Vous êtes sur la page 1sur 13

Journal of Hydrology 371 (2009) 169181

Contents lists available at ScienceDirect

Journal of Hydrology
journal homepage: www.elsevier.com/locate/jhydrol

Effect of spatial trends on interpolation of river bathymetry


Venkatesh Merwade *
School of Civil Engineering, 550 Stadium Mall Drive, Purdue University, West Lafayette, IN 47907, United States

a r t i c l e

i n f o

Article history:
Received 14 October 2008
Received in revised form 18 March 2009
Accepted 25 March 2009

This manuscript was handled by G. Syme,


Editor-in-Chief, with the assistance of Paul
Jeffrey, Associate Editor
Keywords:
River channels
Bathymetry
Spatial trend
Interpolation
River modeling

s u m m a r y
Continuous surface of river bathymetry (bed topography) is typically produced by spatial interpolation of
discrete point or cross-section data. Several interpolation methods that do not account for spatial trend in
river bathymetry produce inaccurate surfaces, thus requiring complex interpolation methods such as
anisotropic kriging. Although isotropic methods are unsuitable for interpolating river bathymetry, issues
that limit their use remain unaddressed. This paper addresses the issue of effect of spatial trend in river
bathymetry on isotropic interpolation methods. It is hypothesized that if the trend is removed from the
data before interpolation, the results from isotropic methods should be comparable with anisotropic
methods. Data from six river reaches in the United States are used to: (i) interpolate bathymetry data
using seven spatial interpolation methods; (ii) separate trend from bathymetry; (iii) interpolate residuals
(bathymetry minus trend) by using all seven interpolation methods to get residual surfaces, (iv) add the
trend back to residual surfaces; and (v) compare resulting surfaces from (iv) with surfaces created in (i).
Quantitative and qualitative comparison of results through root mean square error (RMSE), semi-variograms, and cross-section proles show that signicant improvement (as much as 60% in RMSE) can be
accomplished in spatial interpolation of river bathymetry by separating trend from the data. Although
this paper provides a new simple way for interpolating river bathymetry by using (otherwise deemed
inappropriate) isotropic methods, the choice of trend function and spatial arrangement of discrete
bathymetry data play a vital role in successful implementation of the proposed approach.
2009 Elsevier B.V. All rights reserved.

Introduction
River bathymetry (bed topography) plays a critical role in
numerical modeling of ow hydrodynamics, sediment transport,
ecological and geomorphologic assessments. Conventional way of
measuring river bathymetry is through cross-sectional surveys
where ground proles are collected at certain locations along the
river depending on available resources, river morphology and end
use of the data. A recent technological development in bathymetry
measurement includes the use of boat-mounted SONAR (Sound
Navigation And Ranging) devices such as single or multi-beam
echosounder in conjunction with global positioning system (GPS)
to give a series of (x, y, z) bathymetry points (Vermeyen, 2006;
Rogala, 1999). Although the spatial resolution of bathymetry points
collected through echosounding techniques can be much better
compared to cross-sections, these data still represent a discrete
sample of a continuous bathymetric surface. Continuous mapping
of shallow river bathymetry over large areas through air-borne
techniques is also an active area of research these days (Hilldale
and Raff, 2008; Legleiter et al., 2004; Marcus et al., 2003), but for
deeper rivers discrete eld data are still the best way for creating
* Tel.: +1 765 494 2176; fax: +1 765 494 0395.
E-mail address: vmerwade@purdue.edu
0022-1694/$ - see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.jhydrol.2009.03.026

accurate bathymetric surfaces (e.g., Lampe and Morlock, 2007;


White and Hodges, 2005).
Traditional approach to studying ow, sediment, and aquatic
habitat in river channels is through one-dimensional models that
take bathymetric information in the form of cross-sections (e.g.,
Gard, 2005; Lee et al., 2006; Martin, 2003; Torizzo and Pitlick,
2004; Yang et al., 2006). Although the normal ow in rivers can
be assumed to be one-dimensional in the main channel, this
assumption becomes invalid during high oods, thus necessitating
the use of 2D/3D hydrodynamic models (e.g., Carrivick, 2006;
Dutta et al., 2007; Taye et al., 2007). Similarly sh habitat are
increasingly being related to three-dimensional nature of river
hydraulics, thus requiring 2D/3D models for ecological assessment
(e.g., Booker, 2003; de Jalon and Gortazar, 2007; Mouton et al.,
2007; Shen and Diplas, 2008). Multi-dimensional models are also
becoming popular in sediment transport and geomorphology
(e.g., Khosronejad et al., 2007; Mekonnen and Dargahi, 2007; Yue
et al., 2008). Although much progress has been made in representation and simulation of river processes in 2D/3D, successful application of these models is directly linked to accurate bathymetric
representation (Buttner, 2007; Crowder and Diplas, 2000; Horritt
et al., 2006; Lane et al., 2002). Bathymetry data are incorporated
into 2D/3D models by interpolating observed discrete data (points
or cross-sections) to get elevations at model nodes (e.g., nodes of a

170

V. Merwade / Journal of Hydrology 371 (2009) 169181

nite element mesh). Therefore, the accuracy of bathymetric surfaces represented in 2D/3D models is dependent upon the ability
of interpolation methods in making accurate predictions at unmeasured locations using discrete data. Recent studies have shown that
commonly available interpolation methods such as triangulation,
inverse distance weighting (IDW), splines or kriging, which assume
isotropy in data, yield inaccurate river bathymetric surface (Goff
and Nordfjord, 2004; Merwade et al., 2006). The isotropic assumption in most spatial interpolation methods ignores the trend in river bathymetry that is linear (bed slope) in ow direction and
nonlinear (cross-sectional shape) across ow direction. As a result,
methods that account for river ow direction and topographic
trend are recommended for interpolating discrete bathymetry
data.
Additional constraints imposed by river ow direction and
topographic trend limit the choice of methods available for interpolating river bathymetry to only a few specialized ones such as
anisotropic kriging, or custom modications of existing isotropic
methods (e.g., elliptical IDW by Merwade et al., 2006; anisotropic
IDW by Tomczak, 1998). Although anisotropic methods provide
better results compared to isotropic methods, ways to improve
the results from application of simple isotropic methods for interpolating river bathymetry remain unexplored. Specically the effect of separating existing trends in river bathymetry before
interpolating discrete points is not studied. Such a study can: (a)
potentially overcome the limitations of isotropic methods, thus
making them widely applicable for interpolating river bathymetry;
(b) demonstrate the effect of trend on interpolating river bathymetry; and (c) provide information on accuracy that can be gained by

using complex interpolation methods, such as anisotropic kriging,


compared to isotropic methods if the trend is excluded from river
bathymetry. This paper addresses the issue of effect of bathymetry
trend by presenting results from a study that involved interpolation of discrete bathymetry point dataset with and without removal of trend at six river reaches in the United States.
Although the topic of spatial trends and anisotropy has received
considerable attention in the eld of soil science and groundwater
hydrology (e.g., Petersen et al., 2008; Crawford and Hergert, 1997;
Pfannkuch and Winter, 1984), the present study explores this topic
for river channels, thus providing a new way of interpolating river
bathymetry using simple isotropic methods in the absence of specialized complex techniques.
Study areas and datasets
Bathymetry data collected at six river reaches in the United
States are used in this study (Fig. 1a). Of these six reaches, four
(Brazos, King Ranch, Rainwater and Sulphur) are located in Texas,
and two (Kentucky and Ohio) are located in Kentucky. Except for
King Ranch and Rainwater which are located along the Guadalupe
River in Texas, names of all other reaches reect the names of corresponding rivers. Data for Texas Rivers were provided by Texas
Water Development Board (TWDB); whereas data for Kentucky
and Ohio reaches were provided by North Carolina Water Science
Center (NCWSC). Data for Texas Rivers were collected for Instream
Flow Program (TWDB, 2008), and data for Ohio and Kentucky were
collected for water quality and sediment transport studies
(Wagner and Mueller, 2001, 2002).

Fig. 1. Bathymetry dataset for: (a) Brazos (B) River; (b) Ohio (O) River; (c) King Ranch (KR) reach of Guadalupe River; (d) Rainwater (R) reach of Guadalupe River; (e) Sulphur
(S) River; and (f) Kentucky River (K).

171

V. Merwade / Journal of Hydrology 371 (2009) 169181


Table 1
Summary of data at study reaches (+ obtained by averaging the widths at six equidistant locations along the reach).
River name

Brazos
King Ranch
Rainwater
Sulphur
Kentucky
Ohio

River characteristics

Bathymetry data

Length
(km)

Mean width
(m)+

Mean depth (m)

Slope

Substrate type

Total points

Points/50 
50 m2 area

Mean
(m)

Std. Dev.
(m)

Min.
(m)

Max.
(m)

7.1
1.6
3.2
1.4
7.2
30

105
42
44
33
106
465

3.1
1.4
1.74
4.22
7.98
4.48

0.03
0.01
0.08
0.04
0.02
0.001

Sandy
Gravel/bedrock
Gravel/bedrock
Silty/Sandy clay
Gravel/bedrock
Cohesive/noncohesive sediment

37288
7602
14955
7732
66682
20554

116
274
243
407
224
4

10.42
143.46
144.8
72.06
142.8
111.97

1.43
0.64
1.16
1.48
2.46
1.81

0.11
141.66
140.85
68.9
138.11
106.99

13.25
144.96
146.48
76.3
150.8
116.45

Data at each reach include bathymetry (x ,y, z) points collected


using a single-beam echosounder and global positioning system
(GPS) mounted on a boat. As the boat moves, the single-beam
echosounder measures the water depth (d) by pinging sound signals and the GPS records (x, y) location of each ping. Typically,
depth sounders ping multiple signals at one point, and then average the data to give a single recording at that point. Additionally,
all GPS recordings are corrected through differential GPS (DGPS)
to give more accurate (x, y) locations. Water depth at each point
is then subtracted from water surface elevation to get z, thus creating a series of (x, y, z) bathymetry points. The number of measurements and their spatial arrangement depends on the speed and
path taken by the boat as shown in Fig. 1. Data for Sulphur and
Ohio reach were modied by TWDB and NSWSC, respectively to
match nite element nodes for a hydrodynamic model, and do
not reect the actual arrangement of eld data. It is possible that
some information was lost during this data transformation for
Ohio and Sulphur datasets, but for the purpose of this study, these
modied data are taken as actual measurements. Data at these six
study sites provide a good representation of different spatial
arrangement ranging from traditional cross-sections (Kentucky)
to irregular spacing (Brazos, Rainwater, King Ranch) to regularly
spaced gridded data (Sulphur and Ohio). Morphologic details at
each reach (length, width, slope) including information on
bathymetry data are provided in Table 1.
Methodology
The methodology involves the following steps: (i) interpolation
of bathymetry data including trend; (ii) separating trend from
bathymetry points by using three trend functions; (iii) interpolation of residuals (bathymetry minus trend) by using seven methods to get residual surfaces; (iv) adding the trend back to
residual surfaces; and (v) comparison of resulting surfaces from
(iv) with surfaces created in (i). Key tasks in the methodology are
described below. All geospatial tasks are performed using spatial
and geostatistical analyst extensions in ArcGIS from Environmental
Systems Research Institute (ESRI).
Testing and validation datasets
Each initial bathymetry dataset is split into two sub-sets: testing and validation. Bathymetry points in testing datasets are used
to create interpolated surfaces, which are then compared against
measured bathymetry in validation datasets. Different approaches
are used in creating testing and validation datasets for each study
area depending on the conguration of the available data. In the
case of Brazos River, testing and validation datasets are used from
a previous study by Merwade et al. (2006), which involved manual
separation of bathymetry points to mimic a cross-sectional conguration in the testing dataset. In the case of King Ranch, Rainwater,
Sulphur, and Ohio River, validation points are extracted through
random selection. In the case of Kentucky River, which had points

along cross-sections, alternate cross-sections are selected for validation. Overall, all datasets except Kentucky River are split such
that 70% of bathymetry points are included in testing dataset and
30% in validation dataset. For Kentucky River, the testing and validation datasets are split to have 50% points each (alternate crosssections).
Mapping of data in (s, n) coordinate system
River bathymetry data are collected and stored using Cartesian
(x, y) coordinates, but use of these data in this coordinate system
can introduce issues related to meandering nature of the channel.
For example, computing distance between two points along a river
using (x, y) coordinates will not give the actual ow length between
these points for a meandering river. To overcome such issues,
channel-tted coordinate system dened by an s-axis along the
ow direction and n-axis across the ow direction (perpendicular
to s-axis) is widely used for rivers (e.g., Johannesson and Parker,
1989; Nelson and Dungan, 1989; Ye and McCorquodale, 1997).
The s-axis can be aligned with either river banks or centerline. In
this study, the geometric centerline of the channel is treated as
the s-axis with upstream end of the river as its origin (s = 0). Similarly, looking downstream, all n coordinates are negative on the
left hand side of s-axis, and vice versa. Plotting of river in (s, n)
coordinates straightens the river so data can be treated with respect to the ow direction as shown in Fig. 2. A GIS procedure
developed by (Merwade et al., 2005) is used for mapping bathymetry in (s, n) coordinates for all six reaches.
Fitting trend to bathymetry points
Mapping of bathymetry in (s, n) coordinates makes the linear
trend (bed slope) a function of s coordinate, and nonlinear trend
(cross-sectional shape) a function of n coordinate. Besides using
the n coordinate, the nonlinear trend is modeled by using other
physical attributes such as channel width, depth and meandering
curvature (see e.g., Deutsch and Wang, 1996; James, 1996;
Legleiter and Kyriakidis, 2008). It is assumed that the quality of
trend surface, and eventually variance in residuals (measurements
trend) will depend on the selected trend function. In addition,
whether a trend function is applied locally (separate function
formulations for individual local areas) or globally (one function
for the entire reach) can also produce different results. Several
techniques such as power functions, polynomials, splines and
probability density function (pdf) can be used to t a trend to
rivers cross-sectional shape. A detailed review of these techniques
including their application in tting river cross-sections can be
found in Merwade and Maidment (2004).
To assess the effect of trend function in spatial interpolation,
three techniques (two local and one global) are employed in this
study. Local techniques include cubic polynomial and a combination of two beta pdf (Eq. (1)), and the global technique includes a
cubic polynomial function.

172

V. Merwade / Journal of Hydrology 371 (2009) 169181

n
s
-n

(a)

(b)

Fig. 2. (s, n) Coordinate transformation for King Ranch dataset. Bathymetry points in: (a) (x, y) coordinates and (b) (s, n) coordinates.


Tz ff xja1 ; b1 f xja2 ; b2 g k w;

where, T(z) is the trend function, w is the channel width, k (0 < k < 1)
is a scaling factor, and a1, a2, b1, and b2 are the parameters of beta
pdf given by Eq. (2)

f xja; b

1
xa1 1  xb1 ; 0 < x < 1a > 0;
Ba; b

b>0

Cubic polynomial is selected for its simplicity and easy availability in many software programs for tting spatial trends. A combination of beta pdf is used because comparison of several
techniques listed earlier for tting cross-sectional trend has shown
that a combination of two beta pdf produce best results (Merwade
and Maidment, 2004). The objective is not to advocate any one
technique over others for tting trends, but to see how different
techniques affect spatial interpolation results of bathymetry
points. For local polynomial trend tting, a circular search neighborhood is dened to include 50 points, and for beta pdf, the neighborhood is dened to include points covering a channel length
equal to the average channel width. The neighborhoods for polynomial and beta pdf are dened (by trial and error) such that the
points included in a neighborhood should be able to dene a complete river cross-section. All parameters in cubic polynomials and
beta functions are optimized to minimize the root mean square error (RMSE, Eq. (5)) between predictions and observations.

Spatial interpolation
Spatial interpolation is preformed for each training dataset in
(s, n) coordinates for two variables: zi and ei where, zi is measured
bathymetry at any point i, and ei ( = zi  T) is corresponding residual after tting the trend function. Seven spatial interpolation
techniques are used in this study. These are IDW, regularized
spline (RS), spline with tension (TS), topogrid (TG), natural neighbor (NN), ordinary kriging (OK) and ordinary kriging with anisotropy (AK). A review of these techniques for interpolating
watershed topography and river bathymetry can be found in Chaplot et al. (2006) and Merwade et al. (2006), respectively, but a
brief description and corresponding equations of each technique
are presented in Table 2. All these techniques are commonly used
in many disciplines including hydrology through several commercial and public domain software programs.
Assessment of trend surfaces
Trend surfaces are assessed for similarity and their ability to
best describe the trend in the data. Such an assessment will help
to understand how much effect a trend surface can have on nal
interpolated bathymetry surface (trend + interpolated residual).
Assessment of trend surfaces is performed by using percentage of

Table 2
Spatial interpolation methods.
Method

Description

Equations

IDW

Value at an unsampled location (z*) is distance-weighted average of nearby


observations (zi). ki = weight at ith point; di = distance between zi and z*; N = total no.
of points

i
z RN
i1 ki zi ; ki RN

Tension spline

1
d2

1
i1 d2
i

RR

y
2
2
fx2 fy2 fxx
fxy
fyy
dxdy;

Regularized spline

A minimum curvature (smoothest) function [I(f)] is dened that passes through a set
of observations. R = 2D euclidian space; T(x, y) = local trend; R(r,rj) = basis function;
/ = tension weight; K0 = modied Bessel function of zero order; c = 0.577215
Third and higher order derivatives are added to I(f)

Topogrid

A variation of thin plate splines developed by Hutchinson (1989, 1993)

Natural neighbor

Similar to IDW, but weights are computed based on area. pi and qi are the Thiessen
areas surrounding zi excluding and including z*, respectively

ki pi  qj =pi

Ordinary kriging

Similar to IDW, but weights are assigned based on distance and spatial correlation.
The objective is to minimize r2 to get unbiased estimate with minimum variance
Ordinary kriging that takes data anisotropy into account

r2 E z  ^z 2 E z  Rni1 ki zi 2  subject to

Anisotropic kriging

If

R2 /

f x; y RN
i1 t i Rr; r i Tx; y;
h  
i
d/
Rr; r j 2p1/2 ln 2j c K 0 dj /

RNi1 ki zi

173

V. Merwade / Journal of Hydrology 371 (2009) 169181

total sum of squares (SST, Eq. (3)), and by conducting t-test


(a = 0.05). Although these techniques are commonly used for comparing the accuracy of surfaces, they can produce spurious results
for spatially correlated bathymetry data. Therefore, quantitative
assessments are augmented with visual inspection of surfaces,
and comparison of experimental semi-variogram plots.

Pn

zi  ^zi
SST 1  Pi1
;
ni1 zi  z2

where z is the mean. A semi-variogram plot is a plot of semivariance (Eq. (4)) as a function of distance between observation
points as shown in Fig. 5. In spatially correlated data such as river
bathymetry, semi-variance is smaller for nearer points, and vice
versa. Therefore, semi-variance increases with distance between
measurement points, until a threshold is reached in the distance
of separation. This threshold is called the range as shown in
Fig. 5b. The semi-variance corresponding to the range is called the
sill, and semi-variance corresponding to zero separation distance
on a semi-variogram plot is called the nugget. Semi-variogram can
be used to compare the spatial distribution of semi-variance captured by different trend and interpolated surfaces. Moreover, tting
of semi-variogram in different directions will show the extent to
which the trend in the data is captured by a trend-tting function.
For example, if the trend in the bathymetry is captured, the semivariograms of residuals in all direction should be identical.

cij h

Ezi  zj 2
;
2

where cij is the semi-variance between two bathymetry points


(zi, zj) separated by distance h.
Cross-validation of interpolated surfaces
After spatial interpolation of points (zi) in testing dataset, the
resulting surfaces are validated against observed bathymetry in
validation datasets. Several methods are available to assess the
quality of interpolated surfaces compared to measured bathymetry
in validation datasets. Commonly used quantitative methods include precision indices such as mean error, mean absolute error
or the root mean square error (RMSE). In this study, RMSE (Eq.
(5)) which provides an overall measure of how close or accurate
the estimated elevations are compared to measurements is used
for quantitative assessment.

"
RMSE

n
1X
zi  ^zi 2
n i1

#1=2
;

where zi is measurement at ith location, ^zi is estimate of zi, and n is


the total number of points in the validation dataset. When ei is
interpolated, ^zi is obtained by adding T(i) to ^ei .
Besides RMSE, surfaces are also compared by using SST, and ttest (a = 0.05). Similar to trend surfaces, quantitative assessments
of interpolated surfaces are augmented with visual inspection,
and comparison of experimental semi-variogram plots.

Results
Assessment of trend surfaces
To keep the terminology simple, beta trend surface will be referred as BT, local polynomial trend as LT and global polynomial
trend as GT. Application of traditional techniques such as Students
t-test and SST for comparing trend surfaces show some interesting
results (Tables 3 and 4). For example, the mean of BT, LT and GT are
not signicantly different for Kentucky, the mean for BT and LT are

Table 3
Students t-test results (a = 0.05) between beta (BT), local polynomial (LT) and global
polynomial (GT) trend surfaces for all six reaches (B = Brazos; KR = King Ranch;
RW = Rainwater; S = Sulphur; K = Kentucky; and O = Ohio).
BT
BT
LT
GT

K, KR
R, S, K

LT

GT

K, KR

R, S, K
B, KR, R, K, O

B, KR, R, K, O

Table 4
SSE results for beta (BT), local polynomial (LT) and global polynomial (GT) trend
surfaces for all six reaches.

Brazos
King Ranch
Rainwater
Sulphur
Kentucky
Ohio

BT

LT

GT

82.05
79.61
90.23
80.22
94.48
79.52

83.85
75.77
84.82
82.95
87.46
81.59

44.89
19.41
60.08
50.15
56.75
51.62

signicantly different for Brazos, Ohio and Sulphur datasets


(Table 3). Overall, the mean is not signicantly different for all
three trend surfaces for many study areas (except for Rainwater)
at 95% condence interval. The results of SST (Table 4) show that
the BT and LT capture more than 80% of variance in the measured
bathymetry for all datasets; whereas GT captures only around 50%
of variance in measured bathymetry for most datasets. Lower SST
for global polynomial trend for all datasets can be attributed to
its inability to capture local variations that exist in river bathymetry. Overall, Table 4 seems to show that beta and local polynomial
trends show similar performance in capturing the variations in
measured bathymetry. Does this mean that these two techniques
will have the same effect on bathymetry interpolation? This will
become clear when results from spatial interpolations are presented in a subsequent sub-section.
A visual comparison of trend surfaces with measured bathymetry (Fig. 3) show that the BT surface is better in capturing local
variations compared to LT. Only results for King Ranch dataset
are presented in Fig. 3 because the overall interpretation of this gure is applicable to all other datasets. Fig. 3a represents the base
surface that is created by interpolating all bathymetry points (testing + validation) using anisotropic kriging for comparison. The
smoothness of surface increases from BT to LT to GT, with GT unable to capture any local variations (Fig. 3bd). Plots of ground proles at two locations (Fig. 4) show how BT, LT and GT perform in
describing individual cross-sections. Cross-sections produced by
GT at both locations are almost identical; whereas BT and LT adjust
to t the trend at individual locations. It is also important to note
that a better trend surface that captures local variations requires
more parameters and computing power. A visual comparison of
overall surfaces and cross-section proles at individual locations
show that BT, LT and GT produce distinct surfaces that can have
different effects on interpolation of residuals, and subsequently
bathymetry surfaces.
Experimental semi-variogram plots of BT, LT, GT, and their corresponding residuals show how the spatial correlation in the
bathymetry is affected (Fig. 5). Results from King Ranch experimental semi-variogram and trend surfaces are consistent with
earlier statistical analysis. Both BT and LT produce similar semivariograms suggesting that these two surfaces have similar spatial
distribution of semi-variance. On the other hand, GT is unable to
capture most of the variance in the bathymetry. Semi-variogram
plots of bathymetry and trend surfaces in s-direction have higher
range compared to other directions due to smaller variations
(or higher correlation) in bathymetry along the ow direction.

174

V. Merwade / Journal of Hydrology 371 (2009) 169181

Fig. 3. Trend surfaces for King Ranch Reach: (a) bathymetry surface for the entire reach; (b) close-up view of the surface in (a); (c) trend surface using beta function; (d) trend
surface using local polynomial; and (e) trend surface using global polynomial.

145.0

144.5
Z

BT

LT

GT

144.5

BT

LT

GT

Elevation (m)

144.0
144.0
143.5

143.5

143.0
143.0
142.5
142.5

142.0
0

10

20

30

40

50

10

20

30

40

50

Distance (m)

a)

b)

Fig. 4. Cross-sections for King Ranch Reach at: (a) x location in Fig. 3b; (b) y location in Fig. 3b (Z represents the base surface; BT = beta trend; LT = local polynomial; and
GT = global polynomial).

Similarly semi-variograms of residuals show that BT and LT capture about 75% of semi-variance in the bathymetry (residual sill
is approximately 0.1 compared to 0.4 for bathymetry); whereas
residuals from GT produce a semi-variogram that looks similar to
observed bathymetry in terms of range and sill, suggesting that
no or very little variance is captured. Similarity of residual semi-

variograms from BT and LT in omni-direction and s-direction suggest that most of the trend is captured by these functions. Although
overall results from King Ranch semi-variograms are applicable to
all datasets, semi-variograms from Brazos that had relatively better GT surfaces is also presented for comparison (Fig. 6). Unlike
other datasets that had similar semi-variograms for BT and LT,

175

V. Merwade / Journal of Hydrology 371 (2009) 169181

0.5

0.5
Z

BT

LT

GT
0.4

0.3

0.3

0.2

0.2

0.1

0.1

LT

GT

Range

Nugget
Sill

0.4

BT

0.0

0.0
0

50

100

150

200

50

100

a)

150

200

b)

0.5

0.5

BT

LT

BT

GT

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

0.0
0

50

100

150

200

LT

GT

0.0
0

50

100

c)

150

200

d)

Fig. 5. Semi-variograms of bathymetry, trend and residual surfaces for King Ranch: (a) bathymetry and trend surfaces in all directions; (b) bathymetry and trend surfaces in
s-direction; (c) residuals in all directions; and (d) residuals in s-direction (Z represents the base surface; BT = beta trend; LT = local polynomial; and GT = global polynomial).

2.5

2.5

2.0

2.0

1.5

1.5

1.0

1.0

0.5

0.5

BT

LT

Z
LT
300

GT

0.0

0.0

100

200

300

400

500

100

200

a)

400

BT
GT
500

b)

1.8

1.8

1.5

1.5

1.2

1.2

0.9

0.9

0.6

0.6

0.3

0.3

BT

LT

BT

GT

0.0
0

100

200

c)

300

400

500

LT

GT

300

400

0.0
0

100

200

500

d)

Fig. 6. Semi-variograms of bathymetry, trend and residual surfaces for Brazos River: (a) bathymetry and trend surfaces in all directions; (b) bathymetry and trend surfaces in
s-direction; (c) residuals in all directions; and (d) residuals in s-direction (Z represents the base surface; BT = beta trend; LT = local polynomial; and GT = global polynomial).

176

V. Merwade / Journal of Hydrology 371 (2009) 169181

semi-variograms for Brazos BT and LT show visible difference.


Again, the question is: are these differences any signicant to affect
spatial interpolation? This question is answered in the next subsection.
Assessment of spatial interpolation techniques
Bathymetry points in all testing datasets are interpolated using
elevations to get rst set of bathymetry surfaces (interpolation
including trend). Next, residuals are interpolated, and added to
corresponding trend surfaces to get second set of bathymetry surfaces (interpolation excluding trend). As a result four bathymetry

surfaces are created for each interpolation technique. These are:


(a) interpolation including trend; (b) BT + interpolation of corresponding residuals; (c) LT + interpolation of corresponding residuals; and (d) GT + interpolation of corresponding residuals. RMSE for
each surface is then computed by using elevations in validation
datasets (Fig. 7). Mean RMSE values for all techniques for each validation dataset are presented in Table 5. The differences between
measured and estimated bathymetry, and between RMSE from
each technique are found to be statistically signicant through Students t-test at 95% condence level.
All techniques for all datasets show improvement in RMSE
values when the trend is excluded from spatial interpolation. The

0.6

0.6

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

0.0

0.0
IDW

NN

RS

TS

TG

OK

AK

IDW

NN

Brazos

RS

TS

TG

OK

AK

King Ranch

0.7

0.8

0.6

0.7

0.5

0.6

0.5

0.4

0.4
0.3

0.3
0.2

0.2

0.1

0.1

0.0

0.0
IDW

NN

RS

TS

TG

OK

AK

IDW

NN

Rainwater

RS

TS

TG

OK

AK

TG

OK

AK

Sulphur
1.0

3.0
2.5

0.8

2.0

0.6
1.5

0.4
1.0

0.2

0.5

0.0

0.0
IDW

NN

RS

TS

TG

Kentucky

OK

AK

IDW

NN

RS

TS

Ohio

Interpolation including trend

LT + interpolated residuals

BT + interpolated residuals

GT + interpolated residuals

Fig. 7. RMSE (m) results for spatial interpolation techniques (BT = beta trend; LT = local polynomial; and GT = global polynomial).

V. Merwade / Journal of Hydrology 371 (2009) 169181


Table 5
Mean RMSE in meter for all reaches by creating surfaces through: (a) interpolation
including trend; (b) BT + interpolation of corresponding residuals; (c) LT + interpolation of corresponding residuals; and (d) GT + interpolation of corresponding residuals.
The numbers in parenthesis represent the percentage improvement in RMSE from (b),
(c) and (d) with respect to (a).

Brazos
King Ranch
Rainwater
Sulphur
Kentucky
Ohio

(a)

(b)

(c)

(d)

0.41
0.40
0.45
0.52
1.83
0.54

0.32 (22)
0.34(17)
0.41 (10)
0.49 (6)
0.73 (60)
0.39 (28)

0.36(14)
0.36(11)
0.41 (10)
0.40 (25)
1.18(35)
0.42 (22)

0.38 (9)
0.37 (9)
0.43 (6)
0.41 (22)
1.33 (27)
0.45(16)

Table 6
Mean percentage reduction in RMSE for each technique compared to interpolation
including trend: (b) BT + interpolation of corresponding residuals; (c) LT + interpolation of corresponding residuals; and (d) GT + interpolation of corresponding residuals.

IDW
Natural neighbor
Regular spline
Tension spline
Topogrid
Ordinary kriging
Anisotropic kriging

(b)

(c)

(d)

39.22
13.02
9.85
20.41
24.74
23.91
17.12

30.52
13.00
6.07
19.29
17.74
17.18
20.33

23.39
10.94
3.03
15.59
13.28
14.01
14.27

177

tion. Other functions that require more parameters such as beta


pdf and local polynomials can provide better results than global
polynomial, but at additional computational cost. Fig. 7 also show
that RMSE results obtained through exclusive treatment of trends
are not consistent among all techniques and datasets. For example,
accounting for trend though beta pdf has negative impact on spatial interpolation (slight increase in RMSE) for Ohio dataset. The
same is true for local and global polynomial trends for Rainwater
dataset.
Quantitative estimates of how each interpolation techniques is
affected by removing trend in bathymetry is presented in Table 6
for all datasets. Table 6 is derived by computing the percentage
change in RMSE for (b), (c) and (d) interpolation categories compared to (a) for all reaches, and then taking a mean of these percentages. Mean values of percentage change (or improvement) in
RMSE considering all datasets show that IDW is most benetted
among all interpolation techniques; whereas regularized spline
show the least improvement in RMSE. In addition, interpolation
techniques that account for anisotropy (e.g., anisotropic kriging),
and the ones that are more robust for anisotropic data (e.g., natural
neighbor) do not show much change in RMSE for different trend
surfaces compared to other isotropic techniques (e.g., IDW and
topogrid). Overall, excluding trends from bathymetry points before
spatial interpolation is found to produce better results in terms of
RMSE. A more qualitative assessment through visual inspection
and cross-section plots is presented in the next sub-section.
Assessment of interpolated surfaces

percentage change in RMSE range from as little as 6% (with BT for


Sulphur) to as much as 60% (with BT for Kentucky). With the
exception of the Sulphur River, bathymetry estimates after modeling the trend with BT show the most reduction in RMSE followed
by LT and GT. Table 5 show that removing trends in bathymetry
even through a simple function such as global polynomial that require only three parameters can have impact on spatial interpola-

Although RMSE provides information on the accuracy of a


bathymetric surface, it is necessary to understand how this
number affects the spatial representation. First, bathymetry plots
(Fig. 8) created using IDW for King Ranch dataset show how the
surface representation is affected by including/excluding bathymetry trend during spatial interpolation. One advantage (that is

Fig. 8. Interpolated surfaces for area surrounding Y in Fig. 3b for King Ranch: (a) Base surface; (b) IDW of z including trend; (c) beta trend + IDW residuals; (d) local
polynomial + IDW residuals; (e) global polynomial + IDW residuals; and (f) natural neighbor interpolation of z including trend.

178

V. Merwade / Journal of Hydrology 371 (2009) 169181

Elevation (m)

145.0
a

144.0

143.0

142.0

10

20

30

40

50

Distance (m)
Fig. 9. Cross-section at location X for surfaces in Fig 8: (a) Base surface; (b) IDW of z
including trend; (c) beta + IDW residuals; (d) local polynomial + IDW residuals; and
(e) global polynomial + IDW residuals.

evident from Fig. 8) of interpolating residuals, and then adding


trend to residual surface is to minimize the local inuence of
sparse points on interpolation results. For example, the dark circular spots in Fig. 8b are caused due to presence of single bathymetry
points that have insufcient neighboring points to create a reasonable surface. As a result, a spike (near banks) or depression (near
thalweg) is created surrounding such points. On the other hand,
when residuals (bathymetry trend) are interpolated, the magnitude associated with each interpolating point is found to be relatively small for all study reaches to create this local effect, and
when the trend is added back, the overall effect is negligible as
demonstrated in Fig. 8c surface. However, the quality of interpolate
surface is dependent on the effectiveness of the trend surface. For
example, global polynomial which is unable to capture local trends
effectively also produces spikes/depressions in the nal surface,
but these spikes/depressions are still not as prominent as they appear in Fig. 8a. Interestingly, natural neighbor which is robust (not

Fig. 10. Interpolated bathymetry surfaces with natural neighbor for Brazos and Kentucky: (a)/(d) Base surface; (b)/(c) interpolation including trend; and (c)/(f) interpolation
of residuals + trend.

179

V. Merwade / Journal of Hydrology 371 (2009) 169181

12

150

(a)

(b)

(c)

(d)

(e)

147

11

(f)

10

144

141

138

8
0

20

40

60

80

Brazos

20

40

60

80

100

Kentucky

Fig. 11. Cross-section proles at location P and Q for Brazos and Kentucky, respectively for surfaces in Fig. 10: (a)/(d) Base surface; (b)/(c) interpolation including trend; and
(c)/(f) interpolation of residuals + trend.

as sensitive to bathymetric trend) compared to other methods produced identical surfaces for all techniques (including/excluding
trends) for King Ranch. Although this may suggest that a robust
technique is unaffected by how the interpolation is performed
(including/excluding trends), this is not true for all datasets as described in the next paragraph. Only one surface created by natural
neighbor interpolation that includes trend is presented in Fig. 8f for
King Ranch whose IDW counterpart is 8b. The difference in surfaces is directly reected in cross-section proles presented in
Fig. 9.
The robustness of natural neighbor as seen for the King Ranch
dataset is not applicable for all datasets. For example, cross-sections from interpolated surfaces including and excluding trends
using natural neighbor interpolation for Brazos and Kentucky show
signicant differences in results (Figs. 10 and 11). The surfaces created by excluding trend are smoother compared to surfaces created by interpolation including trend. The overall representation
of bathymetric surface created by excluding trend is more representative of measured bathymetry compared to surfaces created
by interpolation of points including trend. The difference in nal
results is more distinct and considerably better for the Kentucky
dataset that includes sparsely located bathymetry points along
cross-sections.
Discussion and conclusions
There is a growing need to obtain accurate bathymetry data for
use in multi-dimensional river models (e.g., Dutta et al., 2007; Shen
and Diplas, 2008; Yue et al., 2008). Typically, river bathymetry data
are collected as cross-sections or discrete points, which are interpolated to get elevation at computational nodes in river models.
Because of the presence of spatial trends in river bathymetry, commonly used isotropic interpolation methods that do not account
for spatial trends yield inaccurate results. The objective of this
study was to use discrete bathymetry data from six study areas
to investigate the effect of trend on spatial interpolation methods
for creating continuous surfaces. Although recent studies have
advocated the use of anisotropic interpolation techniques (e.g.,
Merwade et al., 2006; Tomczak, 1998), the present study shows
that isotropic methods can yield improved results if the trend in
the data is excluded during spatial interpolation. The trend from
the bathymetry data should be excluded at two levels. First, the
data should be mapped using ow oriented (s, n) coordinates to remove the spatial heterogeneity in the orientation of river slope
with respect to Cartesian coordinates. Second, the cross-sectional
trend must be removed from the bathymetry by using a nonlinear
function. After removing the trend, the residuals can be interpolated to create a surface to which the trend is added back to get

the nal bathymetric surface. Results show that the bathymetric


surface created by excluding trends not only gives better RMSE
for validation datasets, but the nal surface is more representative
of the actual bathymetry.
When eld data are limited (sparse cross-sections or bathymetry points), it becomes necessary to make the best use of these data
to make accurate predictions in unmeasured locations. Excluding
the trend from the data and then adding it back after interpolating
residuals insure that the trend is restored in the nal interpolated
surface, thus providing more condence at unmeasured locations.
The idea of separating bathymetry into trend and its residuals is
not new, but it is mainly utilized in the context of kriging (a relatively sophisticated approach compared to other simple methods
such as IDW), and thus not implemented in any existing isotropic
interpolation methods. Because most isotropic methods are not designed to handle the trend in the data explicitly, this trend can get
lost during interpolation of sparse data (e.g., cross-sections) thus
affecting the quality of the nal interpolated surface, and consequently affect model results that rely on bathymetry input. This
study shows that results from IDW can improve by as much as
40% with regard to RMSE when the trend is excluded from interpolation. Improvement in other methods is not as signicant as IDW,
but still better compared to interpolation including trend. Results
from Kentucky dataset that had typical cross-sectional data show
that spatial interpolation using any method can be signicantly
improved through explicit treatment of bathymetry trend.
Although it is difcult to argue against collection of more eld
data, results from this study show that more bathymetry points
are not always useful if channel morphology or trend is ignored
while taking measurements. For example, Rainwater dataset has
much higher density of bathymetry points compared to Brazos
and Ohio, but the measurements are taken along ow lines that
are unable to capture the entire cross-section of the channel at
any location. Inadequate description of cross-sections results in
poor trend surface, which in turn affects the nal interpolation results. Consequently, RMSE of interpolated surface for Rainwater is
nearly equal to that of Brazos and Ohio which are relatively larger/
deeper rivers with much lower point density. Therefore, spatial
arrangement of bathymetric points is equally important in collecting eld data to get a satisfactory interpolated surface from discrete bathymetry points.
Channel trend can be modeled in several ways: process-based
predictive model, a model based on relationship between channel
planform and cross-sectional asymmetry, and a least-square
regression model that ts a user-dened function to observations.
In this study, three regression models (two using polynomial and
one using beta pdf) are used to t bathymetric trends. Overall, it
is found that selection of a trend model has effect on nal

180

V. Merwade / Journal of Hydrology 371 (2009) 169181

interpolated surface. Although any model can be used to dene the


trend, a model that makes the most use of the measured discrete
data to t the trend is recommended to get accurate results
through spatial interpolation. Trend function dened using beta
pdf is found to be the best among all three options considered in
this study followed by local polynomial. Considering the heterogeneity in channel bathymetry, tting of local trends is recommended compared to a global trend that uses a single function
for the whole data. If use of local trend model is impractical due
to computational demands or any other reasons, modeling the
trend using a global function and interpolating the residuals can
still provide better results compared to traditional interpolation
including trend. A trend model that can capture even as little as
20% of variance in the data (global polynomial trend for Kingh
Ranch) can have effect on nal interpolation results. Quality of
trend surfaces should be checked by considering spatial properties
(e.g., semi-variogram) rather than using only pure statistical techniques, which can give spurious results.
Although the general conclusion that exclusive treatment of
spatial trend while interpolating river bathymetry provides better
results is applicable to many other rivers, it should be noted that
more research is needed to look at other factors related to this topic. Findings related to spatial arrangement of bathymetry and
number of data points may be different if channel morphology,
geology and substrate type are also taken into account. Compared
to a sandy river, a gravel bed river may need more measurements
to accurately capture the small scale variations in the bathymetry.
Another enhancement to the present study would be to look at the
role of river ow on the bathymetry, and how ow conditions
should be incorporated during data collection, and consequently
spatial interpolation. Similarly, the effect of different techniques
used in this study also needs to be investigated. For example,
how would the results change if the test and validation data are
split 50/50, or all datasets are sampled randomly (Brazos is sampled manually for creating validation dataset). Investigation of
these factors in detail may shed more light into results obtained
in Fig. 7 that show that not all datasets respond equally to different
trend models and spatial interpolation techniques.
The idea of separating bathymetry into trend and residuals
opens a new area of research in bathymetric modeling: stochastic
simulation of river bathymetry. Although the parameters of trends
functions used in this study are developed by using eld bathymetry data, approximate trend for river bathymetry can also be
developed by using channel planform, width and depth, which
can be obtained from other datasets without actual eld measurements. An aerial photograph can be used to get channel width at
any location, and drainage area can be used to get average channel
width and depth through hydraulic geometry relationships. Crosssectional shape can be modeled by relating channel planform to
cross-sectional asymmetry, and this shape (represented as a function of curvature, channel width and depth) can then be used to
create an approximate trend surface at any location along a river.
This trend must then be combined with a surface of residuals
which can be one of the realizations of several random elds.
The author is pursuing this idea by using sequential Gaussian simulation for some of the datasets used in this study.
Acknowledgments
The author is grateful to Barney Austin (Texas Water Development Board), Tim Osting (previously at Texas Water Development
Board) and Chad Wagner (USGS) for sharing the bathymetry datasets used in this study. Constructive comments from two reviewers
and the editor led to signicant improvement of this manuscript,
and their input is greatly appreciated.

References
Booker, D.J., 2003. Hydraulic modelling of sh habitat in urban rivers during high
ows. Hydrological Processes 17 (3), 577599.
Buttner, O., 2007. The inuence of topographic and mesh resolution in 2D
hydrodynamic modelling for oodplains and urban areas. European
Geosciences Union 2007 Geophysical Research Abstracts 9 (08232).
Carrivick, J.L., 2006. Application of 2D hydrodynamic modelling to high-magnitude
outburst oods: an example from Kverkfjoll, Iceland. Journal of Hydrology 321
(14), 187199.
Chaplot, V. et al., 2006. Accuracy of interpolation techniques for the derivation of
digital elevation models in relation to landform types and data density.
Geomorphology 77 (12), 126141.
Crawford, C.A.G., Hergert, G.W., 1997. Incorporating spatial trends and anisotropy in
geostatistical mapping of soil properties. Soil Science Society of America Journal
66 (1), 298309.
Crowder, D.W., Diplas, P., 2000. Using two-dimensional hydrodynamic models at
scales of ecological importance. Journal of Hydrology 230 (34), 172191.
de Jalon, D.G., Gortazar, J., 2007. Evaluation of instream habitat enhancement
options using sh habitat simulations: case-studies in the River Pas (Spain).
Aquatic Ecology 41 (3), 461474.
Deutsch, C.V., Wang, L.B., 1996. Hierarchical object-based stochastic modeling of
uvial reservoirs. Mathematical Geology 28 (7), 857880.
Dutta, D., Alam, J., Umeda, K., Hayashi, M., Hironaka, S., 2007. A two-dimensional
hydrodynamic model for ood inundation simulation: a case study in the lower
Mekong river basin. Hydrological Processes 21 (9), 12231237.
Gard, M., 2005. Variability in ow-habitat relationships as a function of transect
number for PHABSIM modelling. River Research and Applications 21 (9), 1013
1019.
Goff, J.A., Nordfjord, S., 2004. Interpolation of uvial morphology using channeloriented coordinate transformation: a case study from the New Jersey shelf.
Mathematical Geology 36 (6), 643658.
Hilldale, R.C., Raff, D., 2008. Assessing the ability of airborne LiDAR to map river
bathymetry. Earth Surface Processes and Landforms 33 (5), 773783.
Horritt, M.S., Bates, P.D., Mattinson, M.J., 2006. Effects of mesh resolution and
topographic representation in 2D nite volume models of shallow water uvial
ow. Journal of Hydrology 329 (12), 306314.
Hutchinson, M.F., 1989. A new procedure for gridding elevation and stream line
data with automatic removal of spurious pits. Journal of Hydrology 106, 211
232.
Hutchinson, M.F., 1993. Development of a continent-wide DEM with applications to
terrain and climate analysis. In: Goodchild, M.F. et al. (Eds.), Environmental
Modeling with GIS. Oxford University Press, New York, pp. 392399.
James, L.A., 1996. Polynomial and power functions for glacial valley cross-section
morphology. Earth Surface Processes and Landforms 21 (5), 413432.
Johannesson, H., Parkerm, G., 1989. River Meandering. In: Ikeda Syunsuke, Parker
Gary (Ed.). Water Resources Monograph, vol. 12, Washington, DC, pp. 181
213.
Khosronejad, A., Rennie, C.D., Neyshabouri, S.A.A.S., Townsend, R.D., 2007. 3D
numerical modeling of ow and sediment transport in laboratory channel
bends. Journal of Hydraulic Engineering-ASCE 133 (10), 11231134.
Lampe, D.C., Morlock, S.E., 2007. Collection of bathymetric data along two reaches of
the Lost River within Bluespring Cavern near Bedford, Lawrence County,
Indiana, July 2007. US G.S. Karst Interest Group Proceedings, Bowling Green,
Kentucky, May 2729, 2008: US G. S. Scientic Investigations Report 20085023, 142 p. <http://pubs.water.usgs.gov/sir>2008-5023.
Lane, S.N., Hardy, R.J., Elliott, L., Ingham, D.B., 2002. High-resolution numerical
modelling of three-dimensional ows over complex river bed topography.
Hydrological Processes 16 (11), 22612272.
Lee, K.T., Ho, Y.H., Chyan, Y.J., 2006. Bridge blockage and overbank ow simulations
using HEC-RAS in the Keelung River during the 2001 Nari Typhoon. Journal of
Hydraulic Engineering-ASCE 132 (3), 319323.
Legleiter, C.J., Kyriakidis, P.C., 2008. Spatial prediction of river channel topography
by kriging. Earth Surface Processes and Landforms 33 (6), 841867.
Legleiter, C.J., Roberts, D.A., Marcus, W.A., Fonstad, M.A., 2004. Passive optical
remote sensing of river channel morphology and in-stream habitat: physical
basis and feasibility. Remote Sensing of Environment 93 (4), 493510.
Marcus, W.A., Legleiter, C.J., Aspinall, R.J., Boardman, J.W., Crabtree, R.L., 2003. High
spatial resolution hyperspectral mapping of in-stream habitats, depths, and
woody debris in mountain streams. Geomorphology 55 (14), 363380.
Martin, Y., 2003. Evaluation of bed load transport formulae using eld
evidence from the Vedder River, British Columbia. Geomorphology 53 (12),
7595.
Mekonnen, M.A., Dargahi, B., 2007. Three dimensional numerical modelling of ow
and sediment transport in rivers. International Journal of Sediment Research 22
(3), 188198.
Merwade, V.M., Maidment, D. R., 2004. geospatial description of river channels in
three dimensions. University of Texas at Austin Center for Research in Water
Resources Online Report 04-8. <http://www.crwr.utexas.edu/reports/2004/
rpt04-8.shtml>.
Merwade, V.M., Maidment, D.R., Hodges, B.R., 2005. Geospatial representation of
river channels. Journal of Hydrologic Engineering 10 (3), 243251.
Merwade, V.M., Maidment, D.R., Goff, J.A., 2006. Anisotropic considerations while
interpolating river channel bathymetry. Journal of Hydrology 331 (3-4), 731
741.

V. Merwade / Journal of Hydrology 371 (2009) 169181


Mouton, A., Meixner, H., Goethals, P.L.M., De Pauw, N., Mader, H., 2007. Concept and
application of the usable volume for modelling the physical habitat of riverine
organisms. River Research and Applications 23 (5), 545558.
Nelson, J.M., Dungan J.S., 1989. River Meandering. In: Ikeda Syunsuke, Parker
Gary (Ed.). Water Resources Monograph, vol. 12. Washington, DC, pp. 69
102.
Petersen, C.T., Trautner, A., Hansen, S., 2008. Spatio-temporal variation of anisotropy
of saturated hydraulic conductivity in a tilled sandy loam soil. Soil and Tillage
Research 100 (12), 108113.
Pfannkuch, H.O., Winter, T.C., 1984. Effect of anisotropy and groundwater system
geometry on seepage through lakebeds: 1. Analog and dimensional analysis.
Journal of Hydrology 77 (14), 213237.
Rogala, J.T., 1999, Methodologies employed for bathymetric mapping and sediment
characterization as part of the Upper Mississippi River System Navigation
Feasibility Study, ENV Report 13, Interim Report for the Upper Mississippi River
Illinois Waterway System Navigation Study; prepared for US Army Engineers
District, Rock Island, US Army Engineers District, St. Louis, and US Army
Engineers District, St. Paul. <http://www.umesc.usgs.gov/documents/
bathymetry/methods.pdf>.
Shen, Y., Diplas, P., 2008. Application of two- and three-dimensional computational
uid dynamics models to complex ecological stream ows. Journal of
Hydrology 348 (12), 195214.
Taye, V., Lane, S.N., Hardy, R.J., Yu, D., 2007. A comparison of one- and twodimensional approaches to modelling ood inundation over complex upland
oodplains. Hydrological Processes 21 (23), 31903202.
Tomczak, M., 1998. Spatial interpolation and its uncertainty using automated
anisotropic inverse distance weighting (IDW) cross-validation/Jackknife

181

approach. Journal of Geographic Information and Decision Analysis 2 (2),


1830.
Torizzo, M., Pitlick, J., 2004. Magnitude-frequency of bed load transport in mountain
streams in Colorado. Journal of Hydrology 290 (12), 137151.
TWDB, 2008. Texas Water Development Board Instream Flow Program. <http://
www.twdb.state.tx.us/instreamows/index.html>.
Vermeyen, T. B., 2006. Using an ADCP, depth sounder, and GPS for bathymetric
surveys. In: Proceedings of the ASCE World and Environmental Resources
Congress, Omaha, NE, May 2006.
Wagner, C.R., Mueller, D.S., 2001. Calibration and validation of a two-dimensional
hydrodynamic model of the Ohio River, Jefferson County, Kentucky. Water
Resources Investigations Report 01-4091.
Wagner, C.R., Mueller, D.S., 2002. Use of velocity data to calibrate and validate twodimensional hydrodynamic models. In: Proceedings Paper and Presentation to
Federal Interagency Hydrologic Modeling Conference, Las Vegas, NV, July 28
August 1, 2002.
White, L., Hodges, B.R., 2005. Filtering the Signature of submerged large woody
debris from bathymetry data. Journal of Hydrology 309 (14), 5365.
Yang, J., Townsend, R.D., Daneshfar, B., 2006. Applying the HEC-RAS model and GIS
techniques in river network oodplain delineation. Canadian Journal of Civil
Engineering 33 (1), 1928.
Ye, J., McCorquodale, J.A., 1997. Depth-averaged hydrodynamic model in curvilinear
collocated grid. Journal of Hydraulic Engineering 123 (5), 380388.
Yue, Z.Y., Cao, Z.X., Li, X., Che, T., 2008. Two-dimensional coupled mathematical
modeling of uvial processes with intense sediment transport and rapid bed
evolution. Science in China Series GPhysics Mechanics and Astronomy 51 (9),
14271438.

Vous aimerez peut-être aussi