creation, the application of GIS software tools represents a significant simplification of the production process and
an important reduction of creation time and production costs.
Within GIS software, manipulation and analysis of various types of geomorphological information, for example, delineation,
measurement, mathematical operations and others, and the design and production of the mapare possible. Furthermore, the
logical storage structure of geomorphological data enables rapid production of derivative maps with special thematic focus, like
process domains, surface processes, surface material, or other. Geomorphological maps are created using either data gathered
during field campaigns and/or data extracted from digital data sources like aerial photography, satellite imagery, and DEMs. Field
mapping is significantly enhanced using mobile devices like tablets or handheld computers connected with GPS. Field mapping
software, usually a GIS-type software, enables direct collection of observations into a georeferenced database system that can later
be transferred to the desktop GIS used for map creation (Gustavsson et al., 2008; Minar et al., 2005). The mapping process can be
performed manually, automated, or semiautomated. Manual mapping relies on the experience and competence of the mapper
using visual heuristics to identify landforms of interest. The method is simple and rapid to deploy, and accuracy is generally high.
Automated or semiautomated mapping allows generation of more objective and repeatable information but usually falls behind in
accuracy compared to manual approaches. Corresponding methods rely on feature extraction techniques applied to satellite/aerial
imagery or different types of DEMs and their derivatives (see Section “Automated land surface classification”). The representation
of a landform on an image is dependent upon (i) the landform itself, (ii) the data source, and (iii) the visualization method (Otto
and Smith, 2013; Smith, 2011). Smith and Wise (2007) identified three main controls on the representation of landforms on
images: (i) relative size: the size of the landform relative to the spatial resolution, (ii) azimuth biasing: the orientation of the
landform with respect to solar azimuth, and (iii) landform signal strength: the tonal/textural differentiation of the landform.
Consequently, the relative reflectance of the landform in relation to surrounding features determines the detectability of a
landform. DEMs are applied using derivatives of elevation that provide various forms of visualization of inherent information
including relief shading, gradient (slope angle), or curvature classification.
The complex content of geomorphological maps is depicted using compound and often illustrative symbols (Otto et al., 2011).
GIS software provides tools for the creation of custom symbols representing geomorphological features and functions for
cartographic design and map production. Digitally produced maps are easily distributed in various formats ranging from print
maps to online web services, making full use of the data organization structure and the georeferencing of the data (Smith et al.,
2013). Additionally, the standard PDF (Portable Document Format) has been extended into a GeoPDF for display and dissemi-
nation of referenced map data. Geospatial functionality of a GeoPDF includes scalable map display, layer visibility control, access
to attribute data, coordinate queries, and spatial measurements (www.terragotech.com).

Automated Land Surface Classification


General land surface classification
The extraction of discrete entities from a continuous digital image has been a main research field within geomorphometry for
decades. Evans (2012) stresses the need to distinguish between landform and land surface form. The difference between the two
arises from the discontinuity or continuity of the feature, respectively (referring to the concept of general and specific geomorpho-
metry, see the previous discussion). Tasks related to the description of the continuous land surface are here addressed as general
land surface classification, while discontinuous landforms will be treated in the following chapter on specific land surface
classification. Theoretical frameworks for general land surface classification have been developed by several authors, mostly
addressing basic LSPs, such as elevation, slope, aspect or curvature. Based on these characteristics the continuous land surface is
10 GIS Applications in Geomorphology

split into discrete parts termed land surface elements. Their main characteristic is geometric homogeneity (Minár and Evans, 2008).
Dikau (1989), for example, proposed nine landform elements, defined by their profile and plan curvature, to represent the
building blocks of the land surface (Fig. 4). MacMillan et al. (2000) and Schmidt and Hewitt (2004) both used plan and profile
curvature, slope and slope position for automated segmentation of landforms into landform elements based on DEMs, heuristic
rules and fuzzy logic. They extend the model of Dikau (1989) by nine elements with classes for ridges, peaks, valleys, spurs, terraces,
hollows, plains, saddles, and slope position. Shary et al. (2005) presented 12 slope types, while Minár and Evans (2008) even
distinguished between 25 elementary landforms. The main function of these elementary forms lies in their relation with process
dynamics rather than the delineation of discrete real landforms. Curvature changes evoke an acceleration or deceleration of gravity
flow and result in dispersion or concentration of transported matter (Minár and Evans, 2008). However, general land surface
classification can be used as starting point for specific classification approaches (Drăguţ and Blaschke, 2006).

(A) Profile curvature

Convex Profile straight Concave


Concave

X/X SF/X V/X


Plan curvature

Plan straight

X/SL SF/SL V/SL


Convex

X/V SF/V V/V

12°16¢0²E

(B) Relief Class.—Dikau


SF/SL—flat
X/X
SF/X
V/X
X/SL
SF/SL
V/SL
X/V
SF/V
V/V
Lakes
Glaciers
47°8¢0²N

0 0.25 0.5 1
km

Fig. 4 (A) Relief classification using the approach by Dikau (1989). (B) Basic landform elements based on curvature conditions in two directions.
Based on Dikau, R. 1989. The application of a digital relief model to landform analysis in geomorphology. In: Raper, J. F. (ed.) Three dimensional
applications in geographical information systems. London: Taylor & Francis.
GIS Applications in Geomorphology 11

Specific land surface classification


Specific landform classification ambitiously aims at combining single pixels or landform elements to landforms as perceived by
experts and hence is somewhat more subjective and adds a lot of additional complexity (Hengl and Reuter, 2009). Attempts to
classify entire scenes or at least specific landform types have only partially been successful. This is mostly owed to the exceptional
complexity immanent in Earth’s landscapes. The surface of real-world landscapes or even small parts of it can never be exactly
matched by mathematical representations of landforms. As a consequence, variations in the appearance of landforms have to be
considered and measures for similarity have to be applied. Beyond these technical issues, different geomorphological processes
may produce similar landforms, a mechanism that is known as “equifinality” and considerably hampers the interpretation of
landscape shape.
Analysis of remote sensing data including DEMs as well as aerial and satellite imagery is mainly performed in a pixel-based
manner. Each pixel is treated separately and only a few attributes are at hand for characterization and classification. The adoption of
image segmentation and object classification techniques from computer science to GIS applications (Blaschke and Strobl, 2001)
allowed overcoming these limitations which can be particularly useful for the automated interpretation of complex landforms.
Image segments are regions which are automatically merged from pixels referring to one or more criteria of homogeneity in one or
more dimensions of feature space (e.g., spectral reflectance). The resulting objects can be described and classified using additional
spectral (e.g., mean, median, variance), geometrical (e.g., circularity), textural, and hierarchical information (Blaschke, 2010).
Schneevoigt et al. (2008) combined ASTER satellite imagery with DEM data to detect 20 different alpine landform types. They
report an overall accuracy of 92% with good results for talus slopes, free rock faces and cirque walls but detection problems for
fluvial, glacial, and debris flow deposits. Among the most successful approaches for a gapless classification of a landscape into
individual landforms, Anders et al. (2011) employed an object-based approach to extract karst, glacial, fluvial, and denudation
landforms from a high-resolution DTM and reached an overall accuracy of about 70%. Focusing on specific landforms only,
d’Oleire-Oltmanns et al. (2013) and Eisank et al. (2014) reached accuracies around 60% when attempting to automatically map
gullies and drumlins. Object-based image analysis is a promising technique for landform classification but clearly further research
is needed to handle the exceptional complexity of real-world landscapes in automated routines of landform mapping.

Application of Various Geomorphological Indices for Process and Landform Analysis—Case Study Obersulzbach
valley, Eastern Alps, Austria

Instead of reviewing the vast number of studies that apply GIS in geomorphological research, we now present the application of
selected topographic indices and geomorphological analyses for a local test site in the Eastern European Alps. The Obersulzbach
valley is located north of the main divide in the Hohe Tauern range in Austria and represents a southern tributary of the Salzach
river (Fig. 5). It covers an area of 81 km2 of complex high alpine relief between 850 and 3657 m. Glacial imprint dominates surface
morphology and is manifested by numerous cirques in the upper reaches and pronounced U-shaped sections along the main
channel of the valley. The longitudinal profile of the main valley shows two pronounced steps, one located about 4 km into the
valley and the other located 6 km from the valley head (Fig. 6). The first step separates the deeply incised, rather V-shaped lower
level that connects to the Salzach valley from a U-shaped upper level. The upper step separates the U-shaped section from the
cirque-like valley head area. Glaciers are still present in some of the cirques and cover large parts of the valley head, summing up to
17% of glacier coverage in total. The Obersulzbachkees glacier at the valley head covered an area of 15 km2 in the Little Ice Age and
is now split into several glacier parts of 9 km2 area in total (last glacier extent from 2009, data by: Fischer et al., 2015). A large
proglacial lake has formed in the valley (second pronounced step) since the late 1990s.
The calculation of the geomorphological indices is based on a DEM with cell size of 10 m (provided by data.gv.at under the
INSPIRE framework). We used ArcGIS, SAGA GIS, and TAUDEM for the analysis and ArcGIS for visualization. The results can be
accessed also as the WEBGIS service in the online supplementary at the publisher’s homepage.

Hillslopes and Gravitational Processes


Process dynamics are governed by surface topography, impacting on energy potentials, for example, by intensification of
attenuation of friction or diffusion, or concentration of flow. In absence of convergent flow and moving ice, hillslope processes
are mostly controlled by surface slope and curvature. Erosion is most commonly modeled by hillslope diffusion where sediment
transport depends only on slope and a diffusivity constant (Montgomery and Foufoula-Georgiou, 1993; Dadson and Church,
2005; Egholm et al., 2012; Tucker and Bras, 1998; Kirkby, 1987). This makes basic LSPs such as slope and curvature very valuable
for evaluating hillslope processes.
Surface slope controls how gravitational acceleration is split into retentive and detaching forces and profile curvature holds
information on how this relation changes. Convexity indicates areas of increasing slope, increased sediment transport capacity and
hence erosion, while concavity represents reduced transport capacity which leads to deposition of material.
The length and relief of the hillslopes holds information on forcing at a range of scales (e.g., Grieve et al., 2016) and plays an
important role in the universal soil loss equation for estimating soil erosion (e.g., Hickey et al., 1994; Moore and Burch, 1986; Liu
et al., 2000) and in slope susceptibility for shallow landslides (e.g., Carrara, 1983; Gómez and Kavzoglu, 2005). Furthermore,
following Salcher et al. (2014), glacial landscapes tend to show longer hillslopes and greater hillslope relief than fluvial ones. For our
12 GIS Applications in Geomorphology

9˚0⬘0⬙E 10˚0⬘0⬙E 11˚0⬘0⬙E 12˚0⬘0⬙E 13˚0⬘0⬙E 12˚12⬘0⬙E 12˚14⬘0⬙E 12˚16⬘0⬙E 12˚18⬘0⬙E 12˚20⬘0⬙E
49˚0⬘0⬙N (A) (B)

47˚14⬘0⬙N
48˚0⬘0⬙N
47˚0⬘0⬙N

47˚12⬘0⬙N
46˚0⬘0⬙N

47˚10⬘0⬙N
47˚8⬘0⬙N
47˚6⬘0⬙N

0 1 2 4
km

Fig. 5 (A) Location of the study area. (B) Aerial image of the Obersulzbach valley, Eastern European Alps, Austria. Free orthofoto data from basemap.at.

0.45
2500
0.40

0.35
2000
Channel slope ( )

0.30
Elevation (m)

1500 0.25

0.20
1000
0.15

0.10
500
0.05

0 0.00
0 5000 10,000 15,000 20,000 25,000
Distance (m)
Fig. 6 Longitudinal profile of the Sulzbach Creek draining the Obersulzbach Valley. Note the changes in channel slope along the path and the distinct
steps along the profile (cf. text for details).
GIS Applications in Geomorphology 13

12°16¢0²E

Hillslope length (m)


0–150
150–300
300–450
450–600
600–750
750–900
900–1,050
1,050–1,200
1,200–1,350
1,350–1,500
1,500–1,650
Lakes
47°8¢0²N

Glaciers

0 0.25 0.5 1
km

Fig. 7 Hillslope length index in the study area.

study area, we calculated both hillslope length and relief with TauDEM (Tarboton et al., 1991) and observe this trend only in distinct
cirques on the western valley flank. However, less dissected terrain can generally be expected to have longer hillslopes (Fig. 7).
Another example for assessing the impact of surface characteristics on hillslope processes is the surface roughness index.
Roughness, also termed ruggedness or microtopography, describes the local variability of elevation on a given scale. On a large
scale, roughness of a land surface is controlled by the number, size and distribution of landforms and is expressed by landform and
landform elements or breaklines. On small scales, roughness of a single landform results from material properties, processes acting
upon it, and the time since formation (Grohmann et al., 2011). In this case it signifies grain size of surficial deposits, or surface
smoothing due to erosion. The parameter is applied, for example, in fluvial geomorphology to characterize bed morphologies and
fluvial erosion, in hydrological modelling, in weathering studies, for landslide detection and for relative age dating (Smith, 2014).
In permafrost studies surface roughness controls energy transmission into the ground and fosters permafrost formation (Otto et al.,
2012). Numerous approaches exist to quantify surface roughness (see Table 1; Grohmann et al., 2011; Smith, 2014). We have
applied two ways to calculate roughness in the study site: (1) standard deviation (STD) of slope (Fig. 8A) and (2) STD of residual
topography (Fig. 8B). Both calculations use a focal statistics tool with a 3  3 cell analysis window. Residual topography results
from the difference between the surface height and a low pass filtered height. STD of slope describes sensitivity to changes in
curvature, while STD of residual topography represents changes in altitude and hence gradient. Both approaches are sensitive to
major breaklines and steps, especially at the cirque-main valley boundary (Fig. 8). They allow for a delineation of linear landforms,
like ridges, lateral moraines, or steep walls. They also depict differences in roughness between bedrock slope, for example, cirque
walls, and sediment covered cirque floors (see eastern valley flanks).

Glacier Environments
GIS play an important role in analyzing and visualizing glaciers and glacially sculpted landscapes (compare chapter on GIS for
Glaciers). GIS are used by glacial geomorphologists to integrate multisource data, manage multiscale studies, identify spatial and
temporal relationships and patterns in geomorphological data, and to link landform data with numerical models as part of model
calibration and verification (Napieralski et al., 2007).
When looking at the ice itself, mass balance constraints are inevitable to understand glacier development and GIS are ideal to
process and visualize related data. For example, GIS can be used to visualize manually mapped glacier extent or to even apply
automated mapping routines based on DEM and spectral data (Racoviteanu et al., 2008). If glacier extent is known, the altitude of
the equilibrium line (ELA), the virtual border between accumulation and ablation area that is central for mass balance assessment,
can be calculated by applying different methods such as the accumulation area ratio (AAR).
The ELAs for the glaciers of our study area range between 2400 and 3300 m (Fig. 9). The distribution of the glaciers in the
Obersulzbach valley already indicates the importance of climatic influences. While glaciers still exist on the slopes facing towards
east and north, western slopes only show remnants of past ice cover. The climatic influence is further indicated by the relation
14 GIS Applications in Geomorphology

12°16′0″E

(A)
Slope STD ( )
0.0–1.4
1.5–2.7
2.8–4.1
4.2–5.6
5.7–7.3
7.4–9.3
9.4–11.8
11.9–15.1
15.2–27.4
Lakes
Glaciers
47°8′0″N

0 0.25 0.5 1
km

12°16′0″E

(B)
Residual topography (m)
0–0.56
0.57–1.21
1.22–2.06
2.07–2.99
3–4.11
4.12–5.51
5.52–7.29
7.3–9.72
9.73–14.01
14.02–23.82

Lakes

Glaciers
47°8′0″N

0 0.25 0.5 1
km

Fig. 8 Surface roughness calculated using the standard deviation of local slope (A) and the residual topography approach (B) in the study area
(extract). See text for explanations.

between ELA and exposition. While northeast-facing glaciers feature the lowest ELAs, south-facing glaciers have the highest ELAs.
When applying the AAR method for ELA calculation, the mass balance of the reference glaciers needs to be considered. If these
glaciers were not fully adapted to the climatic conditions during ELA measurements, any deviation in mass balance is transferred to
the ELAs calculated with the AAR method.
Once the ELA is known, be it from field campaigns or GIS-based models, the glacier surface can be split into accumulation and
ablation area. Shear stress models can be used to derive ice thickness estimations and hence calculate ice volume and water
equivalent (Huss and Farinotti, 2012; Frey et al., 2013). Mass balance-based models can be applied to calculate ice flux and to
retrieve similar results. We applied the model GlabTOP2 developed by (Frey et al., 2013) to the current glacier extend of our study
GIS Applications in Geomorphology 15

12°16′0″E 12°20′0″E

ELA (m)
2400–2500
2500–2600
2600–2700
2700–2800
2800–2900
2900–3000
3000–3100
3100–3200
3200–3300

Lakes
47°8′0″N

0 0.5 1 2
km

Fig. 9 Equilibrium line altitudes of the glaciers of the Obersulzbach Valley.

site. The model calculated ice thickness from slope and shear stress relationships based on a simplified term presented by Haeberli
and Hölzle (1995). Ice thickness modeled ranged from less than 25 m for many small cirque glaciers to maximum values of more
than 200 m for the larger glaciers at the valley head (Fig. 10).
Glacially sculpted landscapes are also frequently analyzed using GIS, be it the morphometric characterization of moraines or
cirques, or automated mapping of glacial landscapes and glacial imprint based on estimations of cross-sectional valley shape or
hypsometric indices (e.g., Prasicek et al., 2015; Smith et al., 2006). Furthermore, a variety of GIS-based models has been applied to
assess the volume of glacial valley fill (e.g., Mey et al., 2016; Jaboyedoff and Derron, 2005a). The hypsometry of glaciated
landscapes has been used to predict the state of glacial landscape evolution and to distinguish between glacial and fluvial terrain
(Brocklehurst and Whipple, 2004; Sternai et al., 2011).
In our example, we quantify the U-shapedness of the valleys in our study area based on the glaciality index (GI), derived from
parabolas fitted to valley flanks for all flow path cells (Prasicek et al., 2014, 2015). To analyze the glaciers themselves we use
outlines mapped by Abermann et al. (2009). We calculate the approximate ELA of all glaciers in the study area employing an AAR of
0.6, which is in the center of the range of reported values (e.g., Porter, 1975; Benn and Lehmkuhl, 2000; Gross et al., 1977).
For calculating U-shapedness as a proxy for the degree of glacial imprint, we automatically determine valley width using a
multiscale curvature approach (Prasicek et al., 2014) and subsequently fit power-laws to the valley flanks (Prasicek et al., 2015).
16 GIS Applications in Geomorphology

12°16′0″E 12°20′0″E

Ice thickness (m)


5–25
25–50
50–75
75–100
100–125
125–150
150–175
175–200
200–225
225–250

Lakes
47°8′0″N

0 0.5 1 2
km

Fig. 10 Ice thickness modeling using the model GlabTOP2 applied to the glaciers in the study area (Frey et al., 2013).

These calculations are performed for flow path cells only and then averaged for a spatially continuous result. Results show that
deeply incised and dissected parts of the study area have a fluvial GI and hence rather straight valley flanks while the main trough
and distinct cirques show increased glacial imprint. However, the effect of valley fill producing flat valley floors as well as other
deposits on valley cross-sectional shape needs to be considered for interpretation (Fig. 11).

Periglacial Environments
Land surface is strongly interrelated with climate on a regional and local scale. The spatial differentiation of near-ground
atmospheric processes and climate variables is dominantly controlled by topography. This impact of topography on climate is
expressed by the term topoclimate or topoclimatology (Böhner and Antonić, 2009). In geomorphology, topoclimatic effects are
related to the spatial distribution and variable morphological characteristics of landforms. Especially landforms of process domains
sensitive to climatic influences such as glacial, periglacial and fluvial processes show interrelations with topoclimatic parameters,
such as aspect, altitude or slope angle (Olyphant, 1977; Allen, 1998; Humlum, 1998; Sattler et al., 2016). Within periglacial
geomorphology GIS have been used to model and visualize permafrost distribution since the early 1990s (Riseborough et al.,
2008). Two classic models originally applied in the Swiss Alps, namely PERMAKART (Keller, 1992) and PERMAMAP (Funk and
GIS Applications in Geomorphology 17

12°16′0″E 12°20′0″E

Glaciality index ( )

0.71–0.92
0.92–1.04
1.04–1.13
1.13–1.21
1.21–1.28
1.28–1.35
1.35–1.41
1.41–1.48
1.48–1.56
1.56–1.64
1.64–1.73
1.73–1.82
1.82–1.91
1.91–2.04
2.04–2.35
47°8′0″N

Lakes

Glaciers

0 0.5 1 2
km

Fig. 11 Glaciality index of U-shapedness in the study area.

Hoezle, 1992), can be considered as a starting point of GIS applications in permafrost research. The model PERMAKART is based
on empirical geomorphological evidence concerning permafrost occurrence (e.g., lower limit of active rock glacier, basal temper-
ature of snow) and integrates besides perennial snow avalanche deposits (protecting the ground surface from radiation) classic
topoclimatic parameters such as altitude and aspect. These LSPs are used as proxy information for air temperature and solar
radiation, respectively, two influential factors on the formation of mountain permafrost. Additionally, slope position and slope
angle are used representing the influence of snow cover on permafrost. The distribution modeling thus consists of a regionalization
approach based on these LSPs using a DEM and the empirical data on permafrost occurrence classified by altitude and aspect, the
so-called topoclimatic key. Over the years the original model structure has been modified and improved several times. One of the
latest GIS-based empirical permafrost models comprises a topoclimatic key of 24 different relief classes subdivided in eight classes
of aspect, each of them divided into three slope categories (rock, steep slopes and slope foot-positions, see Fig. 12) (Schrott et al.,
2013). The empirical model PERMAKART 3.0 has an index-based permafrost probability ranging from 1 (very unlikely) to
100 (very likely). This allows a more transient and realistic visualization, also because the relief class “rock” displays more
realistically lower permafrost probabilities even in higher altitudes (Fig. 13)
18 GIS Applications in Geomorphology

W E

3500

3000

Slope class Permafrost


2500 index
Bedrock < 50%
> 50%
2000 Steep slopes < 50%
> 50%
S
Foot slopes < 50%
> 50%

Fig. 12 Topoclimatic key used in the GIS-based permafrost distribution model PERMAKART 3.0 (Schrott et al., 2013).

12⬚16⬘0⬙E 12⬚20⬘0⬙E

Permafrost index
Probability
High : 100

Low : 0

Lakes
Glaciers
47⬚8⬘0⬙N

N
0 0.4 0.8 1.6
km

Fig. 13 Map of permafrost index in the Obersulzbach valley (extract). Note the varying index values in similar aspect situations (e.g., on the eastern
cirque walls) that result from differences in the topoclimatic key between rock slopes, steep slopes, and foot slopes.

Fluvial Environments
The investigation of fluvial environments is a key task in GIS applications, not only in the field of geomorphology but also hydrology
and ecology. This is mainly because river networks constitute the backbone of most humid and also semi-arid landscape types
worldwide. Even the course of present-day glacial valleys is determined by pre-glacial fluvial topography. Consequently, the structure
of river networks manifested in basins and watersheds is commonly used to partition the land surface and the fluvial catchment is the
standard unit for geomorphological and environmental analyses. In our example we provide an overview over the standard tools to
extract the river network and subsequently calculate a number of indices to further characterize the drainage system.
GIS Applications in Geomorphology 19

100

Slope ( )

10-1

10-5 10-4 10-3 10-2 10-1 100 101 102 103


2
Drainage area (km )
Fig. 14 Slope-area plot of the Obersulzbach Valley. Note the pronounced kink at a drainage area of approximately 0.1 km2.

First, sinks in the DEM should be filled, if a continuous drainage network is desired. The pit-filled DEM can then be used to
calculate flow direction and flow accumulation. Drainage area is one of the most important LSPs and it is typically derived from
gridded elevation data via standardized GIS routines. It is used as a simple and easily determined proxy for discharge. Thus,
drainage area is a major factor for assessing the erosive power of convergent flow and reveals the architecture of the drainage
network, the backbone of many landscape types worldwide.
A wealth of methods exists to determine drainage area from DEMs. The integration of drainage area in the direction of flow is
central to all of them but they differ in the way flow directions are determined. It is beyond the scope of this book section to discuss
the details of these approaches and the interested reader is referred to the work of, for example, Jenson and Domingue (1988),
Fairfield and Leymarie (1991), Freeman (1991), Tarboton et al. (1991), Costa-Cabral and Burges (1994), or Seibert and McGlynn
(2007). While in the classic procedure of Jenson and Domingue (1988) all flow is assumed to descend via the steepest path, cell
area is partitioned between two or more flow directions in most other approaches—a difference that needs consideration for
geomorphological applications. In single flow direction algorithms, drainage area cannot decrease downstream, which leads to the
formation of distinct flow paths. In contrast, multiple flow direction algorithms allow flow dispersion. Differences between the two
types of algorithms are most pronounced in divergent areas such as ridges, hillslopes and planar valley fills (Erskine et al., 2006).
Consequently, algorithms capable of flow dispersion should be used for detailed analyses of ridges and hillslopes, while single flow
direction algorithms are designed for the extraction of drainage networks, Strahler orders and other derivatives such as flow length
and river longitudinal profiles.
Once the per-cell upstream drainage area has been computed, drainage area–slope relations can be used to identify the drainage
area threshold between divergent and convergent flow, that is, the extent of the valley network, on a regional basis. Montgomery
and Foufoula-Georgiou (1993) showed that the extent of topographically divergent hillslopes, and thus the extent of the valley
network, corresponds to a change in sign of the relation between local slope and contributing drainage area. They further
demonstrated that debris flow-dominated channels can be determined from an inflection in the drainage area–slope relation.
In practice, a regional drainage area threshold for valley network initiation can be determined by establishing drainage area bins
and plotting mean slope against drainage area.
For our study area, such a plot is shown in Fig. 14. In our example, the slope–area relation shows a change in sign around
104 km2, at a scale where convergent terrain starts to establish and colluvial processes come into play. Further breaks are evident
around 102 and 101 km2 and we interpret the latter one to indicate the transition from colluvial to fluvial processes. We thus
chose this drainage area cutoff for river network extraction. The extracted drainage network acts as basis for the calculation of stream
orders (Horton, 1945; Strahler, 1952), subcatchments and drainage density. We calculated drainage density per square kilometer
(i) for the entire study area using a kernel with a size of 1 km2 (Fig. 15A), and (ii) for catchments of Strahler order 2 by dividing the
summarized length of all drainage lines in each catchment by catchment size (Fig. 15B). Results show that elevated and cirque-
shaped parts of the study area have a lower drainage density, probably due to more recent and/or more intensive activity of ice.
In addition to identifying the transition between divergent and convergent flow, drainage area–slope relations and flow
length–slope relations can be used to identify breaks in channel geometry and catchment evolution. Steepness index (ks) and
gradient index (SL) hold information on deviations from fundamental relations between drainage area and slope and flow length
and slope in a fluvial landscape, respectively. On glacially sculpted terrain deviations from fluvial topography can of course be
20 GIS Applications in Geomorphology

12°16′0″E 12°20′0″E

(A)
Drainage density (1/km)
0–0.5
0.5–1
1–1.5
1.5–2
2–2.5
2.5–3
3–3.5
3.5–4
4–4.5
4.5–5
Lakes
Glaciers
47°8′0″N

0 0.5 1 2
km

Fig. 15 (A) Drainage density calculated for the entire Obersulzbach Valley using a kernel of 1 km2. (B) Drainage density calculated for catchments of
Strahler order 2.
(Continued)

expected to be ubiquitous. Nevertheless, both indices clearly indicate disturbances in narrow valley sections where knickpoints with
comparatively steep channel sloped are located (Figs. 5 and 16).
Horton-Strahler orders (Horton, 1945; Strahler, 1957) can also be calculated from the extracted drainage network. Subse-
quently, subcatchments can be delineated for each Strahler order which allows the comparison of landscape patches with similar
drainage topology.
Hypsometry has been used to describe and interpret the topography of both fluvial and glacial landscapes (e.g., Brocklehurst
and Whipple, 2004; Strahler, 1952). The hypsometric integral (HI) can be used in both cases to assess landscape maturity.
However, if landscapes are influenced by both process domains, the interpretation of the HI gets difficult. We nevertheless calculate
the HI in our study area (Fig. 17A), once for Strahler order 2 catchments and once applying a kernel-based approach on the entire
Obersulzbach valley (elevation relief ratio—ERR). The size of the kernel is 1 km2. The ERR is high for convex landscape patches and
becomes lower with increasing concavity (Fig. 17B). The HI calculated for the irregularly shaped catchments of Strahler order
2 show a similar pattern with high HI values for hillslope-dominated catchments and low HI values for catchments that include a
considerable amount of valley floor (Fig. 17A). Both the steepness index and the gradient index can be applied to identify spatial
and/or temporal distortions of the drainage system and are hence very valuable tools for GIS-based geomorphological studies.
GIS Applications in Geomorphology 21

12°16′0″E 12°20′0″E

(B) Drainage density (1/km)


0–0.4
0.4–0.8
0.8–1.2
1.2–1.6
1.6–2
2–2.4
2.4–2.8
2.8–3.2
3.2–3.6
3.6–4
Lakes
Glaciers
47°8′0″N

N
0 0.5 1 2
km

Fig. 15, Cont’d

Sediment Flux and Erosion in Mountain Areas


GIS tools are also applied for the analysis and quantification of sediment flux. One focus in GIS-based sediment flux analysis is the
issue of connectivity. Coupling of geomorphological processes and catchment connectivity are central to the efficiency of sediment
flux and represent a significant impact on the sensitivity of geomorphological systems towards changes (Heckmann and
Schwanghart, 2013; Harvey, 2001). Various approaches exist to visualize and quantify catchment connectivity using GIS tools.
Indices of hydrological connectivity are presented by Borselli et al. (2008) and modified by Cavalli et al. (2013). They are based on
parameters like slope gradient, flow length, surface roughness and contributing area. The tools have been applied in various
environments and locations and used to discuss landform distribution and sediment flux (Gay et al., 2016; Lane et al., 2017;
Messenzehl et al., 2014). Heckmann and Schwanghart (2013) apply a different approach using network analysis and graph theory
to identify sediment cascades and delineate subcatchments of sediment flux. We have applied the connectivity index (IC) by Cavalli
et al. (2013) to our study area (Fig. 18). The IC index depicts highly variable connectivity conditions, a generic characteristic of high
mountain topography. Low connectivity areas (blue) are located in the floodplain and at footslopes. Zones of high connectivity
represent flowlines mainly in channels and gullies. The index visualizes, for example, how sediment production and storage in
cirques on the eastern valley side are connected, or disconnected, from the main valley and the dominant fluvial transport system of
the main river.
22 GIS Applications in Geomorphology

12°16′0″E 12°20′0″E

(A)
Steepness index (m)
0–11.3
11.3–16.8
16.8–28.1
28.1–51.3
51.3–98.9
98.9–196.7
196.7–397.5
397.5–809.9
809.9–1656.7
1656.7–3395.7
Lakes
Glaciers
47°8′0″N

0 0.5 1 2
km

12°16′0″E 12°20′0″E

(B)
Gradient index (m)
0–260.8
260.8–387.8
387.8–648.7
648.7–1184.3
1184.3–2284.2
2284.2–4542.9
4542.9–9181.1
9181.1–18,705.8
18,705.8–38,265.1
38,265.1–78,430.6
Lakes
Glaciers
47°8′0″N

0 0.5 1 2
km

Fig. 16 (A) Normalized steepness index (calculated a concavity index of 0.5) for the Obersulzbach Valley. (B) Gradient index for the Obersulzbach
Valley.
GIS Applications in Geomorphology 23

12°12′0″E 12°16′0″E 12°20′0″E

(A)
Elevation relief ratio ( )
0.07–0.14
0.14–0.21
0.21–0.28
0.28–0.35
0.35–0.42
0.42–0.49
0.49–0.56
47°12′0″N

0.56–0.63
0.63–0.7
Lakes
Glaciers

0 0.5 1 2
km

12°12′0″E 12°16′0″E 12°20′0″E

(B)
Hypsometric integral ( )
0.26–0.30
0.30–0.34
0.34–0.38
0.38–0.42
0.42–0.46
0.46–0.50
0.50–0.54
47°12′0″N

0.54–0.58
0.58–0.62
0.62–0.66
Lakes
Glaciers

0 0.5 1 2
km

Fig. 17 (A) ERR for the lower part of the Obersulzbach Valley. (B) HI for the lower part of the Obersulzbach Valley.
24 GIS Applications in Geomorphology

12°16′0″E

(A)
Connectivity index (IC)
–6.3 to –5.5
–5.4 to –4.7
–4.6 to –4.3
–4.2 to –4.0
–3.9 to –3.9
–3.8 to –3.6
–3.5 to –3.4
–3.3 to –3.0
–2.9 to –2.4
–2.3 to –1.4
Lakes
Glaciers
47°8′0″N

0 0.25 0.5 1
km

12°16′0″E

(B)
47°8′0″N

0 0.25 0.5 1
km

Fig. 18 (A) Connectivity index (IC) for the Obersulzbach Valley (part). The index displays distinct variation between valley floor locations and cirques.
The largest changes applying are present between channels and flats (based on IC tool by: Cavalli et al., 2013). (B) Ortho image of the same area. Data
from: basemap at, http://maps.wien.gv.at/basemap/1.0.0/WMTSCapabilities-arcmap.xml.
GIS Applications in Geomorphology 25

Several approaches have been applied for sediment flux quantification in GIS. Quantification of deposits is achieved by
combining surface and subsurface data on sediment depth, for example, from logging or geophysical surveying. In this approach
sediment volume results from the difference between the landform surface and the level of various sedimentary units or the
bedrock boundary. Since data from geophysical surveying or from logging often represents local information of single points or
along transects, interpolation of subsurface information is required and performed using GIS tools. The density of subsurface data,
the interpolation method applied as well as the resolution of the surface and subsurface data determines the accuracy of the
quantification result and needs to be considered. Due to the often-limited availability of detailed subsurface data due to high efforts
in both time and costs of field campaigns, these approaches usually generate information on a limited number of single landforms.
Examples for single landform quantification with GIS and subsequent calculation of sediment flux or erosion rates include glacial
valley fills, talus slopes, rock glaciers, or moraine deposits (Sass, 2007; Schrott and Adams, 2002). Sediment budgets of entire
catchments or valley fill deposits have been generated based on interpolation of dated sedimentary layers from cores (Tunnicliffe
et al., 2012), on geophysical data (Otto et al., 2009; Hinderer, 2001), or a combination of both (Götz et al., 2013).
An alternative solution to quantify sediment volumes is achieved by approximating the three-dimensional shape of the deposit
or the bedrock boundary using geometric models or mathematical functions. Simple geometric shapes or power-law functions are
applied for the representation of the volumetric body of landforms or the bedrock boundary, respectively (Hoffmann and Schrott,
2002; Schrott et al., 2003a). For single landforms sediment volumes of deposits or erosional features, for example, gullies, fans or
talus slopes, are approximated using geometrical shapes, for example, cone sectors or prisms, that represent the landform. Volumes
are quantified using the outline dimensions of the landform (height, length, width, depth) and the respective mathematical term
corresponding to the geometry type (Campbell and Church, 2003; Curry, 1999; Shroder et al., 1999). Valley cross-sections and
valley fill deposits of formerly glaciated terrain have been quantified using power law or polynomial functions based on the
assumption that bedrock topography sculpted by glaciers can be described mathematically (Harbor and Wheeler, 1992; James,
1996). These mathematical approximations are applied to cross-sections and the results are interpolated in order to quantify entire
valley sections (Schrott et al., 2003a; Jaboyedoff and Derron, 2005b).
Finally, GIS tools are applied to quantify erosion and deposition using DEM data of different points in time. Calculating DEM
of differences between two surfaces results in volumetric surface changes that can be associated with geomorphologic process
activity. This approach is mostly applied for rock fall processes using high-resolution LIDAR or SFM data (Rabatel et al., 2008;
Rosser et al., 2005; Warrick et al., 2017; Bremer and Sass, 2012). Even though spatial resolution of the data used in the presented
cases is high, the temporal resolution depends on the frequency of measurements. The resulting erosion rates thus only represent
current developments. A look into past erosion and more long-term rates can be established using DEM from historical aerial
imagery (Fischer et al., 2011). Micheletti et al. (2015) use seven different scenes of aerial images between 1967 and 2005 to quantify
surface changes in a high mountain environment and relate their observations to decadal climate changes in their study site. Within
this environment, glacial and periglacial landforms show the greatest changes in the time period. They relate phases of warming
with increasing surface displacement and downwasting and identify most dynamic changes in periods of observed increased
precipitation and high temperatures. Bennett et al. (2013) quantified hillslope and channel erosion using a similar approach in the
Illgraben catchment (Switzerland). Here, also periods of temperature increase and pronounced frequency and magnitude of
intense rainfall events are associated with the observed changes. It is important to acknowledge that the resolution of the aerial
images used, as well as the accuracy of the ground control points and the software used for DEM generation, significantly determine
the observable surface changes. The studies presented generated DEMs at average resolution between 0.3–0.5 m (Micheletti et al.,
2015) and 2–4 m (Bennett et al., 2012).
An alternative way to quantify erosion is the approach of geophysical relief (Champagnac et al., 2007; Small and Anderson,
1998). Geophysical relief describes the eroded volume of valleys or entire mountain ranges derived from the difference between the
actual surface heights and an interpolated surface connecting the highest points. The approach is used to analyze long-term
landform evolution on mountain ranges and to differentiate between impacts of tectonic uplift, erosion and isostatic rebound on
relief. We have applied the geophysical relief index on the study area (Fig. 19). The pattern of geophysical relief distribution in the
Obersulzbach valley reveals the greatest removal in the central part above the lower step. Relief values exceed 1000 m compared to
700–900 at the lower valley level. This could hint at increased glacial erosion in this location that may be due to a longer time of
glacier presence, for example, in late glacial times.

Conclusions

Modern quantitative geomorphological research can be regarded inextricably linked with GIS analysis. The availability of both
high-resolution and global data on the land surface contributed significantly to recent fundamental advances in the discipline and
opened new fields of research. Applications of GIS in geomorphology span from pure visualization approaches, landform
classification, land surface and hydrological analysis (usually derived from DEMs), process and erosion modeling, and topographic
change detection to hazard zonation or susceptibility modeling. Herein, statistical analysis and spatial interpolation of field data as
well as graphical visualization and map creation represent key features of GIS applied in geomorphology. Numerous topographic
and geomorphological indices have been developed to study geomorphological form and process configurations using GIS
techniques.
26 GIS Applications in Geomorphology

12°12′0″E 12°16′0″E 12°20′0″E

Geophysical relief (m)

0–100
100–200
200–300
300–400
400–500
500–600
600–700
700–800
47°12′0″N

47°12′0″N
800–900
900–1000
1000–1100
1100–1200
1200–1300

Lakes

Glaciers
47°8′0″N

47°8′0″N

0 0.5 1 2
km

12°16′0″E 12°20′0″E

Fig. 19 Geophysical relief calculated after Small and Anderson (1998). Note the differences between the central valley and the northern valley section
(see text for explanation).

In our paper we presented a selection of GIS-based tools and indices in some classic fields of geomorphology such as fluvial,
gravitational, glacial, and periglacial environments. A promising approach which has become more popular in recent years focuses
on quantifying sediment fluxes and deposits using digital data. The latter can be achieved by combining surface and subsurface data
on sediment depth or by comparing surface data from various points in time.
GIS Applications in Geomorphology 27

Increasing resolution of both DEM and image data, free availability of local and global data sets, and low-cost technology to
generate high-resolution surface information will foster the possibilities of geomorphological analysis using GIS. Challenges,
however, exist with respect to scale and the applicability of tools and parameters originally developed using data of lower
resolution. High level of detail can contribute to scientific insights but may also represent noise that prevents clarity (Drăguţ and
Blaschke, 2006). Scale-dependency of LSPs and objects needs to be carefully considered when performing quantitative landform
analysis.

References
Abermann J, Lambrecht A, Fischer A, and Kuhn M (2009) Quantifying changes and trends in glacier area and volume in the Austrian Ötztal Alps (1969–1997–2006). The Cryosphere
3: 205.
Ahnert F (1970) Functional relationships between denudation, relief, and uplift in large, mid-latitude drainage basins. American Journal of Science 268: 243–263.
Allen TR (1998) Topographic context of glaciers and perennial snowfields, Glacier National Park, Montana. Geomorphology 21: 207–216.
Anders NS, Seijmonsbergen AC, and Bouten W (2011) Segmentation optimization and stratified object-based analysis for semi-automated geomorphological mapping. Remote
Sensing of Environment 115: 2976–2985.
Bagnold RA (1960) Sediment discharge and stream power—a preliminary announcement. Circular Reston, Virginia: US Geological Survey 421.
Baltsavias EP, Favey E, Bauder A, Bosch H, and Pateraki M (2001) Digital surface modelling by airborne laser scanning and digital photogrammetry for glacier monitoring. The
Photogrammetric Record 17: 243–273.
Benn DI and Lehmkuhl F (2000) Mass balance and equilibrium-line altitudes of glaciers in high-mountain environments. Quaternary International 65–66: 15–29.
Bennett GL, Molnar P, Eisenbeiss H, and Mcardell BW (2012) Erosional power in the Swiss Alps: characterization of slope failure in the Illgraben. Earth Surface Processes and
Landforms 37: 1627–1640.
Bennett GL, Molnar P, Mcardell BW, Schlunegger F, and Burlando P (2013) Patterns and controls of sediment production, transfer and yield in the Illgraben. Geomorphology
188: 68–82.
Beven KJ and Kirkby MJ (1979) A physically based, variable contributing area model of basin hydrology/Un modèle à base physique de zone d’appel variable de l’hydrologie du bassin
versant. Hydrological Sciences Bulletin 24: 43–69.
Bishop MP (2013) Remote sensing and GIScience in geomorphology: introduction and overview A2. In: Shroder JF (ed.) Treatise on geomorphology. San Diego: Academic Press.
Blaschke T (2010) Object based image analysis for remote sensing. ISPRS Journal of Photogrammetry and Remote Sensing 65: 2–16.
Blaschke T and Strobl J (2001) What’s wrong with pixels? Some recent developments interfacing remote sensing and GIS. Image Rochester NY 6: 12–17.
Böhner J and Antonić O (2009) Land-surface parameters specific to topo-climatology. In: Tomislav H and Hannes IR (eds.) Developments in soil science. Amsterdam: Elsevier.
Chapter 8.
Borselli L, Cassi P, and Torri D (2008) Prolegomena to sediment and flow connectivity in the landscape: a GIS and field numerical assessment. Catena 75: 268–277.
Bremer M and Sass O (2012) Combining airborne and terrestrial laser scanning for quantifying erosion and deposition by a debris flow event. Geomorphology 138: 49–60.
Brocklehurst SH and Whipple KX (2004) Hypsometry of glaciated landscapes. Earth Surface Processes and Landforms 29: 907–926.
Burbank DW and Anderson RS (2011) Tectonic geomorphology. Hoboken, NJ: John Wiley & Sons, Ltd.
Campbell D and Church M (2003) Reconnaissance sediment budgets for Lynn Valley, British Columbia: Holocene and contemporary time scales. Canadian Journal of Earth Sciences
40: 701–713.
Carrara A (1983) Multivariate models for landslide hazard evaluation. Journal of the International Association for Mathematical Geology 15: 403–426.
Carrara A and Guzzetti F (1995) Geographical information systems in assessing natural hazards. Dordrecht: Springer.
Cavalli M, Trevisani S, Comiti F, and Marchi L (2013) Geomorphometric assessment of spatial sediment connectivity in small Alpine catchments. Geomorphology 188: 31–41.
Chairat S and Delleur JW (1993) Effects of the topographic index distribution on predicted runoff using grass. Water Resources Bulletin 29: 1029–1034.
Champagnac JD, Molnar P, Anderson RS, Sue C, and Delacou B (2007) Quaternary erosion-induced isostatic rebound in the western Alps. Geology 35: 195–198.
Chen A, Darbon J, and Morel J-M (2014) Landscape evolution models: a review of their fundamental equations. Geomorphology 219: 68–86.
Costa-Cabral MC and Burges SJ (1994) Digital Elevation Model Networks (DEMON): A model of flow over hillslopes for computation of contributing and dispersal areas. Water
Resources Research 30: 1681–1692.
Coulthard TJ (2001) Landscape evolution models: a software review. Hydrological Processes 15: 165–173.
Curry AM (1999) Paraglacial modification of slope form. Earth Surface Processes and Landforms 24: 1213–1228.
Dadson SJ and Church M (2005) Postglacial topographic evolution of glaciated valleys: a stochastic landscape evolution model. Earth Surface Processes and Landforms
30: 1387–1403.
Deroo APJ, Hazelhoff L, and Burrough PA (1989) Soil-erosion modeling using answers and geographical information-systems. Earth Surface Processes and Landforms 14: 517–532.
Dikau R (1989) The application of a digital relief model to landform analysis in geomorphology. In: Raper JF (ed.) Three dimensional applications in geographical information systems,
London: Taylor & Francis.
Dikau R (1996) Geomorphologische Reliefklassifikation und -analyse. Heidelberger Geographische Arbeiten 104: 15–36.
Dikau R and Jäger S (1995) Landslide hazard modelling in New Mexico and Germany. In: Mcgregor D and Thompson D (eds.) Geomorphology and land management in a changing
environment, Chichester: John Wiley.
Dikau, R., Brabb, E. E. and Mark, R. M. (1991). Landform classification of New Mexico by computer. Open-File Report.—ed.
D’Oleire-Oltmanns S, Eisank C, Dragut L, and Blaschke T (2013) An object-based workflow to extract landforms at multiple scales from two distinct data types. Geoscience and Remote
Sensing Letters, IEEE 10: 947–951.
Drăguţ L and Blaschke T (2006) Automated classification of landform elements using object-based image analysis. Geomorphology 81: 330–344.
Eash DA (1994) A geographic information-system procedure to quantify drainage-basin characteristics. Water Resources Bulletin 30: 1–8.
Egholm DL, Pedersen VK, Knudsen MF, and Larsen NK (2012) Coupling the flow of ice, water, and sediment in a glacial landscape evolution model. Geomorphology 141–142: 47–66.
Eisank C, Smith M, and Hillier J (2014) Assessment of multiresolution segmentation for delimiting drumlins in digital elevation models. Geomorphology 214: 452–464.
Erskine RH, Green TR, Ramirez JA, and Macdonald LH (2006) Comparison of grid-based algorithms for computing upslope contributing area. Water Resources Research 42: 1–9.
Evans IS (1972) General geomorphometry, derivatives of altitude, and descriptive statistics. In: Chorley RJ (ed.) Spatial analysis in geomorphology. London: Methuen.
Evans IS (2012) Geomorphometry and landform mapping: what is a landform? Geomorphology 137: 94–106.
Fairfield J and Leymarie P (1991) Drainage networks from grid digital elevation models. Water Resources Research 27: 709–717.
Farr TG, Rosen PA, Caro E, Crippen R, Duren R, Hensley S, Kobrick M, Paller M, Rodriguez E, Roth L, Seal D, Shaffer S, Shimada J, Umland J, Werner M, Oskin M, Burbank D, and
Alsdorf D (2007) The shuttle radar topography mission. Reviews of Geophysics 45:RG2004.
Fischer L, Eisenbeiss H, Kääb A, Huggel C, and Haeberli W (2011) Monitoring topographic changes in a periglacial high-mountain face using high-resolution DTMs, Monte Rosa East
Face, Italian Alps. Permafrost and Periglacial Processes 22: 140–152.
28 GIS Applications in Geomorphology

Fischer A, Seiser B, Stocker Waldhuber M, Mitterer C, and Abermann J (2015) Tracing glacier changes in Austria from the Little Ice Age to the present using a lidar-based high-
resolution glacier inventory in Austria. The Cryosphere 9: 753–766.
Flint JJ (1974) Stream gradient as a function of order, magnitude, and discharge. Water Resources Research 10: 969–973.
Freeman TG (1991) Calculating catchment area with divergent flow based on a regular grid. Computers & Geosciences 17: 413–422.
Frey H, Machguth H, Huss M, Huggel C, Bajracharya S, Bolch T, Kulkarni A, Linsbauer A, Salzmann N, and Stoffel M (2013) Ice volume estimates for the Himalaya–Karakoram region:
evaluating different methods. The Cryosphere Discuss 7: 4813–4854.
Gay A, Cerdan O, Mardhel V, and Desmet M (2016) Application of an index of sediment connectivity in a lowland area. Journal of Soils and Sediments 16: 280–293.
Gesch D, Oimoen M, Zhang Z, Meyer D, and Danielson J (2012) Validation of the Aster Global Digital Elevation Model Version 2 over the conterminous United States. In: Proceeding
of the XXII ISPRS Congress, pp. 281–286, Melbourne: International Society for Photogrammetry and Remote Sensing.
Gómez H and Kavzoglu T (2005) Assessment of shallow landslide susceptibility using artificial neural networks in Jabonosa River Basin, Venezuela. Engineering Geology 78: 11–27.
Götz J, Otto JC, and Schrott L (2013) Postglacial sediment storage and rockwall retreat in a semi-closed inner-alpine sedimentary basin (Gradenmoos, Hohe Tauern, Austria).
Geografia Fisica e Dinamica Quaternaria 36: 63–80.
Goudie AS (ed.) (1990) Geomorphological techniques. London: Unwin Hyman.
Gregory KJ and Lewin J (2014) The basics of geomorphology: key concepts. London: Sage.
Grieve SWD, Mudd SM, Hurst MD, and Milodowski DT (2016) A nondimensional framework for exploring the relief structure of landscapes. Earth Surface Dynamics 4: 309–325.
Grohmann CH, Smith MJ, and Riccomini C (2011) Multiscale analysis of topographic surface roughness in the Midland Valley, Scotland. Geoscience and Remote Sensing, IEEE
Transactions on 49: 1200–1213.
Gross G, Kerschner H, and Patzelt G (1977) Methodische Untersuchungen über die Schneegrenze in alpinen Gletschergebieten. Zeitschrift für Gletscherkunde und Glazialgeologie
12: 223–251.
Gruber U and Bartelt P (2007) Snow avalanche hazard modelling of large areas using shallow water numerical methods and GIS. Environmental Modelling & Software 22: 1472–1481.
Gruber FE and Mergili M (2013) Regional-scale analysis of high-mountain multi-hazard and risk indicators in the Pamir (Tajikistan) with GRASS GIS. Natural Hazards and Earth
System Sciences 13: 2779–2796.
Gruber S and Peckham S (2009) Land-surface parameters and objects in hydrology. In: Hengl T and Reuter HI (eds.) Amsterdam: Elsevier. Chapter 7.
Gustavsson M, Kolstrup E, and Seijmonsbergen AC (2006) A new symbol-and-GIS based detailed geomorphological mapping system: renewal of a scientific discipline for
understanding landscape development. Geomorphology 77: 90–111.
Gustavsson M, Seijmonsbergen AC, and Kolstrup E (2008) Structure and contents of a new geomorphological GIS database linked to a geomorphological map—with an example from
Liden, central Sweden. Geomorphology 95: 335–349.
Guzzetti F, Carrara A, Cardinali M, and Reichenbach P (1999) Landslide hazard evaluation: a review of current techniques and their application in a multi-scale study, Central Italy.
Geomorphology 31: 181–216.
Hack JT (1957) Studies of longitudinal stream profiles in Virginia and Maryland. Shorter Contributions to General Geology: 45–97.
Hack JT (1973) Stream-profile analysis and stream-gradient index. Journal of Research of the U.S. Geological Survey 1: 421–429.
Haeberli W and Hölzle M (1995) Application of inventory data for estimating characteristics of and regional climate effects on mountain glaciers: a pilot study with the European Alps.
Annals of Glaciology 21: 206–212.
Hara K, Zhao Y, Harada I, Tomita M, Park J, Jung E, Kamagata N, and Hirabuki Y (2015) Multi-scale monitoring of landscape change after the 2011 tsunami. International Archives of
the Photogrammetry, Remote Sensing and Spatial Information Sciences—ISPRS Archives XL-7/W3: 805–809.
Harbor J and Wheeler DA (1992) On the mathematical description of glaciated valley cross section. Earth Surface Processes and Landforms 17: 477–485.
Harvey AM (2001) Coupling between hillslopes and channels in upland fluvial systems: implications for landscape sensitivity, illustrated from the Howgill Fells, northwest England.
Catena 42: 225–250.
Heckmann T and Schwanghart W (2013) Geomorphic coupling and sediment connectivity in an alpine catchment—exploring sediment cascades using graph theory. Geomorphology
182: 89–103.
Hengl T and Reuter HI (eds.) (2009) Geomorphometry—concepts, software, applications, Oxford: Elsevier.
Hickey R, Smith A, and Jankowski P (1994) Slope length calculations from a DEM within ARC/INFO grid. Computers, Environment and Urban Systems 18: 365–380.
Hinderer M (2001) Late Quaternary denudation of the Alps, valley and lake fillings and modern river loads. Geodinamica Acta 14: 231–263.
Funk M and Hoelzle M (1992) A model of potential direct solar radiation for investigating occurrences of mountain permafrost. Permafrost and Periglacial Processes 3(2): 139–142.
Hoffmann T and Schrott L (2002) Modelling sediment thickness and rockwall retreat in an Alpine valley using 2D-seismic refraction (Reintal, Bavarian Alps). Zeitschrift für
Geomorphologie, Supplement Band 127: 153–173.
Horton RE (1932) Drainage–basin characteristics. EOS, Transactions American Geophysical Union 13: 350–361.
Horton RE (1945) Erosional development of streams and their drainage basins; hydrophysical approach to quantitative morphology. Geological Society of America Bulletin
56: 275–370.
Huabin W, Gangjun L, Weiya X, and Gonghui W (2005) GIS-based landslide hazard assessment: an overview. Progress in Physical Geography 29: 548–567.
Humlum O (1998) The climatic significance of rock glaciers. Permafrost and Periglacial Processes 9: 375–395.
Huss M and Farinotti D (2012) Distributed ice thickness and volume of all glaciers around the globe. Journal of Geophysical Research: Earth Surface 117: F04010.
Jaboyedoff M and Derron MH (2005a) A new method to estimate the infilling of alluvial sediment of glacial valleys using a sloping local base level. Geografia Fisica e Dinamica
Quaternaria 28: 37–46.
Jaboyedoff M and Derron MH (2005b) A new method to estimate the infilling of alluvial sediment of glacial valleys using a sloping local base level. Geografia Fisica e Dinamica
Quaternaria 28: 37–46.
Jäger S (1997) Fallstudien zur Bewertung von Massenbewegungen als geomorphologische Naturgefahr. Heidelberg: Selbstverlag des Geographischen Instituts der Universität.
James LA (1996) Polynomial and power functions for glacial valley cross-section morphology. Earth Surface Processes and Landforms 21: 413–432.
Jenson SK and Domingue JO (1988) Extracting topographic structure from digital elevation data for geographical information system analysis. Photogrammetric Engineering and
Remote Sensing 54: 1593–1600.
Kääb A, Winsvold S, Altena B, Nuth C, Nagler T, and Wuite J (2016) Glacier remote sensing using sentinel-2. Part I: radiometric and geometric performance, and application to ice
velocity. Remote Sensing 8: 598.
Keller F (1992) Automated mapping of mountain permafrost using the program PERMAKART within the geographical information system ARC/INFO. Permafrost and Periglacial
Processes 3: 133–138.
Keutterling A and Thomas A (2006) Monitoring glacier elevation and volume changes with digital photogrammetry and GIS at Gepatschferner glacier, Austria. International Journal of
Remote Sensing 27: 4371–4380.
Kirkby MJ (1987) Modelling some influences of soil erosion, landslides and valley gradient on drainage density and hollow development. Catena Supplement 10: 1–14.
Koethe R and Lehmeier F (1993) SARA—Ein Programmsystem zur Automatischen Relief-Analyse. Zeitschrift für Angewandte Geographie 4: 11–21.
Kruse S (2013) 3.5 Near-surface geophysics in geomorphology A2. In: Shroder JF (ed.) Treatise on geomorphology, San Diego: Academic Press.
Kugler, H. (1975). Das Georelief und seine kartographische Modellierung. Dissertation, Martin-Luther-Universität Halle.
Lan H, Derek Martin C, and Lim CH (2007) RockFall analyst: a GIS extension for three-dimensional and spatially distributed rockfall hazard modeling. Computers & Geosciences
33: 262–279.
Lane SN, James TD, and Crowell MD (2000) Application of digital photogrammetry to complex topography for geomorphological research. The Photogrammetric Record
16: 793–821.
GIS Applications in Geomorphology 29

Lane SN, Bakker M, Gabbud C, Micheletti N, and Saugy JN (2017) Sediment export, transient landscape response and catchment-scale connectivity following rapid climate warming
and Alpine glacier recession. Geomorphology 277: 210–227.
Lautensach H (1959) Carl Troll—Ein Forscherleben. Erdkunde 13: 245–258.
Legleiter CJ and Fonstad MA (2012) An introduction to the physical basis for deriving river information by optical remote sensing. Fluvial remote sensing for science and
management. Hoboken, NJ: John Wiley & Sons, Ltd.
Lehner B, Verdin K, and Jarvis A (2008) New global hydrography derived from spaceborne elevation data. Eos, Transactions American Geophysical Union 89: 93–94.
Leopold LB, Wolman MG, and Miller JP (1964) Fluvial processes in geomorphology. San Francisco: W.H. Freeman and Co.
Li X and Damen MCJ (2010) Coastline change detection with satellite remote sensing for environmental management of the Pearl River Estuary, China. Journal of Marine Systems 82
(Suppl.): S54–S61.
Liu BY, Nearing MA, Shi PJ, and Jia ZW (2000) Slope length effects on soil loss for steep slopes. Soil Science Society of America Journal 64: 1759–1763.
Macmillan RA, Pettapiece WW, Nolan SC, and Goddard TW (2000) A generic procedure for automatically segmenting landforms into landform elements using DEMs, heuristic rules
and fuzzy logic. Fuzzy Sets and Systems 113: 81–109.
Marthews TR, Dadson SJ, Lehner B, Abele S, and Gedney N (2015) High-resolution global topographic index values for use in large-scale hydrological modelling. Hydrology and
Earth System Sciences 19: 91–104.
Messenzehl K, Hoffmann T, and Dikau R (2014) Sediment connectivity in the high-alpine valley of Val Müschauns, Swiss National Park—linking geomorphic field mapping with
geomorphometric modelling. Geomorphology 221: 215–229.
Mey J, Scherler D, Wickert AD, Egholm DL, Tesauro M, Schildgen TF, and Strecker MR (2016) Glacial isostatic uplift of the European Alps. Nature Communications 7: 13382.
Micheletti N, Lambiel C, and Lane SN (2015) Investigating decadal-scale geomorphic dynamics in an alpine mountain setting. Journal of Geophysical Research: Earth Surface
120: 2155–2175.
Minár J and Evans IS (2008) Elementary forms for land surface segmentation: the theoretical basis of terrain analysis and geomorphological mapping. Geomorphology 95: 236–259.
Minar J, Mentlik P, Jedlicka K, and Barka I (2005) Geomorphological information system: idea and options for practical implementation. Geograficky Casopis 57: 247–264.
Montgomery DR and Foufoula-Georgiou E (1993) Channel network source representation using digital elevation models. Water Recources Research 29: 3925–3934.
Moore ID and Burch GJ (1986) Physical basis of the length-slope factor in the universal soil loss Equation1. Soil Science Society of America Journal 50: 1294–1298.
Moore ID, Grayson RB, and Ladson AR (1991) Digital terrain modelling: a review of hydrological, geomorphological, and biological applications. Hydrological Processes 5: 3–30.
Napieralski J, Harbor J, and Li Y (2007) Glacial geomorphology and geographic information systems. Earth-Science Reviews 85: 1–22.
Oguchi T and Wasklewicz TA (2011) Geographic information systems in geomorphology. In: Gregory KJ and Goudie AS (eds.) The SAGE handbook of geomorphology, London:
SAGE.
Olaya V (2009) Basic land-surface parameters. In: Tomislav H and Hannes IR (eds.) Developments in soil science. Amsterdam: Elsevier. Chapter 6.
Olyphant GA (1977) Topoclimate and the Depth of Cirque Erosion. Geografiska Annaler. Series A. Physical Geography 59: 209–213.
Otto J-C and Dikau R (2004) Geomorphologic system analysis of a high mountain valley in the Swiss Alps. Zeitschrift für Geomorphologie, NF 48: 323–341.
Otto JC and Smith M (2013) Section 2.6 Geomorphological mapping. In: Clarke LE (ed.) Geomorphological techniques (online edition). London: British Society for Geomorphology.
Otto J-C, Schrott L, Jaboyedoff M, and Dikau R (2009) Quantifying sediment storage in a high alpine valley (Turtmanntal, Switzerland). Earth Surface Processes and Landforms
34: 1726–1742.
Otto JC, Gustavsson M, and Geilhausen M (2011) Cartography: design, symbolisation and visualisation of geomorphological maps. In: Smith MJ, Paron P, and Griffiths J (eds.)
Geomorphological mapping: methods and applications, London: Elsevier.
Otto J-C, Keuschnig M, Götz J, Marbach M, and Schrott L (2012) Detection of mountain permafrost by combining high resolution surface and subsurface information—an example
from the Glatzbach catchment, Austrian Alps. Geografiska Annaler: Series A, Physical Geography 94: 43–57.
Paul F, Bolch T, Kääb A, Nagler T, Nuth C, Scharrer K, Shepherd A, Strozzi T, Ticconi F, Bhambri R, Berthier E, Bevan S, Gourmelen N, Heid T, Jeong S, Kunz M, Lauknes TR,
Luckman A, Merryman Boncori JP, Moholdt G, Muir A, Neelmeijer J, Rankl M, Vanlooy J, and Van Niel T (2015) The glaciers climate change initiative: methods for creating glacier
area, elevation change and velocity products. Remote Sensing of Environment 162: 408–426.
Penck A (1894) Morphologie der Erdoberfläche, 2nd edn Stuttgart: Engelhorn.
Pike R and Dikau R (1995) Advances in Geomorphometry. Proceedings of the Walter F. Wood Memorial Symposium. Zeitschrift für Geomorphologie/Supplement, vol. 101, 238 p.
Pike RJ and Wilson SE (1971) Elevation-relief ratio, hypsometric integral, and geomorphic area-altitude analysis. Geological Society of America Bulletin 82: 1079–1084.
Porter SC (1975) Equilibrium-line altitudes of late Quaternary glaciers in the Southern Alps, New Zealand. Quaternary Research 5: 27–47.
Prasicek G, Otto J-C, Montgomery DR, and Schrott L (2014) Multi-scale curvature for automated identification of glaciated mountain landscapes. Geomorphology 209: 53–65.
Prasicek G, Larsen IJ, and Montgomery DR (2015) Tectonic control on the persistence of glacially sculpted topography. Nature Communications 6:8028.
Rabatel A, Deline P, Jaillet S, and Ravanel L (2008) Rock falls in high-alpine rock walls quantified by terrestrial lidar measurements: a case study in the Mont Blanc area. Geophysical
Research Letters 35(10): L10502.
Racoviteanu A, Williams M, and Barry R (2008) Optical remote sensing of glacier characteristics: a review with focus on the Himalaya. Sensors 8: 3355.
Riseborough D, Shiklomanov N, Etzelmüller B, Gruber S, and Marchenko S (2008) Recent advances in permafrost modelling. Permafrost and Periglacial Processes 19: 137–156.
Rosser NJ, Petley DN, Lim M, Dunning SA, and Allison RJ (2005) Terrestrial laser scanning for monitoring the process of hard rock coastal cliff erosion. Quarterly Journal of
Engineering Geology and Hydrogeology 38: 363–375.
Rowland JC, Shelef E, Pope PA, Muss J, Gangodagamage C, Brumby SP, and Wilson CJ (2016) A morphology independent methodology for quantifying planview river change and
characteristics from remotely sensed imagery. Remote Sensing of Environment 184: 212–228.
Salcher BC, Kober F, Kissling E, and Willett SD (2014) Glacial impact on short-wavelength topography and long-lasting effects on the denudation of a deglaciated mountain range.
Global and Planetary Change 115: 59–70.
Sass O (2007) Bedrock detection and talus thickness assessment in the European Alps using geophysical methods. Journal of Applied Geophysics 3: 254–269.
Sattler K, Anderson B, Mackintosh A, Norton K, and DE Róiste M (2016) Estimating permafrost distribution in the maritime Southern Alps, New Zealand, based on climatic conditions
at rock glacier sites. Frontiers in Earth Science 4.
Scaioni M, Longoni L, Melillo V, and Papini M (2014) Remote sensing for landslide investigations: an overview of recent achievements and perspectives. Remote Sensing 6: 9600.
Schmidt J and Hewitt A (2004) Fuzzy land element classification from DTMs based on geometry and terrain position. Geoderma 121: 243–256.
Schneevoigt NJ, Van der Linden S, Thamm H-P, and Schrott L (2008) Detecting Alpine landforms from remotely sensed imagery. A pilot study in the Bavarian Alps. Geomorphology
93: 104–119.
Schoeneich P (1993) Comparaison des systémes de légendes francais, allemand et suisse principes de la légende IGUL. Travaux et Recherches 9: 15–24.
Schrott L and Adams T (2002) Quantifying sediment storage and Holocene denudation in an Alpine basin, Dolomites, Italy. Zeitschrift für Geomorphologie N.F. 128(Suppl. Bd):
129–145.
Schrott L and Sass O (2008) Application of field geophysics in geomorphology: advances and limitations exemplified by case studies. Geomorphology 93: 55–73.
Schrott L, Hufschmidt G, Hankammer M, Hoffmann T, and Dikau R (2003a) Spatial distribution of sediment storage types and quantification of valley fill deposits in an alpine basin,
Reintal, Bavarian Alps, Germany. Geomorphology 55: 45–63.
Schrott L, Hufschmidt G, Hankammer M, Hofmann T, and Dikau R (2003b) Spatial distribution of sediment storage types and quantification of valley fill deposits in an alpine basin,
Reintal, Bavarian Alps, Germany. Geomorphology 55: 45–63.
Schrott L, Otto JC, and Keller F (2013) Modelling alpine permafrost distribution in the Hohe Tauern region, Austria. Austrian Journal of Earth Science 105: 169–183.
Seibert J and Mcglynn B (2007) A new triangular multiple flow direction algorithm for computing upslope areas from gridded digital elevation models. Water Resources Research
43: 1–8.
30 GIS Applications in Geomorphology

Shary PA, Sharaya LS, and Mitusov AV (2005) The problem of scale-specific and scale-free approaches in geomorphometry. Geografia Fisica e Dinamica Quaternaria 28: 81–101.
Shroder JF, Scheppy RA, and Bishop MP (1999) Denudation of small alpine basins, Nanga Parbat Himalaya, Pakistan. Arctic Antarctic and Alpine Research 31: 121–127.
Small EE and Anderson RS (1998) Pleistocene relief production in Laramide mountain ranges, western United States. Geology 26: 123–126.
Smith MJ (2011) Digital mapping: visualisation, interpretation and quantification of landforms. In: Smith MJ, Paron P, and Griffiths J (eds.) Geomorphological mapping: methods and
applications. London: Elsevier.
Smith MW (2014) Roughness in the Earth Sciences. Earth-Science Reviews 136: 202–225.
Smith MJ and Wise SM (2007) Problems of bias in mapping linear landforms from satellite imagery. International Journal of Applied Earth Observation and Geoinformation 9: 65–78.
Smith MJ, Rose J, and Booth S (2006) Geomorphological mapping of glacial landforms from remotely sensed data: an evaluation of the principal data sources and an assessment of
their quality. Geomorphology 76: 148–165.
Smith MJ, Hilier J, Otto JC, and Geilhausen M (2013) Geovisualisation. In: Shroder JF (ed.) Treatise on geomorphology, San Diego: Academic Press Elsevier.
Snavely N, Seitz SM, and Szeliski R (2008) Modeling the world from Internet photo collections. International Journal of Computer Vision 80: 189–210.
Sternai P, Herman F, Fox MR, and Castelltort S (2011) Hypsometric analysis to identify spatially variable glacial erosion. Journal of Geophysical Research 116: F03001.
Strahler AN (1952) Hypsometric (area-altitude) analysis of erosional topography. Geological Society of America Bulletin 63: 1117–1142.
Strahler AN (1957) Quantitative analysis of watershed geomorphology. Transactions of the American Geophysical Union 38: 913–920.
Stumpf A, Malet JP, and Delacourt C (2017) Correlation of satellite image time-series for the detection and monitoring of slow-moving landslides. Remote Sensing of Environment
189: 40–55.
Svensson H (1959) Is the cross-section of a glacial valley a parabola? Journal of Glaciology 3: 362–363.
Tarboton DG, Bras RL, and Rodriguez-Iturbe I (1991) On the extraction of channel networks from digital elevation data. Hydrological Processes 5: 81–100.
Theler D, Reynard E, and Bardou E (2008) Assessing sediment dynamics from geomorphological maps: Bruchi torrential system, Swiss Alps. Journal of Maps 4: 277–289.
Tucker GE and Bras RL (1998) Hillslope processes, drainage density, and landscape morphology. Water Resources Research 34: 2751–2764.
Tucker GE and Hancock GR (2010) Modelling landscape evolution. Earth Surface Processes and Landforms 35: 28–50.
Tucker GE and Whipple KX (2002) Topographic outcomes predicted by stream erosion models: sensitivity analysis and intermodel comparison. Journal of Geophysical Research:
Solid Earth 107. ETG 1-1–ETG 1-16.
Tunnicliffe J, Church M, Clague JJ, and Feathers JK (2012) Postglacial sediment budget of Chilliwack Valley, British Columbia. Earth Surface Processes and Landforms
37: 1243–1262.
Van Westen CJ, Castellanos E, and Kuriakose SL (2008) Spatial data for landslide susceptibility, hazard, and vulnerability assessment: an overview. Engineering Geology
102: 112–131.
Vanwesten CJ and Terlien MTJ (1996) An approach towards deterministic landslide hazard analysis in GIS. A case study from Manizales (Colombia). Earth Surface Processes and
Landforms 21: 853–868.
Warrick JA, Ritchie AC, Adelman G, Adelman K, and Limber PW (2017) New techniques to measure Cliff change from historical oblique aerial photographs and structure-from-motion
photogrammetry. Journal of Coastal Research 33: 39–55.
Westoby MJ, Brasington J, Glasser NF, Hambrey MJ, and Reynolds JM (2012) ‘Structure-from-Motion’ photogrammetry: a low-cost, effective tool for geoscience applications.
Geomorphology 179: 300–314.
Whipple KX (2004) Bedrock rivers and the geomorphology of active orogens. Annual Review of Earth and Planetary Sciences 32: 151–185.
Whipple KX and Tucker GE (1999) Dynamics of the stream-power river incision model: implications for height limits of mountain ranges, landscape response timescales, and research
needs. Journal of Geophysical Research: Solid Earth 104: 17661–17674.
Whipple KX, Dibiase RA, and Crosby BT (2013) Bedrock rivers. In: Shroder JF and Wohl E (eds.) Treatise on geomorphology. San Diego: Academic Press.
Wichmann V and Becht M (2006) Rockfall modelling: methods and model application in an Alpine basin. Göttingen: Goltze.
Wilford DJ, Sakals ME, Innes JL, Sidle RC, and Bergerud WA (2004) Recognition of debris flow, debris flood and flood hazard through watershed morphometrics. Landslides
1: 61–66.
Wilson JP and Bishop MP (2013) 3.7 Geomorphometry A2. In: Shroder JF (ed.) Treatise on geomorphology, San Diego: Academic Press.
Wilson JP and Gallant JC (eds.) (2000) Terrain analysis: principles and applications, New York: Wiley.
Wulder MA and Coops NC (2014) Make Earth observations open access. Nature 513: 30–31.
Zink M, Fiedler H, Hajnsek I, Krieger G, Moreira A, and Werner M (2006) The TanDEM-X mission concept. IEEE International Symposium on Geoscience and Remote Sensing
1938–1941.
Zink M, Bartusch M, and Miller D (2011) TanDEM-X mission status. IEEE International Geoscience and Remote Sensing Symposium 2011: 2290–2293.

Relevant Websites

www.geomorphometry.org/.
http://gis4geomorphology.com/.
https://www.qgis.org/en/site/.
http://www.saga-gis.org/en/index.html.
http://serc.carleton.edu/NAGTWorkshops/geomorph/vignettes.html.

View publication stats

Anda mungkin juga menyukai