Techniques for quantifying the accuracy of gridded elevation models and for mapping uncertainty in...

26
Article Techniques for quantifying the accuracy of gridded elevation models and for mapping uncertainty in digital terrain analysis N. Gonga-Saholiariliva Universite ´ Paris Diderot, France; TTI Production, France Y. Gunnell Universite ´ Lumie `re Lyon 2, France C. Petit Universite ´ de Nice, France C. Mering Universite ´ Paris Diderot, France Abstract We first provide a critical review of statistical procedures employed in the literature for testing uncertainty in digital terrain analysis, then focus on several aspects of spatial autocorrelation that have been neglected in the analysis of gridded elevation data. When applied to first derivatives of elevation such as topographic slope, a spatial approach using Moran’s I and the LISA (Local Indicator of Spatial Association) allows: (1) georeferenced data patterns to be generated; (2) error hot- and coldspots to be located; and (3) error propagation during DEM manipulation to be evaluated. In a worked example focusing on the Wasatch mountain front, Utah, we analyse the relative advantages of six DEMs resulting from different acquisition modes (airborne, optical, radar, or composite): the LiDAR (2 m), CODEM (5 m), NED10 (10 m), ASTER DEM (15 m) and GDEM (30 m), and SRTM (90 m). The example shows that (apart from the LiDAR) the NED10, which is generated from composite data sources, is the least error-ridden DEM for that region. Knowing error magnitudes and where errors are located determines where corrections to ele- vation are required in order to minimize error accumulation or propagation, and clarifies how they might affect expert judgement in environmental decisions. Ground resolution issues can subsequently be addressed with greater confidence by resampling the preferred grid to terrain resolutions suited to the landscape attributes of interest. Source product testing is an essential yet often neglected part of DEM analysis, with many practical applications in hydrological modelling, for predictions of slope- to catchment-scale mass sediment flux, or for the assessment of slope stability thresholds. Corresponding author: N. Gonga-Saholiariliva, 33 impasse des Arbousiers, F-30250, Sommie `res, France Email: [email protected] Progress in Physical Geography 1–26 ª The Author(s) 2011 Reprints and permission: sagepub.co.uk/journalsPermissions.nav DOI: 10.1177/0309133311409086 ppg.sagepub.com

Transcript of Techniques for quantifying the accuracy of gridded elevation models and for mapping uncertainty in...

Article

Techniques for quantifying theaccuracy of gridded elevationmodels and for mappinguncertainty in digital terrainanalysis

N. Gonga-SaholiarilivaUniversite Paris Diderot, France; TTI Production, France

Y. GunnellUniversite Lumiere Lyon 2, France

C. PetitUniversite de Nice, France

C. MeringUniversite Paris Diderot, France

AbstractWe first provide a critical review of statistical procedures employed in the literature for testing uncertainty indigital terrain analysis, then focus on several aspects of spatial autocorrelation that have been neglected in theanalysis of gridded elevation data. When applied to first derivatives of elevation such as topographic slope,a spatial approach using Moran’s I and the LISA (Local Indicator of Spatial Association) allows:(1) georeferenced data patterns to be generated; (2) error hot- and coldspots to be located; and(3) error propagation during DEM manipulation to be evaluated. In a worked example focusing on theWasatch mountain front, Utah, we analyse the relative advantages of six DEMs resulting from differentacquisition modes (airborne, optical, radar, or composite): the LiDAR (2 m), CODEM (5 m), NED10(10 m), ASTER DEM (15 m) and GDEM (30 m), and SRTM (90 m). The example shows that (apart fromthe LiDAR) the NED10, which is generated from composite data sources, is the least error-ridden DEM forthat region. Knowing error magnitudes and where errors are located determines where corrections to ele-vation are required in order to minimize error accumulation or propagation, and clarifies how they mightaffect expert judgement in environmental decisions. Ground resolution issues can subsequently be addressedwith greater confidence by resampling the preferred grid to terrain resolutions suited to the landscapeattributes of interest. Source product testing is an essential yet often neglected part of DEM analysis,with many practical applications in hydrological modelling, for predictions of slope- to catchment-scale masssediment flux, or for the assessment of slope stability thresholds.

Corresponding author:N. Gonga-Saholiariliva, 33 impasse des Arbousiers, F-30250, Sommieres, FranceEmail: [email protected]

Progress in Physical Geography1–26

ª The Author(s) 2011Reprints and permission:

sagepub.co.uk/journalsPermissions.navDOI: 10.1177/0309133311409086

ppg.sagepub.com

Keywordsdigital elevation model, LISA, Moran’s I, spatial autocorrelation, topographic slope, uncertainty mapping

I Introduction

The advent of remote sensing technologies has

meant that larger areas can be mapped by fewer

people in less time and at diminishing cost

(Smith and Clark, 2005). The widespread use

of digital elevation models (DEMs) is a direct

consequence of those advantages. Traditional

methods such as field surveying and photogram-

metry can yield high accuracy terrain data, but

they are time consuming and labour intensive

(Liu, 2008). As a result, DEMs have become the

most common method for extracting topo-

graphic information, also allowing the flow of

water across topography to be analysed, with

many practical applications in hydrological

modelling and for predictions of slope- to

catchment-scale mass sediment flux (Desmet

and Govers, 1995; Dietrich et al., 1993; Kamp

et al., 2003; Kirkby, 1990). Topographic infor-

mation in a DEM can be represented and stored

as: (1) a triangulated irregular network (TIN);

(2) discrete landform elements based on the

intersection of flow- and contour lines; or (3) a

grid. Here we examine gridded DEMs, which

despite drawbacks arising from higher demands

on computer memory (Claessens et al., 2005:

462; Yue et al., 2007: 161) nevertheless remain

the most commonly used data source for digital

terrain analysis because of their simple structure

and compatibility with other digitally produced

data (Wise, 2000: 1910).

Slope gradient and aspect, which are exam-

ined in this review, are two of the most basic

parameters affecting landscape-scale natural

processes. The accuracy of a slope map deter-

mines many of the predictions that geomorpho-

logists, engineers, hydrologists, and ecologists,

farmers or foresters make about the behaviour

of environmental variables, about natural

hazards, land-use potential, and for planning.

Errors on slope propagate to more sophisticated

DEM-derived terrain attributes such as slope

curvature, drainage networks, slope channel

lengths, slope widths or topographic wetness

indices. Although seemingly obvious, the

consideration of error distribution and error

propagation in DEMs has nonetheless been

neglected in routine DEM-based terrain studies,

perhaps because the appeal and immediacy of

DEM implementation in commonly available

software packages overrides any concern for

accuracy and error. The imperfections of

gridded DEMs, i.e. the fidelity with which they

represent the ‘real world’, have been extensively

explored in the last decades. Fisher and Tate

(2006) have described errors on gridded data sets

as the difference between a model value and the

true value, i.e. as deviations from reality (called

incorrect elevation values, misplaced elevation

values, data gaps, or missing values). True or

real values are absolute elevation measurements

collected on the Earth’s surface with the instru-

ments of terrestrial geodesy, whether satellite-

assisted or not. Like all numerical models

(Dietrich et al., 2003: 109–110), elevation

models are approximations of reality, but this

does not mean that sources of error are

unknowable, intractable or unquantifiable.

Any step in DEM production could be a source

of error, and error should be detected early in

the batch processes in order to avoid its

insidious propagation to subsequent steps and

its effects on the quality of the final product

(Darnell et al., 2008; Hengl et al., 2010).

The risk of introducing or ignoring distortions

is to generate bias in DEM-based quantitative

analyses of the Earth’s surface and in the accu-

rate characterization of landforms and processes.

The use of terrain models for environmental and

engineering applications, or even for heuristic

purposes, thus calls for an assessment of error

relating to data type, data precision, the way

2 Progress in Physical Geography

the data acquisition tools operate, and to the

comparative advantages of high ground-

resolution (g.r.) data such as LiDAR, SpotLight

(g.r.: 1 m, TerraSarX), Cosmo-SkyMed 4 (g.r.:

2–4 m) or Alos Prism (g.r.: 2.5 m) (Hengl

et al., 2010: 769). At a time when the number

and diversity of satellite, airborne and radar

platforms is on the increase due to the rapidly

evolving technology, this requires at least some

knowledge of data processing techniques and

appropriate means for meaningfully evaluating

the reliability of results used in terrain analysis.

Here we review the variety and relative per-

formance of existing approaches while empha-

sizing from the start that few studies have

embraced the analysis of spatial autocorrelation

as a prerequisite for developing error models.

Only random patterns exhibit no spatial autocor-

relation (Hunter and Goodchild, 1997; Oksanen

and Sarjakoski, 2005). DEM error, however, is

in essence already spatially dependent because

of the interpolation methods used in DEM

generation (Carlisle, 2005: 524; Liu and Bian,

2008: 308). This means that redundancy occurs

in spatial data and can bias the common assump-

tions of independence between values in many

statistical operations (e.g. least-squares regres-

sion analysis). A review of the techniques and

applications of spatial autocorrelation to DEM

analysis will accordingly occupy the later

sections of this paper and will focus around a

worked example providing new results for

illustrative and didactic purposes. Some recent

studies have addressed spatial autocorrelation

as a means of calibrating error models, but have

exclusively dealt with global autocorrelation

rather than with examining patterns of

local-scale variation (Hebeler and Purves,

2009; Lindsay and Evans, 2008; Liu and Bian,

2008; Wechsler and Kroll, 2006), and have

tended to focus on global data (e.g. GTOPO30,

SRTM: Holmes et al., 2000, Wechsler, 1999;

Wechsler and Kroll, 2006) and on the interpola-

tion of high accuracy data (e.g. LiDAR: Arrell

et al., 2008). Limited attention has been paid

to: (1) the potential for analysing local and

global spatial autocorrelation conjointly; (2) the

generation of continuous spatial representations

of error arising from DEM analysis; and (3) the

comparative sensitivity to error of DEMs

obtained from different sensors (optical, radar,

airborne). We will show how local spatial auto-

correlation analysis helps to locate the weak-

nesses of DEM structure, how it highlights the

spatial heterogeneity of error, and how, from the

use of local indices, it is possible to obtain error

maps useful to a wide range of terrain applica-

tions relevant to physical geography.

Accordingly, we will focus on three issues

that arise when different DEM sources are avail-

able for a given geographical area. (1) How can

the accuracy of elevation data be assessed in

such situations? (2) How can errors affecting pri-

mary topographic attributes such as topographic

slope be evaluated, and what confidence can be

attributed to such data for hydrological or

geomorphometric applications? (3) What error

mapping techniques are most advantageous for

detecting regions susceptible to error propaga-

tion and for visually displaying bias in DEM-

based scientific predictions? We first present a

review of current knowledge about DEM

accuracy and the tracking of error propagation

in grid processing. The Local Indicator of

Spatial Association (LISA: Anselin, 1995),

similar to calculating neighbourhood filters in

raster analysis, is then introduced and tested as

an illustration of how to select the most efficient

error model based on clustering results.

We test six DEM grids obtained from differ-

ent categories of sensors and focus on techniques

for mapping the spatial distribution of error

affecting topographic slope values. The chosen

study area is the Wasatch mountain front, Utah,

USA. The data sources are a LiDAR grid, i.e. an

airborne source; the composite CODEM,

NED10 and GDEM, each compiled from multi-

ple data sources, i.e. maps and/or optical and/or

radar; the ASTER DEM (AstD), an optical source;

and the SRTM, a radar source. Each possesses

Gonga-Saholiariliva et al. 3

initial statistical characteristics relating to the

elevation interpolation procedures, to the error

corrections relating to physical acquisition

parameters, and to altimetric configurations

imposed by the initial choice of geoid reference

frame. Although the ground resolutions of the

DEMs analysed also happen to range from

2 m to 90 m, what we test here is the quality

of the initial products. Ground resolution

remains a separate issue. Although the com-

plexity of any land surface is likely to decrease

with increasing cell size, with ground resolu-

tion potentially affecting error distribution as

a consequence, the first step in digital terrain

analysis still involves choosing among an

increasingly wide range of imperfect digital

products, whether these are readily available

on the market or purpose-generated by stereo-

photogrammetry. Whatever the original ground

resolution of the digital product, it becomes

advantageous to resample or degrade the DEM

to the ground resolution best suited to the

application required only after quality checks,

such as those proposed in this paper, have been

performed. After the most satisfactory digital

product available has been selected in this way,

work on terrain roughness indices, which mea-

sure the variability of elevation based on a slid-

ing window specific to each DEM (Grohmann

et al., 2009), is one way of addressing the

effects of scale on the analytical accuracy

affecting the landscape attributes of interest.

II Evaluation of accuracy and errorpropagation in gridded DEMprocessing

According to Hunter and Goodchild (1997: 36),

the notion of uncertainty covers undefined error

(i.e. when a true value is unknown) and struc-

tural error. Undefined error covers a range of

subcategories defined by Thapa and Bossler

(1992: 836), including blunders, gross errors

and systematic errors (see Appendix). Often

these cannot be removed, whether by repeated

observation or by recourse to a deterministic

search function, because they tend to evade any

clearly detectable pattern. After Burrough

(1986) and Wise (1998), Wechsler (1999: 3) has

identified three possible sources of structural

error: (1) errors due to an incomplete density

of observations, or to consequences of spatial

sampling; (2) measurement errors such as posi-

tional inaccuracy, which might arise from

improper orientation of the stereo-model or by

neglecting lens distortion in photogrammetric

procedures; and (3) random errors, which can

normally be detected after removal of gross and

systematic errors (see Appendix) and might be

due to data entry faults, observer bias, or

computer-generated numerical errors, interpola-

tion errors, or problems of classification and

generalization.

Finer differentiation among structural errors

distinguishes between horizontal (positional

accuracy) and vertical components (attributes

accuracy), respectively. These may be difficult

to separate (Ehlschlaeger and Shortridge,

1996) but an expression both in terms of the

metric accuracy of elevation and in terms of

planimetric deviation values is possible (Arrell

et al., 2008: 945). Planimetric and metric accu-

racy have long been evaluated and calculated

from routine statistical methods such as mean

error (ME), mean absolute error (MAE), error

in standard deviation units (S), or as the RMSE,

which is the square root of the variance of a dis-

tribution (e.g. Carlisle, 2005: 522; Erdogan,

2010; Fisher and Tate, 2006: 470; Wechsler and

Kroll, 2006: 1082; see Appendix). Within that

class of statistical test, RMSE-based comparison

of DEM elevation values with sample values

from another, more accurate source seems to

be the most appropriate way of providing error

estimation (e.g. Arrell et al., 2008: 945; Hirano

et al., 2003; Wise, 2007: 1353). However, a

major drawback of the RMSE is that results

merely provide one global value per DEM grid,

i.e. they ignore the finer-scale spatial variability

of the error (Wechsler, 2003: 58) that often

4 Progress in Physical Geography

contains a great deal of useful information for

terrain-based applications. Global indices are

thus insufficient (Li and Wong, 2010: 252) and

any useful study of DEM accuracy should also

investigate the spatial variation of error values

(Carlisle, 2005: 523; Wood, 1994).

Methods of error investigation in DEMs have

been widely explored (Fisher and Tate, 2006;

Giles and Franklin, 1996: 1165; Wechsler,

2003: 58; Wilson and Gallant, 2000: 15). The

simplest methods are based on criteria such as:

(1) differences in elevation between adjacent

points (e.g. Binh and Thuy, 2008; Hannah,

1981); (2) systematic detection by inspection

of anomalous values within a given moving win-

dow (e.g. Chaplot et al., 2006; Felicısimo, 1994;

Zhou and Liu, 2004b); (3) the minimization of

total error, which is the sum of errors originating

from the input data sources and operational

errors generated by data manipulation (e.g.

Walsh et al., 1987; Younan et al., 2002); or

(4) spectral (e.g. Tempfli, 1980) and elevation

histogram analyses (e.g. Carrara et al., 1997;

Holmes et al., 2000), which each focus on the

detection of structure in the statistical distribu-

tion of elevation values and on the identification

of unexpected variations interpreted as indicat-

ing systematic or isolated error. More complex

methods introduce remote sensing and/or

geostatistical processes. Goodchild (1982) and

Polidori (1991: 12–20) have investigated the

value of Brownian processes in a fractal terrain

simulation model for improving DEM accuracy.

Brown and Bara (1994) used semi-variogram

and fractal dimension to describe the pattern of

systematic errors. Lopez (1997, 2000) devel-

oped a method based on principal component

analysis (PCA) to locate random errors in DEMs

and to extract uncorrelated patterns. After

comparisons between Felicisimo’s method

(1994) and Lopez’s method (Lopez, 1997),

Lopez (2000) revised the PCA approach in order

to handle the significant spatial autocorrelation

observed in the data. By identifying gross and

random errors, systematic errors were avoided.

Finally, Arrell et al. (2008) identified and

removed artifacts associated with gross error

using Fast Fourier Transform analysis as a means

of discriminating between signal and noise.

Techniques of error location and visualization

have also been considered. Fisher and Tate

(2006) have reviewed research in this domain

and noted that DEM contour maps could be a

useful tool for error observation, with the

possibility of detecting gross errors and blun-

ders. The reliability of this method, however,

is not proven. The method of Hunter and Good-

child (1995) based on the use of the RMSE as a

threshold criterion for testing the accuracy of

elevation appears more useful. Other more

sophisticated methods have focused on error

visualization by using TIN or discrete data and

by displaying in 3-D the magnitude of local dif-

ferences and curvature errors from contoured

DEM data (Gousie, 2005). Perhaps a more work-

able method, proposed by Oksanen (2003:

2412), detects and visualizes gross errors in

gridded DEMs from relief shading because of its

ability to render a 3-D effect on elevation data.

Wise (2007: 1352) has argued that visualization

techniques used in error identification could also

be efficient at demonstrating strong links

between error patterns and interpolation-related

distortions. Wood (1996: 64) demonstrated this

by developing a spline-fitting algorithm on a

fractal surface that allowed errors and other

impacts of the algorithmic process to be detected

and visually displayed. More recent attention

has been paid to DEM interpolation algorithms

and their error-causing impacts (1) on elevation

in the case of high g.r. data (e.g. LiDAR: Arrell

et al., 2008: 944; Bater and Coops, 2009;

Lindsay and Evans, 2008; Liu, 2008), (2) on the

calculation of mathematical derivatives of ele-

vation (e.g. Florinsky, 1998; Jones, 1998; Liu

and Bian, 2008; Schmidt et al., 2003; Zhou and

Liu, 2004a), but also (3) on flow simulation and

its repercussions on drainage network extraction

(e.g. Li and Wong, 2010; Lindsay and Evans,

2008).

Gonga-Saholiariliva et al. 5

In terrain analysis, error propagation analyses

have usually focused on slope and aspect as test

variables. These mathematical derivatives of

elevation are recognized to provide a more use-

ful test of error location and propagation, and

thus for DEM quality assessment, than elevation

itself because of their greater sensitivity to error

(Wechsler, 2003; Wise, 2000: 1911). Liu and

Jezek (1999: 250) have listed three approaches

for treating the propagation of DEM error. These

include empirical approaches (e.g. Bolstad and

Stowe, 1994; Gatziolis and Fried, 2004;

Lindsay, 2006), analytical methods (e.g.

Florinsky, 1998; Heuvelink et al., 1989; Leick,

1990; Oksanen and Sarjakoski, 2005) and

computer simulations (e.g. Canters et al., 2002;

Ehlschlaeger, 1998; Hebeler and Purves, 2009;

Oksanen and Sarjakoski, 2005; Wechsler and

Kroll, 2006). (1) The empirical approaches rely

on a derivation of the spatial properties of error

based on a more accurate reference data set,

where results for a series of simulations are pre-

sented with error fields characterized by a range

of spatial properties (Lindsay and Evans, 2008:

1590). This compensates for the often unknown

spatial characteristics of error. (2) Analytical

methods can involve Taylor series expansion

(Heuvelink et al., 1989). Under the assumption

that error distribution follows a Gaussian distri-

bution, this approach constitutes the basis for

least-squares adjustment techniques (Hunter and

Goodchild, 1997: 36). (3) The simulation meth-

ods involve stochastic approaches such as Monte

Carlo techniques. This requires a number of

equally probable outcomes upon which selected

statistics are performed. Uncertainty is quanti-

fied by evaluating the statistics associated with

the range of outputs under the assumption of

spatial autocorrelation. A series of random equi-

probable error maps, one of which could be the

‘true’ error map, is provided. Such simulations,

however, are non-unique and do not ensure that

the true error map has been generated by the pro-

cedure (Wechsler, 2006: 2360; Wechsler and

Kroll, 2006).

The analytical treatment of terrain algorithms

is complicated and laborious (Lindsay and

Evans, 2008: 1590). Oksanen and Sarjakoski

(2005) confronted an analytical approach with

Monte Carlo simulations to evaluate the propa-

gation of error in slope and aspect simultane-

ously. They concluded that the two methods

produced similar results even though the data

had not been inspected for spatial autocorrela-

tion in the analytical approach. The work of

Oksanen and Sarjakoski (2005: 1016) and Liu

and Bian (2008: 308) has revived a debate about

the value of performing analytical manipulations

of gridded DEMs that disregard the spatial

autocorrelation of error — the most important

consideration being the choice of autocorrela-

tion parameter rather than the shape of the spa-

tial autocorrelation. These aspects are presently

examined.

III Spatial autocorrelation and theproduction of error maps

Spatial autocorrelation characterizes the correla-

tion between a variable and itself (univariate

autocorrelation), or between that same variable

and another variable’s values (bivariate

autocorrelation), in neighbouring cells. Spatial

autocorrelation functions are used to quantify

the spatial regularity of a phenomenon (a form

of spatial complexity) (Aubry and Piegay,

2001). It is based on the assumption that in all

data sets near things are more related than distant

things (Tobler’s first law of geography). Some

approaches avoid the condition of heteroscedas-

ticity, i.e. the necessity to consider variations

in error value as a function of variations in

another variable. This occurs in the case of

spatial autoregressive models (e.g. Hunter and

Goodchild, 1997) and of semi-variance analysis

(e.g. Carlisle, 2005; Giles and Franklin, 1996).

Other approaches, such as pixel swapping and

spatial moving averages using low-pass filters

(e.g. Fisher, 1998; Wechsler, 2006), introduce

spatial correlation without explicitly addressing

6 Progress in Physical Geography

error patterns or mapping error surfaces

(Hebeler and Purves, 2009: 6). Methods such

as sequential Gaussian simulation (e.g. Oksanen

and Sarjakoski, 2005) use global spatial autocor-

relation information without examining the finer

structure of the error. Given that the study of

spatial autocorrelation and of its potential conse-

quences in gridded DEMs has been unsyste-

matic, here we investigate the definitions of

spatial autocorrelation, its uses, how to deal with

it, and how error maps displaying local autocor-

relation clusters can reveal error propagation

phenomena.

Moran’s I (Moran, 1950) is historically the

first index to have been used for understanding

spatial autocorrelation, and still a standard.

In the context of gridded DEMs, Moran’s I is a

global indicator of spatial association and is used

to measure average deviations in slope or eleva-

tion for an entire grid. This global measure is

given by the following formula:

I ¼ NPi

Pj wij

�P

i

Pj wijzizjPi z2

i

ð1Þ

When there are N units, zi (zi ¼ xi � �x) is

the variable value at particular location, zj

(zj ¼ xj � �x) is the variable value at another

location, is the mean of the variable, and wij

expresses the level of interaction (connectivity)

between units i and j. This contiguity matrix

compares the sum of the cross-products of

values at different locations, two at a time,

weighted by the inverse of the distance between

the locations (if i is next to j, the interaction

receives a weight of 1). Moran’s I ranges

betweenþ1 (strongly positive) and –1 (strongly

negative spatial autocorrelation). A value of

0 indicates a random spatial pattern.

From this index Anselin (1995) later

elaborated the LISA, which can also be used

on continuous DEM grid surfaces. Here we use

Moran’s I and LISA as techniques for

visualizing error propagation, and for adding

information about spatial autocorrelation not

provided by Moran’s I. LISA is used for estimat-

ing the local contribution of each pixel value to

the global index. Local indicators are based on a

gamma statistic (Getis, 1991) given by the fol-

lowing formula (Oliveau, 2005):

G ¼X

i

XjwijWij ð2Þ

where Wij ¼ ðzi�zÞðzj�zÞPiðzi�zÞ �

nm

and Gi ¼P

j wijWij

In these formulae, variables are as in equation

(1), n is the total number of subjects in the cross

section (P

i), m is the neighbourhood pair num-

ber (P

i

Pjwij), and Wij is a measure of relation-

ship in attribute space between data sites i and j.

LISA measurements take into account spatial

and numerical similarities between contiguity

matrices. The LISA of each observation gives

an indication of the extent of statistically signif-

icant spatial clustering of similar values around

the observation point. The LISA is obtained by

local spatial clustering procedures using the

following equation (Anselin, 1995):

Ii ¼ zi

Xj

wijzj ð3Þ

with the sum of local indicators:Xi

Ii ¼X

i

zi

Xj

wijzj ð4Þ

where zi and zj are expressed in deviations from

the mean, the summation of is such that only

neighbouring values j 2 Ji are included, and is

the contiguity matrix of weights. The sum of

local indices is proportional to Moran’s I from

g¼P

i

Pj

wijm2; with m2 ¼P

i

z2i =n. LISA

results can be standardized by dividing the

LISA by. The LISA helps to identify local out-

liers (where negative local spatial autocorrela-

tion shows up on the cluster map as high–low

or low–high occurrences, meaning that the

Gonga-Saholiariliva et al. 7

variable of interest exhibits dissimilar values

among its neighbours); and local clusters

(where positive local spatial autocorrelation

shows up on the cluster map as high–high or

low–low occurrences, meaning that the variable

of interest shows values that are similar to its

neighbours). Local deviations from the global

pattern of spatial autocorrelation can also be

delineated.

Bivariate LISA-type cluster analysis explores

the spatial distribution of DEM errors affecting

elevation and measures its transmission to errors

affecting slope gradient. It follows a scheme of

pattern recognition (Jacquez, 2008: 400) made

up of five components: (1) a null hypothesis,

H0, which describes the spatial pattern expected

when the alternative hypothesis is false (i.e. an

absence of significant clustering of errors in

slope at sites where errors in elevation occur);

(2) an alternative hypothesis, HA, which states

that the clustering of errors in slope closely

matches the spatial distribution of errors in

elevation; (3) a statistical significance test that

quantifies the spatial pattern of the data; (4) a

neutral spatial model, which is a numerical

device for generating the reference distribution

based on a Monte Carlo simulation (achieves

randomly operated permutations of values);

(5) a reference distribution, which is the

distribution of the statistical test scores when

the null hypothesis is true, and is set by the

p-values, which in statistics measure strength

of evidence. For the LISA tests, pseudo

p-values are computed as (mþ1) / (rþ1),

where r is the number of replications and m

is the number of instances where a statistic

computed from the permutations is either

equal to or greater than the observed value

(in the case of positive LISA scores), or equal

to or smaller than the observed value (for

negative LISA scores) (Anselin, 1995). Results

are filtered with the choice of an arbitrary

significance threshold, a (0.05, 0.01, 0.001, or

0.0001), which determines confidence intervals

used for assessing results.

IV Patterns of error in griddedDEMs: a worked example fromthe Wasatch Mountains, Utah,USA

Among the few existing studies to have explored

the local structure of spatial autocorrelation in

raster analysis, Goovaerts et al. (2005) detected

anomalies on high-resolution hyperspectral

imagery while testing the significance of LISA

values through randomization. Erdogan (2010)

investigated spatial variation as well as spatial

associations of error by comparing a range of

interpolation methods. Monckton (1994) used

Moran’s I to simulate the spatial structure of

error in elevation data. Results, however, were

inconclusive, probably because of the spatial

lags (which exceeded 250 m) that were used in

dealing with spot heights on Ordnance Survey

maps (Carlisle, 2005: 524).

In the following worked example designed to

illustrate the power of autocorrelation analysis

as a tool in DEM error evaluation, we performed

tests on a *29 km2 study area encompassing

part of the Weber Segment, one of the longest

(61 km) and seismically most active segments

of the Wasatch normal fault in Utah (Figure 1).

On this segment, faceted spurs and their triangu-

lar mountain-front hillslopes are well preserved

(McCalpin and Berry, 1996; Nelson and Perso-

nius, 1993; Petit et al., 2009; Zuchiewicz and

McCalpin, 2000). The geomorphology of fault-

generated range fronts is reflected in the drai-

nage basin morphology as well as in the faceted

spur geometry. In order to extract precise terrain

parameters with minimal bias, DEM evaluation

was examined through slope analysis. Using

both LISA and Moran’s I, the methodology aims

to single out the DEM best suited to geomorpho-

metric or hydrological applications. The best

DEM should provide the model with the

smallest error, i.e. with the highest accuracy and

the lowest level of error clustering.

The grid characteristics of the six DEMs

available for the study area (see Introduction),

8 Progress in Physical Geography

with differing column and row grid-cell sizes,

are as follows (Figure 2) – LiDAR (2 m): 2979

� 2370; CODEM (5 m): 1260 � 857; NED10

(10 m): 630 � 474; AstD (15 m): 420 � 316;

GDEM (30 m): 210 � 857; SRTM (90 m):

70 � 53. General statistics are usually available

for each DEM from supporting documentation,

in which major errors and correction processes

are mentioned. Out of the six in this study,

only the AstD was generated directly from raw

data by the authors using digital stereo-

photogrammetric techniques summarized in

Figure 3. The operational error we use is the dif-

ference between the z values provided by each of

the five tested DEM grids on the one hand, and

the z values provided by the LiDAR data on the

other, the latter being closest to any ‘true’ eleva-

tion measurement. Any observed difference in

z is interpreted as being due to one or several

of the error categories defined in the Appendix.

Those refer to the various defects originally

present in the gridded elevation data. Their

distribution pattern across the grid remains

unknown until we proceed to the mapping tech-

niques proposed later in this worked example.

They depend on the type of sensor from which

the data was collected, and how this data was

preprocessed before making it available for use

by geographers.

1 Vertical accuracy

The first approach to error estimation relied on

the RMSE, which integrates errors relating to

interpolation methods and to geographical posi-

tion in x, y and z. Vertical accuracy assessment at

pixel scale was first calculated from the LiDAR

model (RMSE & 0.086 m) with a regular

sample extraction interval of 100 m, each point

extracted (3000 for each DEM) being a substitute

Figure 1. The Weber Segment in the Wasatch Mountains, Utah, USA. Elevations range from 257 m to 3618 m.The 2-D and 2.5-D views highlight the active normal boundary fault with its triangular faceted spurs.

Gonga-Saholiariliva et al. 9

for ground GPS information (see Appendix for

details). Given that, for comparison purposes,

the grids could not be used with their original

characteristics, DEM resampling was also nec-

essary. Accordingly, for the RMSE calculation

resampling procedures were performed on each

DEM by transforming cell size to the reference

size, i.e. 2 m. To avoid stair-stepping and

jagged artifacts in the resampling process, a

bicubic spline resampling method was used

instead (see Appendix) (Keys, 1981). The

resampled grids were used for a comparative

assessment of vertical accuracy between

DEMs. In order to obtain percent values of

transformation from the resampled DEM, a

new RMSE value, noted RMSEc, was

calculated. The formulae given in the Appendix

were applied to the entire DEM grid and assess-

ment of the vertical accuracy of surface objects

was carried out using the RMSEc criterion.

Figure 4 and Table 1 illustrate the distribution

of error values in elevation among the six DEM

grids. Quartile distributions (first and third)

describe a band of +8 m for minimum and

+22 m for maximum deviations from the mean

elevation (Table 2). Overall, DEM height distri-

butions exhibit a similar pattern despite the

fact that the elevation sources are different

(i.e. airborne, radar, optical or compilation data).

The more outstanding exception is the AstD,

Figure 2. Extract of each studied DEM representing the different grid-cell sizes used in the worked example.(A) LiDAR data (2 m, available at http://gis.utah.gov/elevation-terrain-data/2-meter-bare-earth-lidar). (B)CODEM (5 m, http://gis.utah.gov/elevation-terrain-data/5-meter-auto-correlated-elevation-model-dem).(C) NED10 (10 m, http://gis.utah.gov/elevation-terrain-data/national-elevation-dataset-ned). (D) ASTERDEM (15 m, http://www.echo.nasa.gov/reference/astergdem_tutorial.htm). (E) GDEM (30 m http://aster-web.jpl.nasa.gov/gdem.asp). (F) Shuttle Radar Topography Mission, or SRTM (1 arc second, i.e.*90 m, http://www.cgiar-csi.org/data/elevation/item/45-srtm-90m-digital-elevation-database-v41).

10 Progress in Physical Geography

which presents significant offsets with respect to

the other DEMs. These offsets attain a difference

in values of about +15 m (roughly equivalent to

one AstD pixel) for height and RMSE measure-

ments as well as for statistical moments. This

exception is ascribable to the initial AstD extrac-

tion process, in which the precision of the DEM

tools used was greater than half a pixel (see Fig-

ure 3). The initial RMSE for the AstD (11.8 m)

was accordingly too large. The global RMSE

should be about half a pixel (7.5 m). Additional

ground control points were checked in order to

newly interpolate the AstD, finally providing it

with elevation ranges similar to the other DEMs,

i.e. [1355–1298 m]. As a result, the new value

obtained for the AstD RMSEc (Table 2) fell into

line with the overall RMSEc trend for the other

DEMs, i.e. about 12.9 m. Thus RMSEc results

(incorporating new values given in brackets for

the AstD) yield results similar to those obtained

for the other DEMs: the larger the grid cell size,

the higher the RMSEc [þ5.7;þ18.4] and RMSE

[þ4; þ16 m]. The key conclusion at this stage

is that whatever the DEM’s data source,

Figure 3. Flow chart explaining the ASTER DEM extraction process from the stereoscopic VNIR 3N andVNIR 3B bands. VNIR: visible near-infrared bands; SWIR: short-wave infrared bands; TIR: thermal infraredbands. VNIR, SWIR and TIR are ortho-rectified bands (geometric corrections). GCP are ground controlpoints extracted from the ortho-rectified VNIR bands. The RMSE is calculated for the (x,y) plane coordinatesand the vertical plane z.

Gonga-Saholiariliva et al. 11

DEM generation does not produce large errors in

z. Values for the SRTM and GDEM also fall

close to the reference LiDAR value and thus

perform almost as well as the most accurate

reference grids despite their comparatively

lower initial resolution.

As a further test, linear regression analysis

was carried out between the elevation data of

each DEM and the LiDAR reference grid,

respectively (Figure 5). Overall, the DEMs pro-

duced a goodness-of-fit, measured by the deter-

mination coefficient r2, close to þ1, with

intervals ranging betweenþ0.9979 for the worst

fit and þ0.9999 for the best. The CODEM and

NED10 show the strongest correlation with the

LiDAR, followed by the AstD and the GDEM.

Model fits reveal outliers at different locations

in the AstD, GDEM and SRTM scatter plots.

The outliers reflect some of the error present

within the DEM, which appears largest in the

case of the AstD. However, neither data scatter

in the linear regression models nor vertical accu-

racy constitute sufficient criteria to achieve a full

definition of DEM error because the error is

expressed globally (lacks any indication of spa-

tial distribution), and because when untested

spatial autocorrelation occurs in data, linear

regression models tend to overestimate both

strength of correlation and precision.

2 Residual maps and computation ofprimary topographic attributes

One approach to locating error distribution

patterns visually can be achieved through the

creation of so-called deterministic error

surfaces. These can be generated for a DEM

by using reference data assumed to possess

higher accuracy than the DEM (Hebeler and

Purves, 2009), in this case the LiDAR, but

equally ground control points or other surface

model data (Holmes et al., 2000). In this way,

maps of differences between the reference grid

and other grids, termed here residual maps of

either elevation or slope gradient, can provide

an understanding of error sources and form a

basis for further analysis (Van Niel et al.,

2004). Residual maps were generated by apply-

ing an algebraic operation of pixel by pixel

subtraction between each DEM and the LiDAR.

In order for these tests to be meaningful, degra-

dation of the LiDAR to the relevant ground

resolutions, i.e. from 5 m to 90 m, was necessary.

This operation resulted in a series of numerical

images representing Re, the pixel-wise residual

value (Figure 6).

Figure 6 shows that (1) the range of Re stan-

dard deviations is similar (0–11 m) for the three

most accurate DEMs, namely CODEM, NED10

and AstD. (2) The SRTM and GDEM show the

greatest deviations (up to 41 m) from the refer-

ence surface, which is in keeping with the lower

ground resolution of these two data sets. Further-

more, the bicubic spline resampling procedure

resulted in artificial smoothing of the surfaces,

an aspect clearly apparent in the SRTM residual

map (Figure 6E). (3) High values of standard

deviation tend to cluster around the same sites

in the landscape whatever the grid-cell size, with

detectable variations in the sharpness of the dis-

continuities. (4) The highest values, i.e. greater

Figure 4. Statistical results for vertical accuracyassessment. Cumulative elevation (in %) versus ele-vation values for the studied DEM grids. The offsetin the ASTER DEM curve reveals a shift (systematicerror) requiring systematic correction to achieverealignment with the other grids (see text).

12 Progress in Physical Geography

than 9 m, follow narrow valley floors in the

mountainous relief. Value distributions differ

in the piedmont, where they are located along

wider valleys. (5) High deviations mostly affect

slopes and valley bottoms. Where hillslopes are

affected by anomalous deviations, errors are

likely to propagate to higher order topographic

attributes (e.g. slope curvature), and hence

adversely affect predictions made from the

DEM for applications such as hydrology or nat-

ural hazards. (6) Flat areas, which predominate

in the southern part of the piedmont zone,

display values that are close to the minimal

deviation, i.e. 0 m. This is most visible with the

higher grid-cell sizes in Figure 6, B, C and D.

Flat areas, which are often problematic in hydro-

logical network extraction (e.g. Tarboton, 1997),

thus seem to be less affected by high deviations.

However, this finding does not exclude the

possibility that erratic features, such as pits and

holes, could be present in this kind of area.

Figure 6E (SRTM) shows that regions affected

by large error are larger, form continuous

patches, and are concentrated in valleys and on

elevated ridges whereas they are absent or more

discontinuous in the other DEMs.

The following step in the procedure involved

the computation of slope gradient values in order

to assess error patterns in slope gradient. Slope

maps depend on the choice of algorithm and

on grid-cell size. We used the Zevenbergen and

Thorne (1987) algorithm due to existing

consensus over its relative robustness and con-

sistent behaviour in various terrain geometries

(e.g. Florinsky, 1998; Jones, 1998; Liu and Bian,

2008; Zhou and Liu, 2004b). Calculations were

Table 1. Distributions of elevation in the six DEMs.Min and Max are the minimum and maximum values; 1stand 3rd Qu. are the first and third quartile. For the ASTER DEM, corrected values (see text) are in brackets.For other DEMs, tabulated values are initial values.

Elevation value (m)

Data source SRTM GDEM ASTER NED10 CODEM LiDAR

Min 1355 1355 1340 (1355) 1357 1360 13511st Qu. 1775 1769 1754 (1769) 1769 1776 1771Median 2076 2073 2057 (2072) 2070 2078 2071Mean 2073 2069 2051 (2065) 2067 2073 20693rd Qu. 2366 2358 2338 (2352) 2358 2363 2359Max 2897 2884 2871 (2886) 2892 2900 2900

Table 2. RMSE results for the six DEMs.STD is the standard deviation and ME is the mean error for eachDEM grid. RMSE is the initial RMSE with respect to the reference geoids used for the production of eachDEM.RMSEc is the new RMSE calculated with respect to the LiDAR reference grid.

Error values (m)

DEM Grid cell size STD ME RMSE RMSEc RMSEc–RMSEr

LiDAR (reference) 2 387.3 Ø þ0.86 Ø ØCODEM 5 408.3 4 þ4 þ5.6 þ1.6NED10 10 406.5 1.9 &7 þ6.7 þ0.3ASTER DEM 15 403.9 21.8 (3.1) þ11.8 (þ9.4) þ25.9 (þ12.9) þ14.1 (þ3.5)GDEM 30 405.1 0.5 þ9.35 þ13.6 þ4.25SRTM 90 407.0 3.7 [þ9;þ16] þ18.4 [þ2.4;þ9.4]

Gonga-Saholiariliva et al. 13

Figure 5. Scatter plots of elevation values (z) produced for various permutations of the studied DEM grids

Figure 6. Maps of Re standard deviation for elevation in the Weber Segment (colour version available online).The residual values in maps A to E express the difference between z values in each DEM and the z valuesprovided by the LiDAR reference. (A) CODEM, cell size¼ 5 m. (B) NED10, cell size¼ 10 m. (C) ASTER DEM,cell size ¼ 15 m. (D) GDEM, cell size ¼ 30 m. (E) SRTM, cell size ¼ 90 m.

14 Progress in Physical Geography

carried out on each of the six grids (Figure 7).

Slope values range from 0� to 89.9� (Table 3).

Lower values (0–1�) indicate flat interfluves or

valley bottoms, whereas the highest slope angles

represent unusual terrain values, typically

greater than 35� and located on faceted spurs.

Standard deviation and mean slope decrease

with grid-cell size, in line with previous research

on this topic (e.g. Kienzle, 2004; Warren et al.,

2004) but unlike results obtained for DEMs

extracted from contour lines (e.g. Ziadat, 2007).

The distribution of slope values is representa-

tive of the general slope trend for all the DEMs

(Figure 8). Values of 0–30� represent 98% of the

slope frequency range in the SRTM, whereas

other DEMs exhibit a greater proportion of

Figure 7. Slope maps of the Weber Segment as generated from each of the DEMs (colour version availableonline). (A) LiDAR, cell size ¼ 2 m. (B) CODEM, cell size ¼ 5 m. (C) NED10, cell size ¼ 10 m. (D) ASTERDEM, cell size ¼ 15 m. (E) GDEM, cell size ¼ 30 m. (F) SRTM, cell size ¼ 90 m.

Table 3. Slope values for the six DEMs.Slope-angle range is the interval between minimum and maximumslope; mean slope angle is the arithmetic mean of the slope value distribution; Std slope is the standard devia-tion of that distribution.

DEM Slope-angle range(degrees) Mean slope angle(degrees) Std(degrees)

LiDAR 0–89.9 26.33 1.52CODEM 0–75.8 26.08 1.37NED10 0–89.7 26.26 1.34ASTER 0–89.6 23.43 1.29GDEM 0–89.4 23.14 1.30SRTM 0.3–42.2 22.37 1.06

Gonga-Saholiariliva et al. 15

extreme values (for example, slope angles

greater than 40� account for 10% of the slope

frequency distribution). Field data have revealed

that the Weber Segment mountain front is char-

acterized by slope angles never exceeding 43�

(Eblen, 1995; Stock et al., 2009), with mean val-

ues close to 25–26�. The highest values, between

45� and 90�, are thus anomalous and represent

errors due to DEM uncertainty, which were propa-

gated to the slope map during transformation of the

elevation data. The geographical location of large

errors in slope coincides for each DEM with that of

the largest errors in elevation (Figure 7). For

angles up to 45�, slope values also increase with

cell size. However, anomalous values are fewer

for the larger grid-cell sizes than for the smaller

ones (CODEM, NED10, AstD). This also means

that the larger the grid-cell size, the greater propor-

tion of values <45� and the fewer outliers present.

Although results are so far directly ascribable

to misfits originally affecting z, results differ

between the mean and lower slope values. The

lower 50% of the distribution frequency is

marked by several offsets between the highest

and smallest grid-cell sizes (Figure 8): for exam-

ple, contrary to the expectation that error is

consistently proportional to grid-cell size,

between 0� and 25� the slope distribution values

of the CODEM increase faster than for the

NED10. Also, in the low range (i.e. 0� to less than

15�) the AstD curve locally crosses the SRTM

and GDEM curves. Consequently, lower and

mid-range slope values appear to reflect isolated

discrepancies in the DEM elevation structure.

The standard deviation of residuals (difference

between reference and calculated slope) was also

calculated (Figure 9), illustrating for each DEM

the deviation in mean distribution values for each

grid. In the six DEMs, the thalwegs of larger

valleys are well marked. Note also that the loca-

tions of slope errors are almost identical to the

locations of the errors in elevation except on

the main central facets of the Weber Segment

(Figure 9). The higher the terrain resolution, the

more sensitive the gridded data to errors in slope,

which are thus also linked to terrain roughness

characteristics. Accordingly, slope deviations

range between 0� and 14� for the larger grid-

cell sizes (SRTM, GDEM) and between 0� and

25� for the smaller (LiDAR, CODEM, NED10,

AstD; Figure 9).

The strength of the relationship between

errors in slope and errors in elevation (Figures

6 and 9) was tested by performing a linear

regression between elevation (Re on elevation)

and slope residuals (Re on slope). A non-linear

relationship between those two variables was

found (Figure 10). However, the form and inten-

sity of the relationship is biased by several out-

liers that stretch the scatter plot in several

directions. Determination coefficients with

values <0.3 for the SRTM and CODEM and

<0.1 for the AstD, GDEM and NED10, indicate

weak to no dependency between the errors in

slope and in elevation. In other words, only

10–30% of anomalous slope values (in particular

among the lower and mid-range values) can be

explained by anomalous elevation values.

Figure 8. Comparative distributions of slope values for each DEM grid

16 Progress in Physical Geography

Such a conclusion, however, is potentially

spurious because r2 is not a robust criterion for

explaining either the numerical or the spatial

behaviour of errors affecting slope. Accordingly,

as a further tool for evaluating whether higher

error affecting elevation was positively correlated

with higher error affecting slope, we inspected the

data grids for spatial autocorrelation.

3 Spatial autocorrelation analysis

Calculation and 2.5-D mapping of Moran’s I and

LISA were used to identify areas with a high risk

of error caused by pixels being surrounded by

similar areas or by spatial outliers, and as tools

for visualizing the location of uncertainty.

A spatial weighting matrix was used for each

variable, and was defined at varying distance

lags corresponding to the respective DEM reso-

lutions. The weighting process was carried out

using Rook’s case–contiguity kernel, with sym-

metrical spatial weights wij ¼ 1 when locations

i and j are adjacent, and wij ¼ 0 in other cases.

Rook’s case considers the neighbourhood of

four locations adjacent to each cell (Elobaid

et al., 2009). A Monte Carlo randomization pro-

cedure allowed us to assess significance levels

by performing 999 permutations (i.e. tests) for

each grid (Table 4).

The randomization test avoids relying on

assumptions about data distribution and devel-

oping statistical moments under a null hypoth-

esis, and avoids testing all types of statistic

analytically (Aubry and Piegay, 2001). Signifi-

cance levels are given by the p-values (arbitrary

threshold), are calculated for each cluster, and

Figure 9. Maps of Re standard deviation for slope angle in the Weber Segment (colour version available online).Key as in Figure 6.

Gonga-Saholiariliva et al. 17

allow significant autocorrelation values to be

identified. After trial comparisons with other

thresholds (0.0001, 0.001, and 0.01) for each

DEM, p-value significance levels were set to

0.05 because this configuration yielded the high-

est and most sharply differentiated levels of data

clustering. Table 4 summarizes global clustering

levels for bivariate and univariate variables,

each time for a run of 999 permutations.

The CODEM grid, with a Moran’s univariate

I of about –0.06, illustrates a situation of low

negative spatial autocorrelation, implying that

neighbours are dissimilar and that error values

affecting elevation do not have a unique source

or cause. In contrast, error values in the other

DEMs obtain a Moran’s index close to þ0.9,

indicating a strong spatial link between errors

in elevation. Error in slope systematically attains

values close to þ1, with a high level of spatial

dependence showing strong positive spatial

autocorrelation values. Where high spatial auto-

correlation scores were observed for elevation,

values likewise increased for slope, thus con-

firming a mechanism that drives the dispropor-

tionate (i.e. non-linear) propagation of errors in

elevation to errors in slope gradient. The bivari-

ate Moran’s index, which describes the magni-

tude and relationship of global error between

elevation and slope, is characterized by positive

values but none of these are close to þ1.

The strongest positive relations are observed for

the GDEM (þ0.4870) and SRTM (þ0.3048)

grids, whereas the lowest occur in the AstD

(þ0.1125). Results are thus similar to the uni-

variate values except for the AstD. The SRTM,

GDEM and NED10 present a high level of

clustering, with the CODEM ranking last.

The LISA helps to bring out local magnitudes

of error clustering. Each observation provides a

measure of the extent of statistically significant

Figure 10. Scatter plots highlighting links between errors in elevation and in slope for various permutationsof the studied DEM grids

18 Progress in Physical Geography

spatial clustering of similar error values around a

data point. LISA results reveal a high level of

differentiation within each DEM (Figure 11).

High–high and low–low clusters define a high

error value surrounded by high values and a low

value surrounded by low values, respectively.

Such situations are present in the SRTM, GDEM

and to a lesser extent the NED10 grids. A higher

risk of error propagation occurs where high–

high values are found. Relative stability, i.e.

a low risk of error propagation to higher order

topographic derivatives, prevails in low–low

situations. High–high clusters are found in the

same locations as in the univariate elevation and

slope error maps. In those cases, therefore, the

LISA maps merely corroborate previous results.

Uncertainty in slope is thus highly correlated

with initial elevation, and thus with error in

elevation, which in turn conditions the spatial

distribution of errors.

Table 4. Autocorrelation results for errors on elevation and on slope.Moran’s I elevation and Moran’s Islope are the univariate Moran’s I calculations for the dependent (elevation) and independent (slope) vari-ables, respectively. Bivariate Moran’s I refers to the errors in elevation and in slope and is the result of themultivariate LISA test.

DEM r2 Moran’s I elevation error Moran’s I slope error Bivariate Moran’s I

CODEM 0.24392 –0.0573 0.9149 0.1403NED10 0.07332 0.7523 0.9226 0.2772ASTER 0.01236 0.6729 0.9218 0.1125GDEM 0.09374 0.7702 0.9261 0.4870SRTM 0.20950 0.8869 0.9149 0.3048

Figure 11. Uncertainty cluster maps based on the LISA (colour version available online)

Gonga-Saholiariliva et al. 19

For grids other than the SRTM, GDEM and

NED10, LISA maps show somewhat different

results. AstD maps show all the peak occur-

rences already found in the SRTM and GDEM

grids, but low–high and high–low clusters are

more prominent between high–high clusters.

High–high levels are typical of major thalwegs

and low-lying areas. In the case of the NED10,

error distribution patterns bring out geomorpho-

logical features (particularly ridges and escarp-

ments) more sharply than the SRTM and the

GDEM. This arises whatever the level of p-value

significance (tested from 0.0001 to 0.05). The

CODEM stands as the exception: uncertainty

already detected in previous maps (Figures 7 and

9) is well represented in the CODEM LISA map

but fails to show the high or low clustering levels

encountered in the other DEM grids. Clustering

is less intense (note the predominance of half-

tones), illustrating uncertainty in the data, i.e.

uncertainty not linked to terrain characteristics

and presumably caused by other unknown fac-

tors which would remain to be investigated.

IV Summary and conclusion

The aims of this article were: (1) to review suit-

able tools for assessing error in remotely sensed

data and for calibrating DEM studies that focus

on primary topographic attributes; (2) to investi-

gate a statistical tool more often used in ecology

and human geography than in digital terrain

analysis, namely LISA clustering, which helps

to produce and evaluate error maps in order to

track DEM error propagation; and (3) to assist

in decision-making when selecting among a

range of DEM products that are readily available

for a chosen study area. In a worked example

involving terrain analysis in the Wasatch Moun-

tains of Utah, we were able to use LiDAR data as

a surrogate for field control points. We tested

five other, freely available, DEM grids against

this very high-resolution reference grid.

Despite the existing catalogue of methods for

removing various categories of error in gridded

data, the overriding conclusion at present is that

errors can never be totally erased from DEMs.

Empirical approaches work best when

higher-quality reference data are available for

comparison and allow a range of statistical rank-

ing procedures to be carried out. Initially, basic

quantitative tools such as the RMSEc were of

little use in determining DEM accuracy. It was

shown that the larger the grid-cell size, the lower

the accuracy, with nevertheless a fair degree of

reliability in z whatever the sensor type and

whatever the original data ingredients entered

into the DEM. The CODEM, NED10, AstD,

GDEM and SRTM could therefore be used

almost indiscriminately for basic displays of

topography, depending simply on choice of

resolution for graphic output.

As an improvement on RMSE procedures we

generated a series of residual elevation and

residual slope maps (Re), which provided an out-

line of the spatial distribution of errors based on

calculations of the standard deviation of Re.

Two main conclusions arose from these find-

ings: (1) errors in slope and in elevation are

linked, but with a variable strength of depen-

dency as shown by linear regression results; and

(2) the SRTM, GDEM and NED10 stand out for

their lower error scores. Irrespective of their

ground resolutions, the CODEM and the AstD

presented either more extreme error values or

more confused results. This finding argues

against preferring these data sources for

morphometric or hydrological applications. Re

results emulated a further approach in error

analysis using Moran’s I and LISA maps.

This approach has revealed that the intensity of

spatial autocorrelation is not linked to grid-cell

size. It is more closely related to the intrinsic

characteristics of DEM products and to proce-

dures dealing with slope gradient calculation.

The LISA maps also help to define error hot- and

coldspots, i.e. high–high and low–low clusters.

Among the five DEMs analysed in this study, the

autocorrelation tests finally allowed the NED10

to stand out as the DEM least affected by the

20 Progress in Physical Geography

presence of error and its propagation. Like the

CODEM, the NED10 is a hybrid DEM compiled

from a range of digital and non-digital sources

rather than from unique radar or optical sources,

and also happens to be of a lower native resolution.

The trends and topics reviewed here have

emphasized the value of exploring gridded data

structures for detecting spatial error patterns

because results can be used for ranking DEM

quality. Such procedures seem critical for most

practical terrain applications in natural catch-

ments, but also for using the DEM grids as input

substitutes for ‘real topography’, i.e. as initial

conditions for running predictive surface pro-

cess models (e.g. Codilean et al., 2006) and

observing how the landscape evolves under

given conditions (heuristic terrain modelling).

Despite its advantages, the LISA still remains

a semi-quantitative tool (values express varia-

tions in clustering as ‘high’ or ‘low’). A move

towards more fully quantitative characteriza-

tions of error in DEMs, where, in addition to its

location, the intensity of error is provided

(deviation of elevation in metres, and/or slope

in degrees), could perhaps be achieved by

coupling the LISA with other error modelling

methods such as geostatistical simulations based

on kriging, or geographically weighted regres-

sion (Fotheringham et al., 2002).

Appendix

(1) Definitions of error

(a) Artifacts, blunders and gross errors

Reuter et al. (2009) defined artifacts as erratic

features that appear on a DEM as holes, ghost

lines, stripes and so on. Artifacts and associated

errors could even have been invisible in the

input DEM but can be inferred from unrealistic

values occurring on elevation derivative maps.

Blunders are vertical errors associated with data

collection. They can be mistakes caused by

misreading contours, by transposing numerical

values, by erroneous correlations, or by careless

observations (Wechsler, 2003). Gross errors

occur when a measurement process is subject

to occasionally large inaccuracies and generates

data outliers. It includes topographical misinter-

pretations, and long linear features along some

boundaries of the calculation window used in

DEM interpolation (Oksanen, 2003). Some

techniques such as ‘data snooping’ are efficient

at detecting them (Karras and Petsa, 1993), and

inspection of a shaded relief map can also help

(Gousie, 2005).

(b) Systematic errors

Sytematic errors result from deficiencies in

measurement or processing and affect the accu-

racy of the final data, i.e. the degree to which the

final data agree with reality. Systematic errors

are also identified as residual systematic error

surfaces caused by slightly inaccurately esti-

mated lens distortion parameters (Wise, 2000),

by translation effects, or false features such as

phantom tops, ridges, benches or striations

(Raaflaub and Collins, 2006). In general, they

reflect bias inherent in the data collection

method (Reuter et al., 2009).

(c) Random errors

Random errors result from mistakes such as

inaccurate surveying or improper recording of

elevation information. Often associated with

signal noise, random errors are extracted after

known blunders and systematic errors have

been removed (Podobnikar, 2009; Reuter

et al., 2009; Wechsler, 1999, 2006; Zhang and

Cen, 2008).

(d) Uncertainty

Uncertainty is a term employed when a mod-

elled value deviates from its true value, with a

confidence interval (Hebeler and Purves,

2009). Lack of knowledge about measurement

reliability and about its representation of true

values is referred to as its uncertainty and is a

measure of what we do not know (Wechsler,

1999, 2006). DEM uncertainty thus covers lack

of knowledge about parameters, model assump-

tions, and scenarios, i.e. spatial evolution mod-

els (Liu and Bian, 2008).

Gonga-Saholiariliva et al. 21

(2) Equations and formulas

(a) Root mean square error and mean error

RMSEðzÞ ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1

n

Xn

i¼1

zdem � zref

� �?

s

ME ¼ 1

n

Xn

i¼1

zdem � zref

� �

where zdem is DEM elevation, zref is reference

elevation, and n is the number of test points.

(b) Bicubic spline formula

pðx; yÞ ¼X3

i¼0

X3

j¼0

aijxiyi

where p is the interpolated surface, x and y are

rectangular coordinates of point (i,j), and a is the

polynomial coefficient. This interpolation

maintains the continuity of the function and its

first derivatives (both normal and tangential)

across cell boundaries.

(c) Residual maps and standard deviation

algorithmsPixel by pixel algebraic subtraction

of elevation values provides a series of residual

values (Re) that can be mapped:

Reði;jÞ ¼ zði;jÞdem � zði;jÞref

where i and j are lines and columns of the grid,

and and are the elevations of the DEM and the

reference grid, respectively. The standard

deviation of Re is calculated as follows:

Std ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPni¼1 zdem � zref �ME� �2

n� 1

s

where ME is the mean error of the DEM. Stan-

dard deviations were calculated each time for

the entire image, cell by cell, yielding a map

of Re values in which deviations in elevation

values from those of the reference model can

be located.

Acknowledgements

We are grateful to Editor-in-Chief Nicholas Clifford

for steering our manuscript through a high-quality

reviewing process, with fine reports from several

anonymous referees who greatly improved the scope

and structure of this contribution.

References

Anselin L (1995) Local indicators of spatial association –

LISA. Geographical Analysis 27: 93–115.

Arrell K, Wise S, Wood J, and Donoghue D (2008)

Spectral filtering as a method of visualizing and

removing striped artifacts in digital elevation data.

Earth Surface Processes and Landforms 33: 943–961.

Aubry P and Piegay H (2001) Pratique de l’analyse de

l’autocorrelation spatiale en geomorphologie:

Definitions operatoires et test. Geographie Physique

et Quaternaire 55: 111�126.

Bater CW and Coops NC (2009) Evaluating error associ-

ated with lidar-derived DEM interpolation. Computers

and Geosciences 35: 289–300.

Binh TQ and Thuy NT (2008) Assessment of the influence

of interpolation techniques on the accuracy of digital

elevation model. VNU Journal of Science, Earth

Sciences 24: 176–183.

Bolstad P and Stowe T (1994) An evaluation of DEM accu-

racy: Elevation, slope and aspect. Photogrammetric

Engineering and Remote Sensing 60: 1327–1332.

Bolstad P and Stowe T (1994) An evaluation of DEM accu-

racy: Elevation, slope and aspect. Photogrammetric

Engineering and Remote Sensing 60: 1327–1332.

Brown D and Bara T (1994) Recognition and reduction of

systematic error in elevation and derivative surfaces

from 7.5-minute DEMs. Photogrammetric Engineering

and Remote Sensing 60: 189�194.

Burrough PA (1986) Principles of Geographical Informa-

tion Systems for Land Resources Assessment. New

York: Oxford University Press.

Canters F, Genst WD, and Dufourmont H (2002) Asses-

sing effects of input uncertainty in structural landscape

classification. International Journal of Geographical

Information Science 16: 129�149.

Carlisle BH (2005) Modelling the spatial distribution of

DEM Error. Transactions in GIS 9: 521–540.

Carrara A, Bitelli G, and Carla R (1997) Comparison of

techniques for generating digital terrain models from

contour lines. International Journal of Geographical

Information Science 11: 451–472.

22 Progress in Physical Geography

Chaplot V, Darboux F, Bourennane H, Leguedois S,

Silvera N, and Phachomphon K (2006) Accuracy of

interpolation techniques for the derivation of digital

elevation models in relation to landform types and data

density. Geomorphology 77: 126–141.

Claessens L, Heuvelink GBM, Schoorl JM, and Veldkamp

A (2005) DEM resolution effects on shallow landslide

hazard and soil redistribution modeling. Earth Surface

Processes and Landforms 30: 461–477.

Codilean AT, Bishop P, and Hoey TB (2006) Surface

process models and the links between tectonics and topo-

graphy. Progress in Physical Geography 30: 307–333.

Darnell AR, Tate NJ, and Brunsdon C (2008) Improving

user assessment of error implications in digital

elevation models. Computers, Environment and Urban

Systems 32: 268–277.

Desmet PJJ and Govers G (1995) GIS-based simulation of

erosion and deposition patterns in an agricultural

landscape: A comparison of model results with soil

map information. Catena 25: 389–401.

Dietrich WE, Bellugi DG, Sklar LS, Stock JD, Heimsath

AM, and Roering JJ (2003) Geomorphic transport laws

for predicting landscape form and dynamics. In:

Wilcock PR and Iverson RI (eds) Prediction in

Geomorphology, AGU Geophysical Monograph Series

135. Washington, DC: American Geophysical Union,

103–132.

Dietrich WE, Wilson CJ, Montgomery DR, and McKean J

(1993) Analysis of erosion thresholds, channel

networks, and landscape morphology using a digital

terrain model. Journal of Geology 101: 259–278.

Eblen JS (1995) A probabilistic investigation of slope

stability in the Wasatch Range, Davis County, Utah.

Utah Geological Survey Contract Report 95-3. Utah:

Department of Geological Survey – Department of

Natural Resources.

Ehlschlaeger CR (1998) The stochastic simulation

approach: Tools for representing spatial application

uncertainty. Unpublished doctoral thesis, University

of California, Santa Barbara.

Ehlschlaeger CR and Shortridge A (1996) Modeling eleva-

tion uncertainty in geographical analyses. Proceeding

of the International Symposium on Spatial Data Han-

dling, Delft, Netherlands, 9B.15–9B.25. Available at:

http://chuck.ehlschlaeger.info/older/SDH96/

paper.html.

Elobaid RF, Shitan M, Ibrahim NA, Ghani ANA, and Daud

I (2009) Evolution of spatial correlation of mean

diameter: A case study of trees in natural dipterocarp

forest. Journal of Mathematics and Statistics 5: 267–269.

Erdogan S (2010) Modelling the spatial distribution of

DEM error with geographically weighted regression:

An experimental study. Computers and Geosciences

36: 34–43.

Felicısimo A (1994) Parametric statistical method for error

detection in digital elevation models. ISPRS Journal of

Photogrammetry and Remote Sensing 49: 29–33.

Fisher PF (1998) Improved modeling of elevation error

with geostatistics. GeoInformatica 2: 215–233.

Fisher PF and Tate NJ (2006) Causes and consequences of

error in digital elevation models. Progress in Physical

Geography 30: 467–489.

Florinsky IV (1998) Accuracy of local topographic vari-

ables derived from digital elevation models. Interna-

tional Journal of Geographical Information Science

12: 47–62.

Fotheringham AS, Brunsdon C, and Charlton M (2002)

Geographically Weighted Regression, the Analysis of

Spatially Varying Relationships. Chichester: Wiley.

Gatziolis D and Fried JS (2004) Adding gaussian noise to

inaccurate digital elevation models improves fidelity of

derived drainage networks. Water Resources Research

40: W02508.

Getis A (1991) Spatial interaction and spatial autocorrela-

tion: A cross-product approach. Environment and

Planning 23: 1269�1277.

Giles PT and Franklin SE (1996) Comparison of derivative

topographic surfaces of a DEM generated form

stereoscopic SPOT images with field measurements.

Photogrammetric Engineering and Remote Sensing

62: 1165–1171.

Goodchild MF (1982) The fractal Brownian process as a

terrain simulation model. Modelling and Simulation

13: 1133–1137.

Goovaerts P, Jacquez MG, and Marcus A (2005) Geosta-

tistical and local cluster analysis of high resolution

hyperspectral imagery for detection of anomalies.

Remote Sensing of Environment 95: 351–367.

Gousie MB (2005) Digital elevation model error detection

and visualization. In: Gold C (ed.) The 4th Workshop on

Dynamic and Multi-dimensional GIS, Pontypridd,

Wales, UK. International Society for Photogrammetry

and Remote Sensing (ISPRS), 42–46.

Grohmann CH, Smith MJ, and Riccomini C (2009)

Surface roughness of topography: A multiscale analy-

sis of landform elements in Midland Valley, Scotland.

Gonga-Saholiariliva et al. 23

Proceeding of Geomorphometry 2009, 31 August–2

September. Zurich: Zurich University, 140–148.

Hannah MJ (1981) Error detection and correction in digital

terrain models. Photogrammetric Engineering and

Remote Sensing 47: 63–69.

Hebeler F and Purves RS (2009) The influence of elevation

uncertainty on derivation of topographic indices.

Geomorphology 111: 4–16.

Hengl T, Heuvelink GBM, and Van Loon EE (2010) On

the uncertainty of stream network derived from eleva-

tion data: The error propagation approach. Hydrology

and Earth Sciences Discussions 7: 767–799.

Heuvelink GBM, Burrough PA, and Leenaers H (1989)

Propagation of errors in spatial modelling with GIS.

International Journal of Geographical Information

Systems 3: 303–322.

Hirano A, Welch R, and Lang H (2003) Mapping from

ASTER stereo image data: DEM validation and accu-

racy assessment. ISPRS Journal of Photogrammetry

and Remote Sensing 57: 356–370.

Holmes KW, Chadwick OA, and Kyriakidis PC (2000)

Error in a USGS 30-meter elevation model and its

impact on terrain modeling. Journal of Hydrology

333: 154–173.

Hunter G and Goodchild M (1995) Dealing with error in

spatial databases: A simple case study. Photogram-

metric Engineering and Remote Sensing 61: 529–537.

Hunter G and Goodchild M (1997) Modeling the

uncertainty of slope and aspect estimates derived from

spatial databases. Geographical Analysis 29: 35�49.

Jacquez GM (2008) Spatial cluster analysis. In: Fothering-

ham S and Wilson J (eds) The Handbook of Geographic

Information Science. Oxford: Blackwell, 395–416.

Jones KH (1998) A comparison of algorithms used to

compute hill slope as a property of the DEM. Computer

and Geosciences 24: 315–323.

Kamp U, Bolch T, and Olsenholler J (2003) DEM Gener-

ation from ASTER satellite data for geomorphometric

analysis of Cerro Sillajhuay, Chile/Bolivia.

Proceedings Annual Meeting Imaging and Geospatial

Information Society (ASPRS), 5–9 May, Anchorage,

USA.

Karras GE and Petsa E (1993) DEM matching and detec-

tion of deformation in close-range photogrammetry

without control. Photogrammetric Engineering and

Remote Sensing 59: 1419�1424.

Keys H (1981) Cubic convolution interpolation for digital

image processing. IEEE Transactions on Signal

Processing, Acoustics, Speech, and Signal Processing

29: 1153–1160.

Kienzle SW (2004) The effect of DEM raster resolution on

first order, second order and compound terrain deriva-

tives. Transactions in GIS 8: 83–111.

Kirkby MJ (1990) The landscape viewed through models.

Zeitschrift fur Geomorphologie NF, Supplement Band

79: 63–81.

Leick A (1990) GPS Satellite Surveying. New York:

Wiley.

Li J and Wong WS (2010) Effect of DEM sources on

hydrologic applications. Computers, Environment and

Urban Systems 34: 251–261.

Lindsay JB (2006) Sensitivity of channel mapping

techniques to uncertainty in digital elevation data.

International Journal of Geographical Information

Science 20: 669–692.

Lindsay JB and Evans MG (2008) The influence of

elevation error on the morphometrics of channel

networks extracted from DEMs and the implication for

hydrological modeling. Hydrological Processes 22:

1588–1603.

Liu H and Jezek KC (1999) Investigating DEM error

patterns by directional variograms and Fourier analysis.

Geographical Analysis 3(3): 249–266.

Liu X (2008) Airborne LiDAR for DEM generation:

Some critical issues. Progress in Physical Geography

32: 31–49.

Liu X and Bian L (2008) Accuracy assessment of DEM

slope algorithms related to spatial autocorrelation of

DEM Errors. In: Zhou Q, Lees B, and Tang GA (eds)

Advances in Digital Terrain Analysis, Lecture Notes

in Geoinformation and Cartography. Berlin: Springer,

307–322.

Lopez C (1997) Locating some types of random errors.

International Journal of Geographical Information

Science 11: 677–698.

Lopez C (2000) Improving the elevation accuracy of digi-

tal elevation models: A comparison of some error

detection procedures. Transactions in GIS 4: 43–64.

McCalpin JP and Berry ME (1996) Soil catenas to estimate

ages of movements on normal fault scarps, with an

example from the Wasatch fault zone, Utah, USA.

Catena 27: 265�286.

Monckton C (1994) An investigation into the spatial struc-

ture of error in digital elevation data. In: Worboys MF

(ed.) Innovations in GIS 1. London: Taylor and Francis,

201–211.

24 Progress in Physical Geography

Moran PAP (1950) Notes on continuous stochastic phe-

nomena. Biometrika 37: 17�23.

Nelson AR and Personius SF (1993) Surficial geologic

map of the Weber segment, Wasatch fault zone, Weber

and Davis Counties, Utah. US Geological Survey Mis-

cellaneous Investigations. Series Map I-2199, pamph-

let, scale 1:50,000, 22 pp.

Oksanen J (2003) Tracing the gross errors of DEM – visualiza-

tion techniques for preliminary quality analysis. Proceed-

ings of the 21st International Cartographic Conference,

10–16 August, Durban, South Africa, 2410�2416.

Oksanen J and Sarjakoski T (2005) Error propagation

of DEM-based surface derivatives. Computers and

Geosciences 31: 1015–1027.

Oliveau S (2005) Les indicateurs locaux d’association

spatiale (LISA) comme methode de regionalisation.

Septiemes Rencontres de Theo Quant, January.

Available at: http://thema.univ-fcomte.fr/theoq/pdf/

2005/TQ2005%20ARTICLE%2029.pdf.

Petit C, Gunnell Y, Gonga-Saholiariliva N, Meyer B, and

Seguinot J (2009) Faceted spurs at normal fault scarps:

Insights from numerical modeling. Journal of Geophy-

sical Research 114: B05403.

Podobnikar T (2009) Methods for visual quality

assessment of a digital terrain model. S.A.P.I.EN.S 2.

Available at: http://sapiens.revues.org/738.

Polidori L (1991) Validation de modeles numeriques de

terrain, application a la cartographie des risques

geologiques. Unpublished doctoral thesis, Universite

Paris Diderot.

Raaflaub LD and Collins MJ (2006) The effect of error in

gridded digital elevation models on the estimation of

topographic parameters. Environmental Modelling and

Software 21: 710�732.

Reuter HI, Hengl T, Gessler P, and Soille P (2009)

Preparation of DEMs for geomorphometric analysis.

In: Hengl T and Reuter HI (eds) Geomorphometry –

Concepts, Software, Applications, Developments in

Soil Science 33. Amsterdam: Elsevier, 87–119.

Schmidt J, Evans IS, and Brinkmann J (2003) Comparison

of polynomial models for land surface curvature

calculation. International Journal of Geographical

Information Science 17: 797–814.

Smith MJ and Clark CD (2005) Methods for the visualiza-

tion of digital elevation models for landform mapping.

Earth Surface Processes and Landforms 30: 885–900.

Stock GM, Frankel KL, Ehlers TA, Schaller M, Briggs

SM, and Finkel RC (2009) Spatial and temporal

variations in denudation of the Wasatch Mountains,

Utah, USA. Lithosphere 1: 34–40.

Tarboton DG (1997) A new method for the determination

of flow directions and upslope areas in grid digital

elevation models. Water Resources Research 33:

309–319.

Tempfli K (1980) Spectral analysis of terrain relief for

the accuracy estimation of digital terrain models. ITC

Journal 3: 478–510.

Thapa K and Bossler J (1992) Accuracy of spatial data

used in geographic information systems. Photogram-

metric Engineering and Remote Sensing 58: 835–841.

Van Niel KP, Laffan SW, and Lees BG (2004) Effect of

error in the DEM on environmental variables for

predictive vegetation modeling. Journal of Vegetation

Science 15: 747�856.

Walsh SJ, Lightfoot DR, and Butler DR (1987) Recogni-

tion and assessment of error in geographic information

systems. Photogrammetric Engineering and Remote

Sensing 53: 1423–1430.

Warren SD, Hohmann MG, Auerswald K, and Mitasova H

(2004) An evaluation of methods to determine slope

using digital elevation data. Catena 58: 215–233.

Wechsler SP (1999) Digital elevation model (DEM)

uncertainty: Evaluation and effect on topographic

parameters. Proceedings 19th Annual ESRI User Con-

ference, San Diego, California, USA. Available at:

http://gis.esri.com/library/userconf/proc99/proceed/

papers /pap262/p262.htm.

Wechsler SP (2003) Perceptions of digital elevation model

uncertainty by DEM users. Urban and Regional Infor-

mation Systems Association Journal 15: 57–64.

Wechsler SP (2006) Uncertainties associated with digital

elevation models for hydrologic applications: A

review. Hydrology and Earth System Sciences Discus-

sions 3: 2343–2384.

Wechsler SP and Kroll CN (2006) Quantifying DEM

uncertainty and its effect on topographic parameters.

Photogrammetric Engineering and Remote Sensing

72: 1081–1090.

Wilson JP and Gallant JC (2000) Digital terrain analysis.

In: Wilson JP and Gallant JC (eds) Terrain Analysis:

Principles and Applications. New York: Wiley, 1–28.

Wise SM (1998) The effect of GIS interpolation errors on

the use of digital elevation models in geomorphology.

In: Lane SN, Richards KS, and Chandler JH (eds)

Landform Monitoring, Modelling and Analysis.

Chichester: Wiley, 139–164.

Gonga-Saholiariliva et al. 25

Wise SM (2000) Assessing the quality for hydrological

applications of digital elevation models derived from

contours. Hydrological Processes 14: 1909�1929.

Wise SM (2007) Effect of differing DEM creation methods

on the results from a hydrological model. Computers

and Geosciences 33: 1351–1365.

Wood JD (1994) Visualising contour interpolation accu-

racy in digital elevation models. In: Hearnshaw HM

and Unwin DJ (eds) Visualization in Geographical

Information Systems. Chichester: Wiley, 168–180.

Wood JD (1996) The geomorphological characterisation

of digital elevation models. Unpublished doctoral the-

sis, University of Leicester.

Younan NH, Lee HS, and King RL (2002) DTM error mini-

mization via adaptive smoothing (LiDAR forest measure-

ments). Geoscience and Remote Sensing Symposium.

IGARSS’02.2002 IEEE International 6: 3611–3613.

Yue TX, Du ZP, Song DJ, and Gong Y (2007) A new

method of surface modeling and its application to DEM

construction. Geomorphology 91: 161–172.

Zevenbergen LW and Thorne CR (1987) Quantitative

analysis of land surface topography. Earth Surface and

Landforms 12: 47�56.

Zhang T and Cen M (2008) Robust DEM co-registration

method for terrain changes assessment using least

trimmed squares estimator. Advances in Space

Research 41: 1827–1835.

Zhou Q and Liu X (2004a) Error analysis on grid-based

slope and aspect algorithms. Photogrammetric

Engineering and Remote Sensing 70: 957–962.

Zhou Q and Liu X (2004b) Analysis of errors of derived

slope and aspect related to DEM data properties.

Computers and Geosciences 30: 369–378.

Ziadat FM (2007) Effect of contour intervals and grid cell

size on the accuracy of DEMs and slope derivatives.

Transactions in GIS 11: 67–81.

Zuchiewicz WA and McCalpin JP (2000) Geometry of

faceted spurs on an active normal fault: Case study

of the central Wasatch Fault, Utah, USA. Annales

Societatis Geologorum Poloniae 70: 231�249.

26 Progress in Physical Geography