You are on page 1of 19

133: Rainfall-runoff Modeling of Ungauged

Catchments

GUNTER BLOSCHL
Institute for Hydraulic and Water Resources Engineering, Vienna University of Technology,
Vienna, Austria

Catchments where no runoff data are available are termed ungauged catchments. For these catchments, the
parameters of rainfall-runoff models cannot be obtained by the calibration on runoff data and hence need to
be obtained by other methods. Model parameters that require calibration are usually transposed from similar
gauged catchments. This article reviews concepts for identifying hydrologic similarity as well as methods for
transposing the parameters of both event models and explicit soil moisture accounting (ESMA) models. Model
parameters that are physically based are usually measured or inferred from other data within the ungauged
catchment of interest. This article summarizes the most important methods and discusses the issues of using
point scale field data in rainfall-runoff models. Alternatives to runoff data for model calibration are suggested.
The value of soft data and qualitative field observations is emphasized.

INTRODUCTION principles, so most of the model equations are empirical


in nature and tend to depend on the hydrological setting
Most catchments of the world are ungauged. Even though (Chapter 1 On the fundamentals of hydrological sciences.
more than 60 000 stream gauges are installed worldwide Volume 1). Calibration can account for the effects of the
(WMO, 1995), there are a couple of orders of magni- hydrological setting in a particular catchment. Second,
tude more catchments where no runoff data are available. hydrological models are very much dependent on their
These are termed ungauged catchments. In these catch- boundary conditions, and these are often poorly defined.
ments, rainfall-runoff models must be used to calculate Calibration can adjust for biases in the inputs, for example,
runoff and other variables, given rainfall (and other cli- as a result of orographic effects and instrument biases.
mate variables) as an input. There are both operational and Third, and probably most important, the media properties
academic drivers for pursuing rainfall-runoff modeling of (both soil and vegetation) are highly heterogeneous and
ungauged catchments. The former include design applica- essentially always unknown or at least poorly known. Soil
tions (of spillways, culverts, and embankments), forecasting properties can change dramatically in space but change very
applications (flood warning and hydropower operation), little with time, so parameter calibration can significantly
and catchment management applications (water allocation, enhance the performance of rainfall-runoff models. This
climate impact studies), the latter are geared toward under- led Beven (2000) to remark that: However such models
standing the catchment functioning and how the individual are constructed, they will have some parameters that will
processes combine to produce catchment response. need to be defined for each application site. While the
The main challenge with rainfall-runoff modeling in calibration on runoff data has served hydrology well in
ungauged catchments is the lack of local runoff data that the past, this is not an option in ungauged catchments.
could be used for calibrating model parameters. Parameter Alternatives are needed which are the subject of this article.
calibration is important for a number of reasons. First, with The most efficient method of addressing the issue of
the exception of the water balance equation, there is no ungauged catchments is to glean the model parameters (and
unique hydrological equation that can be derived from first perhaps the model structure) from analogue catchments in

Encyclopedia of Hydrological Sciences. Edited by M G Anderson.


2005 John Wiley & Sons, Ltd.
2 RAINFALL-RUNOFF MODELING

the region, that is, from one or more catchments that one can regions with approximately homogeneous model parame-
expect to behave similarly to the catchment of interest. This ters. The regions are found from an analysis of a small
article therefore first discusses hydrological similarity in the number of gauged catchments in each of the regions (e.g.
context of rainfall-runoff modeling (Section Hydrological Nathan and McMahon, 1990) by exercising expert judg-
similarity in rainfall-runoff modeling). Rainfall-runoff ment supported by whatever hydrologic information that is
models use two broad categories of model parameters: available. Hydrologic information may consist of hydrogeo-
physically based parameters that in principle can be logic maps, climate maps, soil and vegetation maps, and the
observed or estimated directly from measurements in a seasonality of hydrologic response as an indicator of hydro-
catchment; and calibration parameters that appear in empir- logic processes (e.g. Merz et al., 1999). The regions found
ical relationships and need to be back-calculated from in this way will be plausible but it is not straightforward
rainfall and runoff data. The calibration parameters need to to formalize the procedure. Also, the uncertainty associ-
be transposed from similar, gauged catchments and these ated with identifying the regions is not usually known. As
methods are reviewed in Section Transposing calibration an alternative to contiguous regions, interpolation methods
parameters from similar, gauged catchments. Physically such as kriging have been used more recently as one way
based parameters, on the other hand, can be observed in the of exploiting spatial proximity as a similarity measure (e.g.
catchment of interest, although with some restrictions. This Merz and Bloschl, 2004, 2005).
is discussed in the Section Measuring or inferring physi-
2. Similar catchment attributes: The second similarity con-
cally based model parameters in an ungauged catchment.
cept consists of using measurable catchment attributes as
Even though in an ungauged catchment no runoff data are
indicators of hydrological similarity. The assumption is that
available, other hydrologic response data may be available
if catchment attributes such as soil type, vegetation type,
and could be used as an alternative for model calibration
and topographic characteristics of two catchments are simi-
and testing. This is dealt with in the Section Alternatives
lar, one would expect the hydrologic response to be similar.
to runoff data for model calibration.
Again, this is a plausible assumption, although few stud-
ies have conclusively demonstrated the predictive power of
HYDROLOGICAL SIMILARITY IN catchment attributes. Catchment attributes can be used in
RAINFALL-RUNOFF MODELING different ways. The first method is to use them to define
a group of hydrologically similar catchments. The similar-
If the rainfall-runoff processes in two catchments are
ity between catchments is usually quantified by a distance
similar, their hydrologic responses will be similar too.
measure as a function of the differences of the catchment
This is the notion of hydrologic similarity. Similarity of
attributes in the two catchments. The distance measure is
hydrological processes can be defined in various ways.
zero if the catchment attributes in the two catchments are
Dunne (1978) suggested that runoff processes are mainly
identical and increases as the attributes get more dissim-
controlled by physioclimatic controls and identified three
ilar. From the matrix of distance measures, the grouping
main types: Infiltration excess runoff which is generated
can then be obtained by a range of statistical methods such
from partial areas where surface hydraulic conductivities
as cluster analysis, principal component analysis, and clas-
are low; saturation excess runoff which is generated in
sification trees (Nathan and McMahon, 1990; Bates, 1994;
areas with shallow water tables or near-channel wetlands;
Breiman et al., 1984). One particular variant of this method
and subsurface storm flow which is likely to be active and
is the region of influence approach where for each catch-
dominant on steep, humid forested hillslopes with very
ment of interest a separate pooling group is formed (Burn
permeable surface soils. In two similar catchments, the
and Boorman, 1993). Once the groups are identified, the
relative role of each of these processes would be similar.
model parameters (and perhaps the model structure) can
The characteristics of these processes in natural catchments
be transferred from an analogue gauged catchment to an
are never known in full detail so a number of similarity
ungauged catchment within the same group.
concepts have been proposed in the literature that attempt
The second method of using catchment attributes consists
to represent these processes to various degrees.
of defining relationships between model parameters and
1. Spatial proximity: In the first concept, catchments that catchment attributes. There is a wide spectrum of methods
are close to each other are assumed to behave in hydrolog- ranging from multiple regression methods for calibration
ically similar manner. The rationale of this concept is that parameters (see Section Explicit soil moisture account-
the controls on the rainfall-runoff relationship are likely to ing models) to empirical formulae in engineering design
vary smoothly in space, so one can expect spatial proxim- (see Section Event models). The structure of these rela-
ity to be a good indicator of the similarity of catchment tionships is often dictated by parsimony and convenience
response. This has been one of the traditional methods of although some degree of process reasoning can come in.
regionalizing rainfall-runoff model parameters and is usu- The parameters of the relationships are usually based on
ally based on delineating on a map, spatially contiguous the analysis of a large number of gauged catchments. In
RAINFALL-RUNOFF MODELING OF UNGAUGED CATCHMENTS 3

multiple regressions, one may encounter problems with over impervious bedrock. HRUs are often the method
multicollinearity if one or more of the catchment attributes of choice when inferring physically based model param-
is highly correlated with another catchment attribute or with eters from landscape attributes. The main difficulties with
some linear combination of them. Multicollinearity will the use of HRUs is that the estimation of parameters for
reduce the reliability of the regression coefficients (Hirsch multiple HRUs in a catchment is not simple, either in cal-
et al., 1992). One therefore limits the number of catchment ibration or for ungauged catchments (see Chapter 122,
attributes used in the regression, sometimes combining a Rainfall-runoff Modeling: Introduction, Volume 3), and
number of attributes into an index that is deemed represen- that measurable landscape attributes and model parameters
tative of one particular aspect of the rainfall-runoff model are often poorly correlated (see Sections Limitations and
(such as a baseflow index, IH, 1999). Sequential regression challenges and Local scale measurements and relation to
may assist in identifying robust parameter estimates (Calver landscape attributes).
et al., 2004; Lamb and Kay, 2004). The extension of this 3. Similarity indices: Similarity indices are based on some
general approach of using catchment attributes are methods understanding of the structure of runoff generation and
that derive physically based model parameters from land- runoff routing and are usually defined as a dimensionless
scape or catchment attributes (see Sections Local scale number. Two catchments or two parts of the landscape
measurements and relation to landscape attributes and would then behave hydrologically similar if they are
Upscaling local measurements to catchment/model ele- associated with the same value of the similarity index
ment scale). Sometimes, the first and second methods are (Bloschl and Sivapalan, 1995; Aryal et al., 2002; see
combined in using relationships between model parameters also Chapter 3, Hydrologic Concepts of Variability and
and catchment attributes only within regions that are homo- Scale, Volume 1). Similarity indices differ in terms of
geneous with respect to a certain set of catchment attributes the processes they aim to represent. Similarity in climate
(e.g. Burn and Boorman, 1993). A formal way of combining can be quantified by the aridity index of Budyko (1974),
the two methods results in the CART model (Breiman et al., for example, which is the ratio of long-term potential
1984; Laaha and Bloschl, 2005a). In CART models, the evaporation to precipitation. Atkinson et al. (2002) and
independent variables are the catchment attributes and the Farmer et al. (2003) used this type of climatic index for
dependent variables are the model parameters. Regression analyzing suitable rainfall-runoff model complexities as a
trees then divide a heterogeneous domain into a number of function of timescale and climate characteristics. Similarity
more homogeneous regions by maximizing the homogene- in channel flow processes can be quantified by the stream
ity of model parameters and catchment attributes within ordering and other characteristics of the map view of the
each group simultaneously. Regression trees have a number channel topology (e.g. Rodrguez-Iturbe and Valdes, 1979;
of advantages over other models. Their structure is non- Rinaldo et al., 1995). Similarity in catchment processes
parametric, small trees are readily interpretable, there is no can be quantified by topographic indices (Moore et al.,
global sensitivity to outliers, and they are able to handle 1991). Examples include the topographic wetness index
nonlinear relationships well. However, big trees are diffi- of Beven and Kirkby (1979), which is a function of the
cult to interpret, there is a lack of smoothness, and there area drained per unit contour length at a given point
are potential problems with overfitting the data, so the trees and the local slope gradient. Catchment characteristics
need to be pruned (see, e.g. Breiman et al., 1984). such as soil depths may be incorporated in the index.
The third method of using catchment attributes in defin- Each patch in the landscape exhibiting the same value of
ing hydrological similarity has originated from the use of the index is considered to produce the same hydrologic
geographical information systems in hydrological model- response. Two catchments are then deemed similar if their
ing. In this method, patches of the landscape are classified distributions of the index are identical. On the basis of
into groups of similar topographic slope, aspect, elevation, this similarity index, Sivapalan et al. (1987) identified five
vegetation type, soil type, and precipitation distribution, nondimensional similarity parameters that represented the
for example, by overlaying maps of the different types of interrelationships of topography, soil, and rainfall, which
information. Each group is termed a hydrological response lead to similar catchment responses. Larsen et al. (1994)
unit (HRU) and all the elements in a group are assumed and Robinson and Sivapalan (1995) extended this work
to exhibit the same hydrological response characteristics concentrating on the relative dominance of the infiltration
(Leavesley and Stannard, 1995). This again is based on excess and saturation excess mechanisms. The advantage
the assumption that similar catchment attributes would of this type of index is that they are not purely empirical
be associated with similar hydrologic processes. The way so their structure may be more defendable than that of
the layers of information are combined can have various the empirical relationships between model parameters and
degrees of process representation. Flugel (1995), for exam- catchment attributes discussed previously. The drawback is
ple, combined the layers by reasoning such as rangeland that the inputs needed for the indices are often difficult
on gley soil at the valley floor with shallow groundwater to specify in ungauged catchments. Also, tests of some
4 RAINFALL-RUNOFF MODELING

of these indices against spatial soil moisture data have 4. Setting up models relating each rainfall-runoff model
given mixed results, as their applicability depends on parameter to a set of catchment attributes. In this
catchment conditions and climate (e.g. Lamb et al., 1998; step, commonly, multiple linear regressions are used
Western et al., 1999; Blazkova et al., 2002). Similarity (see Section Hydrological similarity in rainfall-runoff
indices are therefore currently not widely used in practice modelling), and some or all of the catchment attributes
although they constitute an active and promising area of are (e.g. logarithmically) transformed.
research. 5. Testing the strength of the relationship of (4), for
The various concepts of hydrological similarity are example, by some goodness-of-fit measure such as a
potentially useful for inferring a suitable model structure correlation coefficient.
in ungauged catchments, but their main application lies 6. Estimating each parameter of the rainfall-runoff model
in transposing model parameters in space. The following for the ungauged catchment from the (regression)
section focuses on the case of calibration parameters. model.
7. Simulating runoff for the ungauged catchment of inter-
est by applying the same model as in (2), using the
TRANSPOSING CALIBRATION PARAMETERS regionally transposed model parameters.
FROM SIMILAR, GAUGED CATCHMENTS 8. Testing the transposition by cross-validation. A gauged
Methods catchment is assumed to be ungauged, runoff is sim-
ulated as in (7) and then compared with the locally
Calibration parameters are those that cannot be measured or observed runoff.
inferred from measurements but need to be transposed from
gauged catchments in the region. These gauged catchments Some of these steps can be skipped depending on the
should be hydrologically similar to the catchment of interest data availability and the similarity concept chosen. In the
in the sense discussed earlier. Rainfall-runoff models can simplest case of using model parameters from one or more
be either event models or ESMA models (see Beven 2001a donor catchments directly in the ungauged catchment, steps
and Chapter 122, Rainfall-runoff Modeling: Introduc- (3 6) can be skipped. Alternatively, results from the lit-
tion, Volume 3). Event models require initial conditions erature can be used for any of the steps. The less the
on the soil moisture status of the catchment and are usually information available in the region of interest, the fewer
run over several hours. ESMA models, in contrast, account the steps that can be carried out, but this is likely at the
for the changes of soil moisture stores in a catchment cost of decreasing model performance. In each of these
through taking stock of evapotranspiration, precipitation, steps it is prudent to not only find the best relationships or
and runoff and are usually run over several years. Different parameters but also to find some measure of the uncertainty
model parameters exist in each type of model: runoff coef- associated with the estimates. In step (1) alternative regions
ficient (or loss) parameters and lag time (or time to peak) and/or donor catchments can be examined and their suitabil-
parameters in the case of event models; and evaporation ity can be tested by cross-validation (step 5); in step (2) for-
parameters, runoff generation parameters and storage coef- mal methods of parameter uncertainty analysis can be used
ficients in the case of ESMA models. The procedures for or less formal sensitivity analyses and split sample methods
transposing these parameters from similar, gauged catch- (see Chapter 131, Model Calibration and Uncertainty
ments in the region to an ungauged catchment are, however, Estimation, Volume 3); in step (3) alternative catchment
similar for the two model types. The transposition typically attributes or combinations of catchment attributes can be
involves the following steps: used and tested in a similar way; and in step (4) alternative
transposition models can be tested. The advantage of the
1. Delineation of homogeneous regions and/or identifi- cross-validation test in step (5) over other methods of
cation of one or more gauged catchments, termed assessing uncertainty, such as sensitivity analyses, is that it
donor catchments, based on any of the similarity mea- examines a combination of a number of sources of uncer-
sures discussed in Section Hydrological similarity in tainty including the data, the model parameters, and the
rainfall-runoff modeling. structure of both the runoff model and the transposition
2. Estimation of model parameters for the donor catch- model.
ments by manual or automatic calibration on observed
runoff data (see Chapter 131, Model Calibration and Event Models
Uncertainty Estimation, Volume 3).
3. Selection of catchment attributes that are deemed to As an example for estimating model parameters for event-
affect catchment response to rainfall. This is either based rainfall-runoff models, the procedure recommended
based on an a priori understanding of what attributes in the UK Flood Estimation Handbook (FEH) (IH, 1999:
may be relevant to a particular model parameter or on http://www.nwl.ac.uk/ih/feh/) will be very briefly
some goodness-of-fit measure. summarized. The general recommendation in the FEH is to
RAINFALL-RUNOFF MODELING OF UNGAUGED CATCHMENTS 5

use runoff data wherever possible including records as short Explicit Soil Moisture Accounting Models
as a year. In ungauged catchments, this is obviously not
Little is known on the process relationships between the
possible. The preferred method for ungauged catchments is
parameters of conceptual ESMA models and the measurable
to use the parameters from a similar (donor) catchment in
catchment attributes. Because of this, the most widely used
the region. As a last resort, relationships between model
method of estimating ESMA model parameters is either by
parameters and catchment attributes can be used. These
using uniform zones or regions (e.g. lithological zones in
involve the largest errors. It is also recommended to
Drogue et al., 2002), or by multiple regressions from catch-
combine the last two methods and scale the parameters
ment attributes. However, most of the case studies pub-
from the donor catchment by the ratio of the parameter
lished in the literature have found rather low correlations.
estimates from catchment attributes in the catchment of
Sefton and Howarth (1998), for example, compared cali-
interest and those in the donor catchment. It is suggested
brated parameters of the IHACRES model with attributes of
to estimate the time-to-peak parameter, if needed, from the 60 catchments in England and Wales. The best correlations
following catchment attributes: mean drainage path slope, they obtained were R 2 = 0.59 between a routing param-
mean drainage path length, extent of urban land cover, eter and percentage of aquifers, and R 2 = 0.69 between
and a parameter representing the proportion of time when an evaporation parameter and mean annual precipitation.
soil moisture deficit was below 6 mm during a standard For the storage parameters no significant correlations were
period. The runoff coefficient can be estimated as the sum obtained. Seibert (1999) related the model parameters of the
of three terms. The first represents the normal capacity of a HBV model (Bergstrom, 1976) to attributes of 11 Swedish
catchment to generate runoff. It varies between catchments catchments within the NOPEX area. The relationships
but does not vary between storms and is estimated from the between forest percentage and snow parameters could be
Hydrology of Soil Types (HOST) soil classes (Boorman interpreted on hydrological grounds but other relationships
et al., 1995). The second term represents the variation in could not. The rank correlation coefficient between a non-
runoff depending on the antecedent soil moisture and can linearity parameter of runoff generation and catchment area
be estimated from mean annual precipitation. The third was R 2 = 0.87 but most other parameters exhibited hardly
term depends on the storm rainfall depth and represents the any significant correlations with catchment attributes. These
nonlinearity in runoff generation. There is also a correction typically low correlations are likely to translate into rather
for urbanization. low model performances for the ungauged catchment case.
There are numerous recommendations in the engineering In Seiberts (1999) study, the median Nash and Sutcliffe
hydrology literature of how to obtain model parameters (1970) model efficiency decreased from 0.81 to 0.79 when
in ungauged catchments, but they all fall into one of moving from calibrated parameters to regionalized param-
the two categories of transpositions from similar, gauged eters for the same set of 11 catchments, but the median
catchments and empirical formulae (e.g. USACE, 1994; efficiencies decreased to 0.67 for a separate set of 7 catch-
Bates, 1994). In the empirical formulae, lag time parameters ments. A model efficiency of 1 indicates a perfect fit while
are often a function of morphometric parameters such smaller values indicate poorer fits. In a Norwegian study,
as topographic slope, stream slope, and flow length (e.g. Beldring et al. (2002) used 141 catchments for calibrating
USACE, 1994). Loss parameters are often a function a version of the HBV model. They then treated 43 addi-
of land cover, soil type, and antecedent soil moisture tional catchments as ungauged and regionalized the model
(such as in the Curve Number method of the US Soil parameters as a function of land use classes. For both sets
conservation service (SCS, 1973; Mishra and Singh, 2003; of catchments they found median NashSutcliffe efficien-
Merz et al., 2005). There is also the option of making use cies of 0.68 and concluded that the regionalization method
of the structure of the empirical equations and adjusting represented the main features of the landscape well. How-
their coefficients by analyzing runoff data in the region ever, for 20% of the second set of stations the efficiencies
which, from a practical perspective, is probably more were less than 0.3.
meaningful than applying any of the (black box) regression In a recent study, Merz and Bloschl (2004) simulated
equations in use. In most studies, the emphasis is on the water balance dynamics of 308 catchments in Aus-
finding a best-fit curve between the runoff model parameters tria using a lumped ESMA model involving 11 calibration
and their controls. The uncertainty around these best-fit parameters. The parameters were calibrated separately on
curves is often immense (Cordery and Pilgrim, 1983). two nonoverlapping 11 year periods of daily runoff data to
Because of this, similar to the suggestion of the FEH, assess the reliability of the calibrated model parameters.
the general recommendation is to analyze runoff data Figure 1 shows two of the parameters. Figure 1(a) gives
from similar gauged catchments in the region whenever the calibrated beta parameter, which is the nonlinearity
possible and use them by one method or another for parameter in the function relating runoff generation to the
estimating the model parameters in the ungauged catchment soil moisture state. There are distinct patterns of low val-
of interest. ues in the West and high values in the East. The regional
6 RAINFALL-RUNOFF MODELING

9
Beta ()
8
7
6
5 Period: 198797 Period: 197686
4
3
2
1
0
(a)

20
k1 (days)
18
16
14 Period: 198797 Period: 197686
12
10
8
6
4 100 km
(b)

Figure 1 Patterns of calibrated model parameters (left: calibration period 1987 1997, right: calibration period
1976 1986). (a) Nonlinearity parameter, beta (). (b) Fast storage coefficient k1 (days) (From Merz and Bloschl, 2004 by
permission of Elsevier). A color version of this image is available at http://www.mrw.interscience.wiley.com/ehs

differences in beta imply a relatively linear rainfall-runoff NashSutcliffe model efficiency decreased from 0.67 to
relationship and large runoff coefficients in the wetter alpine 0.57 when moving from gauged to ungauged catchments
catchment in the West, and a nonlinear rainfall-runoff rela- (Table 1, using kriging for regionalization). For the veri-
tionship and small runoff coefficients in the dryer lowland fication period, the median decreased from 0.63 to 0.56.
catchments in the East. Figure 1 (b) shows the spatial pat- This means that the uncertainty introduced by moving from
terns of the fast storage coefficient. There is a tendency for gauged to ungauged catchments is about twice the uncer-
faster responses in the prealpine catchments of the North tainty of moving from the calibration to the verification
than in the alpine catchments of the South. The regional period. Of particular interest is the change in the bias of
patterns of the parameters for the two periods are similar simulated runoff (i.e. volume errors). While the median val-
although large local differences occur. This suggests that the ues are always small, the scatter between catchments (quan-
calibrated parameters are able to represent the regional or tified by the difference of the 75% and 25% quantiles) dou-
large-scale differences in the hydrological conditions. One bles when moving from the calibration to the verification
would therefore assume that it is possible to derive regional period, but increases by a factor of 5 when moving from
relationships between the calibrated parameter values and gauged to ungauged catchments. This illustrates the value
catchment attributes. Merz and Bloschl (2004), however, of model calibration in reducing bias in gauged catchments
found that the correlation coefficients were relatively low and the uncertainties in ungauged catchments where calibra-
with maximum values of R 2 = 0.27. They then compared tion is not an option Parajka et al. 2005a extended the anal-
various methods for estimating the model parameters in ysis of Merz and Bloschl (2004) by modifying the model
ungauged catchments in terms of their ability of simulating structure and testing additional regionalisation methods.
runoff, judged against runoff data (i.e. a cross-validation; Their model efficiencies were higher than those of Merz and
Table 1, Figure 2). The regionalization methods signifi- Bloschl (2004) but their main findings remained similar.
cantly increased the predictive performance from the cases
of preset parameters (Figure 2a) and global average param- Limitations and Challenges
eters (Figure 2b). The increase in performance depended on
the regionalization method. Average parameters of imme- There are three potential explanations of the relatively
diate upstream and downstream neighbors (Figure 2f) and poor correlations between model parameters and catchment
regionalization by kriging (Figure 2 g) performed somewhat attributes, and the relatively poor performance of both
better than multiple regressions with catchment attributes event-based and ESMA models when parameters estimated
(Figure 2d). For the calibration period, the median of the from catchment attributes are used.
RAINFALL-RUNOFF MODELING OF UNGAUGED CATCHMENTS 7

1 1 1

ME-global regression
ME-preset values

ME-global mean
0.8 0.8 0.8
0.6 0.6 0.6
0.4 0.4 0.4
0.2 0.2 0.2
0 0 0
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
(a) ME-calibration 198797 (b) ME-calibration 198797 (c) ME-calibration 198797

ME-opt. local regression

ME-nested neighbours
1 1 1
ME-local regression

0.8 0.8 0.8


0.6 0.6 0.6
0.4 0.4 0.4
0.2 0.2 0.2
0 0 0
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
(d) ME-calibration 198797 (e) ME-calibration 198797 (f) ME-calibration 198797

1 1
ME-kriging not nested

0.8 0.8
ME-kriging

0.6 0.6
0.4 0.4
0.2 0.2
0 0
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
(g) ME-calibration 198797 (h) ME-calibration 198797

Figure 2 Nash Sutcliffe model efficiencies of regionalized versus calibrated parameters for the period 1987 1997.
(a) preset parameters, (b) global mean of all catchments, (c) global regression with catchment attributes using all
catchments, (d) local regression within a 50 km neighborhood, (e) optimized local regression, (f) average parameters
of immediate upstream and downstream (nested) neighbours, (g) kriging, (h) kriging without nested neighbours (From
Merz and Bloschl, 2004 by permission of Elsevier)

Table 1 Model performance for gauged catchments of pedotransfer functions (see Section Local scale
and ungauged catchments (kriging regionalization) by measurements and relation to landscape attributes).
cross-validation, both for the calibration and the verifi-
cation periods. ME: median Nash Sutcliffe efficiency of
Catchment attributes are usually static indicators and
runoff plus or minus difference of 75% and 25% quantiles may therefore not be very representative of the hydro-
of efficiencies, that is, a measure of scatter. Bias: Median logic functioning of catchments. Also, much of what
relative volume errors of runoff plus or minus difference of is of interest in rainfall-runoff modeling happens
75% and 25% quantiles of relative volume errors. Statistics
beneath the surface while most measurable catch-
are for 308 catchments in Austria. (From Merz and Bloschl,
2004 by permission of Elsevier) ment attributes (such as topographic characteristics and
vegetation) may be representative of the surface but
Gauged Ungauged
catchment catchment
are not normally representative of the hydrologically
active zone in the subsurface. This will be expanded
ME Calibration period 0.67 0.10 0.57 0.18 in Section Measuring or inferring physically based
ME Verification period 0.63 0.11 0.56 0.18
Bias Calibration period 0.00 0.04 0.04 0.21 model parameters in an ungauged catchment of this
Bias Verification period 0.01 0.10 0.04 0.21 article. One avenue to proceed would be to identify
catchment attributes that are more representative of the
processes relevant to rainfall-runoff modeling, perhaps
1. One explanation is that the measurable catchment based on similarity indices, but so far it is not quite
attributes may not be very relevant for catchment clear how to define them at the regional scale.
response. This is certainly the case for soil type 2. The second explanation is that there may be signif-
as reflected by the usually low predictive power icant uncertainty in the calibrated parameter values,
8 RAINFALL-RUNOFF MODELING

which may cloud the underlying relationship between parameters of the Sacramento model from catchment
calibrated model parameters and catchment attributes attributes on the basis of hydrologic reasoning.
(e.g. Gottschalk, 2002). There are methods of account- There are additional sources of uncertainty including
ing for parameter uncertainty in the regionalization of errors in the structure of the rainfall-runoff model and
model parameters (see, e.g. Campbell and Bates (2001), data errors. These are discussed in more detail in Beven
in the case of an event scale rainfall-runoff model illus- (2001a) and also in Chapter 122, Rainfall-runoff Model-
trated on 39 catchments in southwestern Australia). The ing: Introduction, Volume 3, Chapter 131, Model Cali-
examples in Figure 1 suggest that the uncertainty is bration and Uncertainty Estimation, Volume 3.
not very large in this case, but it does seem important
to better constrain model parameters in the calibra-
MEASURING OR INFERRING PHYSICALLY
tion process. There are two possibilities. The first is BASED MODEL PARAMETERS IN AN
to use additional data on state variables in the calibra- UNGAUGED CATCHMENT
tion procedure (Gupta et al., 1998; Mroczkowski et al.,
1997; also see Chapter 131, Model Calibration and Methods
Uncertainty Estimation, Volume 3). For the study Physically based parameters are those that can be observed
area of Figure 1, Parajka et al. (2005b) used snow in the field or estimated directly from measurements of
depth measurements in a multiobjective calibration pro- catchment or channel characteristics without the use of
cedure which lead to better constrained parameters. runoff data. They possess a physical meaning beyond the
The second possibility is regional calibration in which particular model used. Examples include roughness parame-
the model parameters of a number of catchments are ters such as Mannings n, hydraulic conductivity, soil depth,
calibrated simultaneously (e.g. Fernandez et al., 2000; and surface albedo. There are, however, two types of dif-
Szolgay et al., 2002; Engeland and Gottschalk, 2002). ficulties with using measured parameters in rainfall-runoff
Regional calibration will provide more robust param- models that are related to scale (Western and Bloschl, 1999;
eters that are likely to translate into a reduction in Bloschl, 1999). The first difficulty is that the measurement
uncertainty when transposing them to ungauged catch- volume (or support scale) is usually much smaller than the
ments. Merz et al. (2004) also showed that regional model element size. The second is that in most cases there
calibration may increase the correlations between cali- are only a few measurement locations within a catchment
brated model parameters and catchment attributes. and the spacing between the measurements (the spacing
3. The third explanation is that the structure of the model scale) is hence large. This means that the measured param-
relating catchment attributes and model parameters eter is not defined in exactly the same way as in the model
may not be suitable. Indeed, the choice of a linear even if it shares the same name (Beven, 1989). In principle,
regression model is one of convenience rather than one both scale disparities can be addressed by upscaling proce-
based on know relationships. An additional complica- dures. In practice, there is no generally accepted upscal-
tion with the linear regression model is that it is an ing theory (see Section Upscaling local measurements to
unbiased linear estimator of the model parameters but catchment / model element scale) and one often neglects
the incompatibility related to the support and addresses
because the rainfall-runoff models are nonlinear, the
the incompatibility related to the spacing by some sort of
regression model results in biased runoff simulations
interpolation procedure (see Section Upscaling local mea-
in the ungauged catchment. Also, estimating the model
surements to catchment / model element scale). In many
parameters independently from catchment attributes
instances, particularly when remote sensing data are used,
may lead to parameter combinations that are not model parameters are inferred from surrogates (e.g. rough-
very realistic. Nonlinear models (such as CART mod- ness from land use). As the model parameters are often
els, Section Hydrological similarity in rainfall-runoff not very well correlated with the surrogate, this introduces
modelling) exist, and thresholds or bounding condi- additional uncertainties.
tions may be introduced. For example, in the United Because of these difficulties, in gauged catchments phys-
Kingdom, the baseflow index (IH, 1999) may take on ically based parameters are often allowed some degree
a full range of values in low rainfall areas of the country of calibration within a physically justifiable range to
but is always low in high rainfall areas of the country. adjust some of the measurement or estimation biases. For
There have been a number of other attempts to include ungauged catchments, it may therefore be of value to
process reasoning into the estimation of calibration transfer physically based parameters from similar catch-
parameters of ESMA models. Schumann et al. (2000), ments in the region by one of the methods outlined in
for example, imposed the structure on some of the Section Transposing calibration parameters from similar,
equations of estimating the model parameters. Another gauged catchments, in addition to measuring the param-
attempt is that of Koren et al. (2003) who derived eters in the field (or estimating them from remote sensing
RAINFALL-RUNOFF MODELING OF UNGAUGED CATCHMENTS 9

1. Soil hydraulic characteristics: Soil hydraulic character-


istics such as the saturated conductivity, porosity, and soil
water release characteristics are usually estimated from
Predictive performance

infiltration experiments at the plot scale in the field or,


alternatively, from laboratory core tests. If no or only a
few measurements are available, they are sometimes esti-
mated from relationships to soil type (often defined by
% sand, silt, clay, organic matter, and perhaps bulk den-
sity). These relationships are termed pedotransfer functions.
An excellent review of pedotransfer functions and their
d use in hydrology is given in Wosten et al. (2001). The
ure
Deg
ree m eas ters appeal of pedotransfer functions is that soil texture data
e
of c
alibr t of aram
ten are now widely available in databases such as the State
atio
n Ex del p
mo Soil Geographic Data Base (STATSGO) in North America
(USDA, 1991) and the European soil database (Jamagne
Figure 3 Schematic of the value of calibration model et al., 2002). The justification of using pedotransfer func-
parameters versus measuring model parameters for a
typical rainfall-runoff application tions is that the grain-size distribution (defined by the soil
type) should also be relevant to the pore-size distribution
(which in turn is related to soil hydraulic properties). Unfor-
data). In fact, estimating the parameters from two or more tunately, this is not often the case because peds and cracks,
different sources will always increase their reliability. The rather than the grain-size distribution, tend to dominate
distinction between calibration parameters and physically the hydraulic properties. It is therefore not uncommon for
based parameters is hence not a sharp one, there is a gradual the soil properties to vary as much between soil types as
transition from what one would call a calibration parameter within a soil type (e.g. Warrick et al., 1990) and for other
and a physically based parameter, depending on the type influences such as terrain to be important to soil hydraulic
and extent of information available in any particular case. properties (Gessler et al., 1995). This makes Wosten et al.
Figure 3 shows a schematic of the relative value of the (2001) conclude that pedotransfer functions are sufficiently
calibration and the measurement of model parameters for accurate for interpolation purposes between soil hydraulic
a typical rainfall-runoff model application. Both calibrat- measurements in the catchment of interest whereas they are
ing and measuring a model parameter tend to enhance the not recommended to be used in catchments where no mea-
predictive performance of the model. Ideally one should surements are available. If no measurements are available,
pursue both avenues. The schematic also indicates that, a minimum requirement for the use of pedotransfer func-
while it is useful to measure physically based parameters, it tions would seem to be that the conditions in the catchment
is even more useful to have access to runoff data in a par- of interest are similar to those in the catchments in which
ticular catchment to calibrate or adjust one or more of the they were derived (see Section Hydrological similarity in
model parameters. The following section reviews methods rainfall-runoff modeling), though it is worth noting that
of inferring model parameters from measurements in the the pedotransfer functions are generally derived from mea-
catchment of interest. surements on small soil cores or samples.
Local Scale Measurements and Relation to 2. Surface roughness: Surface roughness parameters such
Landscape Attributes as Mannings n are usually determined by in situ exper-
iments on irrigation plots (e.g. Hessel et al., 2003). If no
Physically based model parameters can be inferred from measurements are available, tabulated values in the litera-
field measurements or from remote sensing data. The ture are often used, which are a function of land cover and
former are usually local scale measurements and hence sometimes topographic slope (Engman, 1986). Land-cover
best suited for small catchments while the latter are more
type can be obtained either from field surveys or from anal-
appropriate for large catchments. Often, parameters are
yses of satellite data. Roughness can also change with the
inferred from qualitative information or surrogates. In the
flow rate (e.g. Sepaskhah and Bondar, 2002). Again, ide-
following, a brief summary is given of the three main
ally, the conditions of the application should be similar to
categories of model parameters as used in rainfall-runoff
those where the roughness values were measured.
models and how they can be related to surrogates that
are more widely available than the model parameters 3. Vegetation characteristics: The Leaf Area Index (LAI),
themselves. A practical discussion of some of the parameter the Fraction of Green Vegetation (FGV), and the Fraction
estimation methods is given in Duan et al. (2001) and Vieux of Absorbed Photosynthetic Active Radiation (fAPAR ) are
(2001). vegetation characteristics that can be used in rainfall-runoff
10 RAINFALL-RUNOFF MODELING

models to estimate evapotranspiration. These vegetation generation mechanism that occurred for a given catchment
characteristics can be related to land-cover classes although state in the 4.6 km2 Wernersbach catchment in Germany.
the relationships are not always unique (Hall et al., 1995; They then developed an expert system, known as FLAB,
Gorte, 2000; Kite and Droogers, 2000). Land-cover classes, that estimates the dominant runoff mechanism for a given
in turn, can be estimated from a range of satellite data point in the landscape and a given event size. The mech-
on the basis of indices such as the normalized difference anisms included were Hortonian overland flow, saturation
vegetation index (ndvi). There exist numerous satellite- area overland flow, interflow, recharge, and storage. The
(e.g. AVHRR and Landsat) based land-cover maps such as expert system is based on rules. For example, potentially
the European CORINE land-cover data set (Buttner et al., contributing areas are identified on the basis of connectivity
2002), data sets for North America (see, e.g. Gallo et al., to the stream, terrain slope, and soil characteristics includ-
2001), as well as global data sets (e.g. Hansen et al., 2000; ing layering. If, for a given event, the saturation deficit is
Tucker et al., 2004). Scale incompatibilities may be less exceeded, these areas become active saturation areas and
stringent with satellite data than with ground data although contribute to saturation area overland flow. The rules are
scale issues do remain (e.g. Brunsell and Gillies, 2003). based both on hydrologic reasoning and field mapped runoff
mechanisms for certain event types. Figure 4 shows a test
Upscaling Local Measurements to of the FLAB model against the field mapped areas for two
Catchment/Model Element Scale runoff mechanisms, saturation area runoff (Figure 4 (a)),
and interflow (Figure 4 (b)). Each figure is a composite of
As mentioned previously, the support scale incompati-
the observed and FLAB simulated binary pattern of the
bility is usually neglected but some sort of interpolation
presence/absence of the given runoff mechanism. Black
is often used to bridge the spacing scale incompatibil-
indicates areas where both model and field observations
ity (Bloschl and Sivapalan, 1995; see also Chapter 6,
identify the mechanism, grey indicates areas where pres-
Principles of Hydrological Measurements, Volume 1).
ence of the mechanism is modeled but not observed, and
The interpolation procedures usually build on the surrogate
light grey indicates areas where the presence of the mecha-
variables, discussed in the section Local scale measure-
nism is observed but not modeled. Overall, there is a close
ments and relation to landscape attributes, either based
match. Saturation areas mainly occur close to the stream,
on different classes of the surrogates (land-cover class,
interflow occurs at some distance from the stream, and this
soil type class, etc.) or on regression relationships or a
general pattern is modulated by soils and terrain slope. The
combination of the two. The classification approach is
strength of the FLAB approach is that the rules are based on
closely related to the Hydrologic Response Unit concept
field scale observations rather than on laboratory-scale mea-
(Section Hydrological similarity in rainfall-runoff model-
surements. This general method has the potential of being
ing) where each HRU is a unique combination of attribute used in ungauged catchments for identifying runoff mech-
classes (land cover, soil type, terrain slope, etc.). Both anisms and hence upscaling local measurements. Although
regressions and classifications based on HRU concepts are the rules are not likely to be universal, identifying these
widely used in the literature to upscale measurements (see, rules for certain catchment regime types seems a promising
Bloschl and Grayson, 2000; see also Chapter 11, Upscal- area of research (Scherrer and Naef, 2003; Woods, 2003).
ing and Downscaling Dynamic Models, Volume 1).
Examples include Busch et al. (1999) and Bormann et al.
Soft Data and Qualitative Field Observations
(1999) who identified HRUs based on a cluster analysis of
actual evapotranspiration, groundwater recharge, and sur- Strategies of upscaling such as those of Peschke et al.
face runoff, and then performed one single simulation for (1999a,b) are partly based on qualitative information which
each HRU. Upscaling methods based on remotely sensed one may term soft data. In addition to upscaling local
data are summarized in Stewart et al. (1996). Upscaling measurements, this type of soft information can be used
methods based on geostatistical concepts are reviewed in for estimating or, at least, for making an educated guess
Bierkens et al. (2000) and also in Chapter 9, Statistical on the magnitude of model parameters in ungauged catch-
Upscaling and Downscaling in Hydrology, Volume 1. ments. Clearly, site visits will be instrumental in this type
There is considerable uncertainty associated with any of of assessment. The value of qualitative information is illus-
these upscaling methods and the magnitude of the uncer- trated by an example from the Austrian Alps. Wienerbruck
tainty is not always clear. While the standard procedure is to and Mitterbach are two adjacent catchments that are similar
relate the model parameters to landscape attributes either by in size and are similar in terms of their catchment attributes
classification or regression methods (which are both black as available at the regional scale (Table 2). Wienerbruck
box methods), there are also initial attempts at incorpo- is slightly steeper which would suggest somewhat faster
rating hydrological understanding. An excellent example response but the channel lengths are slightly larger, so
is presented in Peschke et al. (1999a,b) who, based on most empirical equations suitable for ungauged catch-
many years of field experience, mapped the type of runoff ments would give similar estimates of time-to-peak model
RAINFALL-RUNOFF MODELING OF UNGAUGED CATCHMENTS 11

Saturation area runoff Interflow

Modeled only
500 m
Observed only
Observed and modeled
(a) (b)

Figure 4 Comparison of the FLAB model estimates against field mapped areas of the presence of two runoff
mechanisms (a): saturation area runoff, (b): interflow, for a 5 h, 25 mm event and wet antecedent soil moisture
status (Peschke et al., (1999b). Reproduced with permission of IHI Zittau). A color version of this image is available at
http://www.mrw.interscience.wiley.com/ehs

parameters in the two catchments. Similarly, surrogate- as the dominant geologic formation in both catchments.
based estimates of physically based model parameters A more detailed survey indicates that the difference in
would give similar values, as land use and soil types are response is likely a function of a number of factors, includ-
similar. Since these two catchments are gauged, one can ing topography, geology, and the valley bottom infill. In
test this assessment. Table 3 shows the event characteris- both catchments, dolomites from the Upper Trias prevail
tics of the largest event on record (August 2, 1991). While although in slightly different variants (Dachsteindolomit in
runoff coefficients in the two catchments were similar, the Mitterbach and Wettersteindolomit in Wienerbruck). In Mit-
response time (here in terms of the recession constant near terbach, moraines have accumulated on some of the valley
peak) in the Mitterbach was almost 10 times that of the bottoms while in Wienerbruck the valley bottoms are nar-
Wienerbruck catchment. It would be clearly impossible to row without infill. In a small part of the upper Mitterbach
predict the differences in catchment response between the catchment, karstic limestone appears to exist giving rise to
two catchments on the basis of the quantitative catchment karstic springs. Soils tend to be deeper in Mitterbach than
attributes. In contrast, soft information obtained through in Wienerbruck. In a post hoc analysis one can speculate
a visual examination of the catchments during site vis- that the steeper slopes along with more rainfall have lead
its may help tremendously. Figure 5 shows photographs to feedback effects between runoff processes, soil devel-
that are representative of the landforms in each of the opment, and vegetation development through some sort of
catchments. In the slow response catchment (Mitterbach, coevolution (see e.g. Hoosbeek and Bryant, 1993; Lucas,
(a)) the stream channel is mossy. Quite clearly, there is 2001; see also, Chapter 12, Co-evolution of Climate, Soil
very little hydrologic activity. In contrast, the fast-response and Vegetation, Volume 1). In an a priori mode, as needed
catchment (Wienerbruck, (b)) exhibits signs of erosion rills in ungauged catchments, there is clearly a very important
and deeply incised channels, that is, this is a hydrologi- role of this type of field assessment.
cally highly active catchment. From the visual assessment There have been few attempts at formalizing the soft
it would not be impossible to predict the magnitude differ- information from an expert assessment of the catchment
ences in response timing. It is clear that physical controls functioning of the type in the above example. One notable
do lead to these differences in catchment response but these exception is presented in Seibert and McDonnell (2002).
are not apparent in the quantitative catchment attributes They proposed a method where qualitative knowledge from
usually available in rainfall-runoff analyses. The geologic field surveys is made useful through defining a fuzzy
database available at the regional scale indicates limestone membership function that can be used in calibrating a
12 RAINFALL-RUNOFF MODELING

Table 2 Catchment attributes of two adjacent catchments ALTERNATIVES TO RUNOFF DATA FOR
in the Austrian Alps MODEL CALIBRATION
Otscherbach Groe Erlauf
at Wienerbruck at Mitterbach
So far this article has focused on methods of parameter
estimation that do not use any runoff data for calibration
Catchment area (km2 ) 36.1 29.7 in the catchment of interest as the article is concerned with
Mean topographic 1013 984 the case of ungauged catchments. Even if no runoff data are
elevation (m)
Mean topographic 30.0 22.3 available there may exist other hydrologic response data
slope (%) that can be used to calibrate some or all of the model
Maximum channel 10.2 8.7 parameters. In the following, two examples are given to
length (km) illustrate the potential of hydrologic response data. The first
Geology Limestone Limestone example is set in the 90 km2 Schneealpe area in the Austrian
Soils Rendzina Rendzina
Forest (%) 84.0 81.0 Alps (Bloschl et al., 2002). In this study, calibration of
Grass (%) 14.2 16.7 model parameters on runoff data was not possible because
Rock (%) 1.9 0.0 the catchment was highly Karstic and it was unclear what
Mean annual 1680 1415 the hydrologic catchment area was. The catchment was
precipitation (mm) therefore treated as ungauged and the model parameters
were calibrated on alternative response data. The main
Table 3 Characteristics of the largest flood event on interest in the study was on snow processes, so snow cover
record of the two catchments of Table 2 and recession patterns were deemed to be suitable response data. Snow
constants of five largest events cover patterns for the years 19982000 were derived from
Otscherbach at Groe Erlauf Systeme Probatoire pour lObservation de la Terre SPOT
Wienerbruck at Mitterbach XS images based on an unsupervised isodata technique
Event rainfall depth (mm) 255 167 (Jensen, 1996), separately for different illumination classes.
of event on August 2, The classification produced three class patterns of snow,
1991 no-snow, and partial coverage. The rainfall-runoff model
Runoff coefficient ( ) of 0.78 0.73 focused on the snow component and simulated the snow
event on August 2, 1991 pack evolution and snow melt for each grid cell on the
Recession constant near 3.8 36.1
peak (h) of event on basis of an energy balance approach. Wind drift of snow
August 2, 1991 was represented by a wind drift factor, which was a
Recession constant near 6.4 1.8 35.9 7.9 function of terrain elevation, slope, and curvature. For the
peak (average of 5 calibration of the model parameters, four SPOT-derived
largest events cover patterns were chosen from each of the years 1998
standard deviation) (h)
and 1999. These eight patterns were compared with the
simulated patterns for the same dates. The calibration
rainfall-runoff model. Seibert and McDonnell tested their proceeded in two steps. In the first step, snow albedo
method in the Maimai research catchment in New Zealand. and the threshold air temperature for separating snowfall
They achieved very good fits for a three-box model and rainfall were calibrated. The objective function was
when optimizing the parameter values with only runoff based on differences in the snow-covered area of simulation
data, but parameter sets obtained in this way showed in and observation, both lumped for the entire study area
general a poor goodness of fit for other criteria such and stratified by terrain slope and aspect. This made it
as the simulated new water contributions to peak runoff. possible to separate the effects of albedo from those of
Inclusion of soft data criteria in the model calibration the threshold temperature through differential melting on
process resulted in lower runoff model efficiency values north and south facing slopes. In the second step, the
but led to a better overall performance, as interpreted by the wind drift factor was refined. To this end, a combined
experimentalists view of catchment runoff dynamics. One error map of cover for the eight patterns was calculated
of the difficulties with this type of analysis is that normally (Figure 6a). This map represents the average percent error
it will be necessary to introduce additional parameters in snow cover estimation and it was assumed that this
to make use of tracer data (be it artificial tracers or error stemmed largely from the representation of snow
natural tracers including geochemical data) because wave drift. The parameters of the snow-drift model were then
velocities and flow velocities are not necessarily the same as individually calibrated for each pixel to minimize the error.
assumed in their study. However, this general approach of Figure 6(b) shows the error pattern for the calibrated model
including soft information does appear to have considerable and illustrates the significant improvement over the pattern
potential for application in ungauged catchments provided in Figure 6(a). Remaining error is due to sources other than
soft information in the catchment of interest is available. wind drift, or that cannot be explained by the structure
RAINFALL-RUNOFF MODELING OF UNGAUGED CATCHMENTS 13

(a) (b)
70

60

50
Discharge (m3s1)

40

30

20

10

0
0 20 40 60 80 100 120 140 160 180
(c) Time (hrs)

Figure 5 (a) Photographs that are representative of the landforms in the Mitterbach (a) and the Wienerbruck
(b) catchments (Tables 2 and 3). (c) Hydrograph of the largest flood on record (thick line: Mitterbach; thin line:
Wienerbruck). A color version of this image is available at http://www.mrw.interscience.wiley.com/ehs

of the wind drift representation. The model was then of the rainfall-runoff model cannot be estimated from a
applied to a different snow season (2000) in a classical comparison with observed snow patterns, the snow related
split sample test. Figure 6(c) shows the error map for this parameters clearly can.
year and illustrates that, while there is some deterioration The second example is taken from Bauer (2004). His
in simulation compared to the fully calibrated year, the study is set in the Okavango Delta which is a large alluvial
revised snow-drift model is a major improvement over fan situated in northwestern Botswana. This wetland is
the original form (Figure 6a). While subsurface parameters fed by the Okavango River, which flows from the tropical
14 RAINFALL-RUNOFF MODELING

(a)

(b)

(c)
50% 0% 50%
Underestimation Overestimation

Figure 6 Bias in snow cover patterns based on residuals of comparisons between observed and simulated
patterns for the Schneealpe region. (a) precalibration using eight observed patterns in 1998 1999, (b) postcalibration
using eight observed patterns in 1998 1999, (c) validation using four observed patterns in 2000 (Bloschl
et al. (2002). Reproduced with permission of Spinger-Verlag Wien). A color version of this image is available at
http://www.mrw.interscience.wiley.com/ehs

highlands of Angola into the Kalahari basin. The core wet- (NOAA), Advanced Very High Resolution Radiometer
land, which is permanently flooded, is around 6000 km2 (AVHRR), and Landsat satellite images by unsupervised
in size; the surrounding seasonal floodplains add another classification (McCarthy et al., 2003). A total of 150 inun-
6000 km2 . Runoff data would be extremely difficult to dation patterns over the period of 19702000 was available
acquire as the flow occurs in numerous channels and on for calibration. The hydrological model used was physically
the marsh land. The catchment was therefore treated as based and represented surface routing by Mannings equa-
ungauged and the model parameters were calibrated on tion, infiltration into the unsaturated and saturated zones
alternative response data. The main interest in the study by a simple leakage approach, two-dimensional regional
was on the runoff processes in the wetland, so inunda- groundwater flow, and evaporation from the saturated and
tion patterns were deemed to be suitable response data. unsaturated zones. The main calibration parameters were
The inundation patterns (either land or water) were derived Mannings roughness of surface flow, the hydraulic con-
from National Oceanic and Atmospheric Administration ductivity of the swamp cells, the transmissivity of the sand
RAINFALL-RUNOFF MODELING OF UNGAUGED CATCHMENTS 15

aquifer, and the vertical leakance between the surface and as land use and soil type but, again, there are signifi-
the subsurface layer (i.e. an infiltration parameter). Three cant uncertainties and comparisons with adjusted param-
objective functions were used in a manual calibration pro- eters from similar, gauged catchments in the same region
cedure and they all involved a comparison of simulated are extremely useful. As an alternative, or in addition to
and observed inundation patterns for the same date. The these parameter estimation methods, hydrologic response
first objective function was based on the total size of the data other than runoff may be available in the catch-
flooded area, the second on a pixel-by-pixel comparison, ment of interest and these may assist in calibrating model
and the third attempted to match the number of fringe parameters.
points of the observed and simulated inundation patterns. Some guidance for selecting the model structure may
Resulting parameters were all within a physically plausible also be needed in ungauged catchments. Similar to gauged
range. catchments, model structure choice depends on the problem
There are more alternatives to runoff that can be used at hand, data availability, and the runoff processes one is
for calibrating rainfall-runoff models in ungauged catch- to represent (see Chapter 122, Rainfall-runoff Modeling:
ments. Suitable variables depend on the type of hydrologi- Introduction, Volume 3). An analysis of similar, gauged
cal process of interest (Grayson et al., 2002). Observed soil catchments in the same region can provide some assis-
moisture would be an obvious choice for explicit soil mois- tance in the latter step. Acquiring confidence in the model
ture accounting models, but it is difficult to obtain spatially through model tests is not straightforward in the ungauged
representative soil moisture measurements in catchments catchment case but pursuing more than one avenue of
(Western et al., 2002, 2004). Ground measurements may parameter estimation can be of assistance. For example,
be representative of the entire root zone but are usually transposing parameters from similar gauged catchments,
limited to a few spots in a catchment. Spaceborne esti- inferring them from measurements in the ungauged catch-
mates of soil moisture can be retrieved for large areas ment, and testing the model against hydrologic response
(e.g. Wagner et al., 2003) but their main limitations are data that may be available in the ungauged catchment
the shallow penetration depths, which are much smaller may provide some indications on model reliability through
than the root depth represented in many hydrologic models. complimentary information. Cross-validation is also use-
Although there are methods of dealing with this incom- ful as it tests a number of uncertainties in a combined
patibility in a simplified way (Houser et al., 2000; Walker way. Additionally, some of the methods of uncertainty
et al., 2001; Schuurmans et al., 2003), challenges remain. assessment reviewed in Chapter 131, Model Calibra-
Various types of field based soft data, as discussed in tion and Uncertainty Estimation, Volume 3 are appli-
Section Soft data and qualitative field observations, are cable to the ungauged catchment case. In all these steps
also potentially useful for assisting in model calibration as there is a very important role of qualitative field obser-
demonstrated in Seibert and McDonnell (2002). Soft data vations, if only by visual examination of the landforms.
include saturation areas as mapped in the early work of Qualitative field observations may greatly assist in judg-
Dunne and Black (1970) and Dunne et al. (1975). There ing the catchment functioning and hence in testing the
is renewed interest in saturation patterns as illustrated by plausibility of model parameters and the model struc-
a number of mapping projects occurring (e.g. Kirnbauer ture.
et al., 2005) and their application in constraining param- It may seem strange to end a review of rainfall-
eter estimation in rainfall-runoff modeling (e.g. Franks runoff modeling in ungauged catchments with a note
et al., 1998). on the value of runoff data, but that in my opinion
is the state of the science. As highlighted throughout
this article, there are methods for dealing with rainfall-
CONCLUDING REMARKS runoff modeling in ungauged catchments, but signifi-
cant uncertainties remain. Short-runoff data series may
In ungauged catchments, no runoff data are available for contain very valuable information (Vogel and Kroll,
calibrating model parameters, so alternative methods are 1991; Laaha and Bloschl, 2005b) so the single best
needed. This article has reviewed numerous methods that recommendation on the issue of rainfall-runoff model-
depend on the type of parameters, whether they are cal- ing of ungauged catchments may be to install a stream
ibration parameters associated with one particular model, gauge!
or physically based parameters with some meaning beyond
the model used. Calibration parameters are preferably trans-
Acknowledgment
posed from similar, gauged catchments in the same region.
Estimating calibration parameters from catchment attributes I am grateful to Ralf Merz for providing valuable comments
is generally not recommended for a number of reasons. on an early draft of this manuscript and to Keith Beven
Physically based parameters can be inferred from measure- and an anonymous reviewer for comments that helped
ments, often based on widely available surrogates such improve the manuscript. I would also like to thank Dieter
16 RAINFALL-RUNOFF MODELING

Gutknecht, Rodger Grayson, and Siva Sivapalan for the Bloschl G. and Grayson R. (2000) Spatial observations and
numerous insightful discussions on the subject matter over interpolation. In Spatial Patterns in Catchment Hydrology:
the years. Financial assistance from the Austrian Academy Observations and Modelling, Chap. 2, Grayson R. and
of Sciences and the Austrian Science Foundation (project Bloschl G. (Eds.), Cambridge University Press: Cambridge, pp.
no P14478-TEC) is gratefully acknowledged. 17 50.
Bloschl G., Kirnbauer R., Jansa J., Kraus K., Kuschnig G.,
Gutknecht D. and Reszler Ch. (2002) Einsatz von Fern-
erkundungsmethoden zur Eichung und Verifikation eines
FURTHER READING flachendetaillierten Schneemodells (Using remote sensing
Beven K.J. (2001b) How far can we go with distributed methods for calibrating and verifying a spatially distributed
hydrological modelling? Hydrology and Earth Systems snow model). Osterreichische Wasser- und Abfallwirtschaft, 54,
Sciences, 5(1), 1 12. 1 16.
Bloschl G. and Sivapalan M. (1995) Scale issues in hydrological
modelling a review. Hydrological Processes, 9, 251 290.
REFERENCES Boorman D.B., Hollis J.M. and Lilly A. (1995) Hydrology of Soil
Types: A Hydrologically Based Classification of the Soils of the
Aryal S.K., OLoughlin E.M. and Mein R.G. (2002) A similarity United Kingdom, IH Report No. 126, Institute of Hydrology,
approach to predict landscape saturation in catchments. Water Wallingford.
Resources Research, 38(10), 1208, doi:10.1029/2001WR00 Bormann H. Diekkruger B. and Renschler C. (1999)
0864. Regionalisation concept for hydrological modelling on different
Atkinson S.E., Woods R.A. and Sivapalan M. (2002) Climate and scales using a physically based model: results and evaluation,
landscape controls on water balance model complexity over Physics and Chemistry of the Earth, Part B: Hydrology, Oceans
changing timescales. Water Resources Research, 38(12), 1314, and Atmosphere, 24(7), 799 804.
doi:10.1029/2002WR001487. Breiman L., Friedman J.H., Olshen R. and Stone C.J. (1984)
Bates B.C. (1994) Regionalisation of hydrologic data: a Classification and Regression Trees, Wadsworth International
review. Cooperative Research Centre for Catchment Hydrology, Group: Belmont.
Monash University: Victoria, p. 61. Brunsell N.A. and Gillies R.R. (2003) Scale issues in land-
Bauer P. (2004) Flooding and Salt Transport in the Okavango atmosphere interactions: implications for remote sensing of the
Delta, Botswana: Key Issues for Sustainable Wetland Manage- surface energy balance. Agricultural and Forest Meteorology,
ment, Diss., Naturwissenschaften, Eidgenossische Technische 117, 203 221.
Hochschule ETH Zurich, Nr. 15436. Budyko M.I. (1974) Climate and Life, Academic Press: Orlando,
Beldring S., Roald L.A. and Voks A. (2002) Avrenningskart for p. 508.
Norge (Runoff map for Norway, in Norwegian), Report No. 2, Burn D.H. and Boorman D.B. (1993) Estimation of hydrological
Norwegian Water and Energy Directorate, Oslo. parameters at ungauged catchments. Journal of Hydrology, 143,
Bergstrom S. (1976) Development and Application of a 429 454.
Conceptual Runoff Model for Scandinavian Catchments, Busch G., Sutmoller J., Kruger J.-P. and Gerold G. (1999)
Department of Water Resources Engineering, Lund Institute of Regionalization of runoff formation by aggregation of
Technology, Bulletin Series A-52, Swedish Meteorological and hydrological response units: a regional comparison. In
Hydrological Institute: Norrkoping, p. 134. Regionalization in Hydrology, Diekkruger B., Kirkby M.J. and
Beven K. (1989) Changing ideas in hydrology the case of Schroder U. (Ed.), IAHS Publication No. 254, IAHS Press: pp.
physically based models. Journal of Hydrology, 105, 157 172. 45 53.
Beven K.J. (2000) Uniqueness of place and process representa- Buttner G., Feranec J. and Jaffrain G. (2002) Corine Land Cover
tions in hydrological modelling. Hydrology and Earth Systems Update 2000: Technical Guidelines, Technical Report No. 89,
Sciences, 4(2), 203 213. European Environment agency, Copenhagen.
Beven K.J. (2001a) Rainfall-runoff Modelling The Primer, John Calver A., Kay A.L., Jones D.A., Kjeldsen T., Reynard N.S. and
Wiley and Sons, p. 360. Crooks S. (2004) Flood frequency quantification for ungauged
Beven K.J. and Kirkby M.J. (1979) A physically-based, variable sites using continuous simulation: a UK approach. In Com-
contributing area model of basin hydrology. Hydrological plexity and Integrated Resources Management, Transactions of
Sciences Bulletin, 24, 43 69. the 2nd Biennial iEMSs Meeting, Pahl-Wostl C., Schmidt S.,
Bierkens M.F.P., Finke P.A. and de Willigen P. (2000) Upscaling Rizzoli A.E. and Jakeman A.J. (Eds.), International Environ-
and Downscaling Methods for Environmental Research, Kluwer mental Modelling and Software Society (iEMSs): Manno, pp.
Academic Publishers: p. 190. 1214 1218.
Blazkova S., Beven K., Tacheci P. and Kulasova A. Campbell E. and Bates B. (2001) Regionalization of rainfall-
(2002) Testing the distributed water table predictions of runoff model parameters using Markov Chain Monte Carlo
TOPMODEL (allowing for uncertainty in model calibration): samples. Water Resources Research, 37(3), 731 739.
the death of TOPMODEL? Water Resources Research, 38(11), Cordery I. and Pilgrim D.H. (1983) On the Lack of Dependence
10.1029/2001WR000912. of Losses from Flood Runoff on Soil and Cover Characteristics,
Bloschl G. (1999) Scaling issues in snow hydrology. Hydrological Proceedings of the Hamburg Symposium, IAHS Publication No.
Processes, 13, 2149 2175. 140, IAHS: Wallingford, pp. 187 195.
RAINFALL-RUNOFF MODELING OF UNGAUGED CATCHMENTS 17

Drogue G., El Idrissi A., Pfister L., Leviandier T., Iffly J.-F. and Grayson R., Bloschl G., Western A. and McMahon T. (2002)
Hoffmann L. (2002) Calibration of a parsimonious rainfall- Advances in the use of observed spatial patterns of catchment
runoff model: a sensitivity analysis from local to regional scale. hydrological response. Advances in Water Resources, 25,
Proceedings of iEMSs 2002 Integrated Assessment and Decision 1313 1334.
Support 24 27 June 2002 Lugano, Vol. 1, International Gupta H.V., Sorooshian S. and Yapo P.O. (1998) Toward
Environmental Modelling and Software Society: pp. 464 469. improved calibration of hydrologic models: multiple and
Duan Q., Schaake J. and Koren V. (2001) A priori estimation noncommensurable measures of information. Water Resources
of land surface model parameters. In Land Surface Hydrology, Research, 34(4), 751 763.
Meteorology, and Climate: Observations and Modeling. Water Hall F.G., Townshend J.R. and Engman E.T. (1995) Status of
Science and Application, Vol. 7, American Geophysical Union: remote sensing algorithms for estimation of land surface state
pp. 77 94. parameters. Remote Sensing of Environment, 51(1), 138 156.
Dunne T. (1978) Field studies of hillslope flow processes. In Hansen M., DeFries R., Townshend J.R.G. and Sohlberg R.
Hillslope Hydrology, Kirkby M.J. (Ed.), Wiley: Chichester, pp. (2000) Global land cover classification at 1 km resolution
227 293. using a decision tree classifier. International Journal of Remote
Dunne T. and Black R.D. (1970) Partial area contributions Sensing, 21, 1331 1365.
to storm runoff in a small New England watershed. Water Hessel R., Jetten V. and Guanghui Z. (2003) Estimating
Resources Research, 6(5), 1296 1311. Mannings n for steep slopes. Catena, 54(1 2), 77 91.
Dunne T., Moore T.R. and Taylor C.H. (1975) Recognition Hirsch R.M., Helsel D.R., Cohn T.A. and Gilroy E.J. (1992)
and prediction of runoff-producing zones in humid regions. Statistical analysis of hydrological data. In Handbook of
Hydrological Sciences Bulletin, 20, 305 327. Hydrology, Maidment R. (Ed.), McGraw-Hill: New York, pp.
Engeland K. and Gottschalk L. (2002) Bayesian estimation of 17.1 17.55.
parameters in a regional hydrological model. Hydrological Hoosbeek M.R. and Bryant R.B. (1993) Towards the quantitative
Earth System Sciences, 6(5), 883 898. modeling of pedogenesis: a review. Geoderma, 55, 183 210.
Engman E.T. (1986) Roughness coefficients for routing surface Houser P.R., Goodrich D. and Syed K. (2000) Runoff,
runoff. Journal of Irrigation and Drainage Engineering, 112, precipitation, and soil moisture at Walnut Gulch. In Spatial
39 53. Patterns in Catchment Hydrology: Observations and Modelling,
Farmer D., Sivapalan M. and Jothityangkoon C. (2003) Climate, Chap. 6, Grayson R. and Bloschl G. (Eds.), Cambridge
soil and vegetation controls upon the variability of water University Press: Cambridge, pp. 125 157.
balance in temperate and semiarid landscapes: downward Institute of Hydrology (IH) (1999) Flood Estimation Handbook.
approach to water balance analysis. Water Resources Research, Institute of Hydrology: Wallingford.
39(2), 1035, SWC 1 21. Jamagne M., Deroussin J., Eimberck M., King D., Lambert
Fernandez W., Vogel R.M. and Sankarasubramanian A. (2000) J.J., Lebas C. and Montanarella L. (2002) Soil geographical
Regional calibration of a watershed model. Hydrological data base of Eurasia and Mediterranean countries at 1:1m,
Sciences Journal, 45(5), 689 708. Transcations of the 17th World Congress of Soil Science, The
Flugel W.A. (1995) Delineating hydrological response units International Union of Soil Sciences, Symposium 44 , Bangkok,
by geographical information system analyses for regional paper no 494, 14 20 August.
hydrological modelling using PRMS/MMS in the drainage Jensen J.R. (1996) Introductory Digital Image Processing. A
basin of the River Brol, Germany. Hydrological Processes, 9, Remote Sensing Perspective, Second Edition, Prentice-Hall.
423 436. Kirnbauer R., Bloschl G., Haas P., Muller G. and Merz B.
Franks S.W., Gineste P., Beven K.J. and Merot P.h (1998) (2005) Identifying space-time patterns of runoff generation A
On constraining the predictions of a distributed model: the case study from the Lohnersbach catchment, Austrian Alps. In
incorporation of fuzzy estimates of saturated areas into the Global Change and Mountain Regions, Huber U., Reasoner M.
calibration process. Water Resources Research, 34(4), 787 797. and Bugmann H. (Eds.), Springer Wien: New York, pp.
Gallo K., Tarpley D., Mitchell K., Csiszar I., Owen T. and Reed B. 309 320.
(2001) Monthly fractional green vegetation cover associated Kite G. and Droogers P. (Eds.) (2000) Comparing actual
with land cover classes of the conterminous USA. Geophysical evapotranspiration from satellite data, hydrological models
Research Letters, 28(10), 2089 2092. and field data. Special Issue Journal of Hydrology, 229(1 2),
Gessler P.E., Moore I.D., McKenzie N.J. and Ryan P.J. (1995) 1 100.
Soil landscape modelling and spatial prediction of soil Koren V., Smith M. and Duan Q. (2003) Use of a priori parameter
attributes. International Journal of Geographical Information estimates in the derivation of spatially consistent parameters
Systems, 9(4), 421 432. sets of rainfall-runoff modes. In Calibration of Watershed
Gorte B.G.H. (2000) Land-use and catchment characteristics. In Models. Water Science and Application, Vol. 6, American
Remote Sensing in Hydrology and Water Management, Schultz Geophysical Union: pp. 239 254.
G.A. and Engman E.T. (Eds.), Springer: Berlin, pp. 133 156. Laaha G. and Bloschl G. (2005a) A comparison of low flow
Gottschalk L. (2002) Advances in observational hydrology regionalisation methods catchment grouping. Journal of
field experiments and modelling. In Proceedings of Workshop Hydrology, in press.
on the Prediction of Ungauged Basins (PUBs) held at 28 29 Laaha G. and Bloschl G. (2005b) Low flow estimates from short
March 2002 at the Yamanashi University, Takeuchi K. (Ed.), stream flow records a comparison of methods. Journal of
International Association of Hydrological Sciences: Kofu. Hydrology, 306(1 4), 264 286.
18 RAINFALL-RUNOFF MODELING

Lamb R., Beven K.J. and Myrab S. (1998) Use of spatially Parajka J., Merz R. and Bloschl G. (2005a) A comparison
distributed water table observations to constrain uncertainty in of regionalisation methods for catchment model parameters.
a rainfall-runoff model. Advances in Water Resources, 22(4), Hydrology and Earth Systems Sciences Discuss., 2, 509 542.
305 317. www.copernicus.org/EGU/hess/hessd/2/509/ SRef-ID:
Lamb R. and Kay A.L. (2004) Confidence intervals for a spatially 1812-2116/hessd/2005-2-509.
generalized, continuous simulation flood frequency model Parajka J., Merz R. and Bloschl G. (2005b) Regionale
for Great Britain. Water Resources Research, 40, W07501, Wasserbilanzkomponenten fur Osterreich auf Tagesbasis
doi:10.1029/2003WR002428. (Regional water balance components in Austria on a daily
Larsen J.E., Sivapalan M., Coles N.A. and Linnet P.E. basis). Osterreichische Wasser- und Abfallwirtschaft, 57(3/4),
(1994) Similarity analysis of runoff generation processes 43 56.
in real-world catchments. Water Resources Research, 30(6), Peschke G., Etzenberg C., Topfer J., Zimmermann S. and
1641 1652. Muller G. (1999a) Runoff generation regionalization: analysis
Leavesley G.H., and Stannard L.G. (1995) The Precipitation- and a possible approach to a solution. In Regionalization
Runoff Modeling System PRMS. In Computer Models in Hydrology, Diekkruger B., Kirkby M.J. and Schroder U.
of Watershed Hydrology, Singh V.P. (Ed.), Water Resources (Eds.), IAHS Publication No. 254, IAHS Press: pp.
Publications, Highlands Ranch, pp. 281 310. 147 156.
Lucas Y. (2001) The role of plants in controlling rates and Peschke G., Etzenberg C., Muller G., Topfer J. und
products of weathering. Annual Review of Earth and Planetary Zimmermann S. (1999b) Das wissensbasierte System FLAB
Sciences, 29, 135 163. ein Instrument zur rechnergestutzten Bestimmung von Land-
McCarthy J.M., Gumbricht Th.R., McCarthy T.S., Frost Ph.E., schaftseinheiten mit gleicher Abflubildung, IHI-Schriften,
Wessels K. and Seidel F. (2003) Flooding patterns of the H.10, Internat. Hochschulinstitut Zittau.
Okavango Wetland in Botswana between 1972 and 2000. Rinaldo A., Vogel G.K., Rigon R. and Rodriguez-Iturbe I. (1995)
AMBIO: A Journal of the Human Environment, 32(7), 453 457. Can one gauge the shape of a basin? Water Resources Research,
Merz R. and Bloschl G. (2004) Regionalisation of catchment 31, 1119 1128.
model parameters. Journal of Hydrology, 287, 95 123. Robinson J.S. and Sivapalan M. (1995) Catchment-scale model
Merz R. and Bloschl G. (2005) Flood frequency regionalisation of runoff generation by aggregation and similarity analysis.
spatial proximity vs. catchment attributes. Journal of Hydrological Processes, 9(5 6), 555 574.
Hydrology, 302(1 4), 283 306. Rodrguez-Iturbe I. and Valdes J.B. (1979) The geomorphologic
Merz R., Bloschl G. and Parajka J. (2005) Raum-zeitliche structure of hydrologic response. Water Resources Research, 15,
Variabilitat von Ereignisabflussbeiwerten in Osterreich (Spatio- 1409 1420.
temporal variability of event runoff coefficients in Austria). Scherrer S. and Naef F. (2003) The identification of dominant
Hydrologie und Wasserbewirtschaftung, in press. runoff processes on plot-scale based on field information used
Merz R., Parajka J. and Bloschl G. (2004) Raumliche to define hydrologic response units on the catchment-scale.
Muster in der konzeptionellen Wasserbilanzmodellierung Geophysical Research Abstracts, 5, 14 393.
Parameteridentifikation. Proceedings Tag der Hydrologie in Schumann A.H., Funke R. and Schultz G.A. (2000) Applications
Potsdam, Universitat Potsdam. of a Geographic Information System for conceptual rainfall-
Merz R., Piock-Ellena U., Bloschl G. and Gutknecht D. runoff modeling. Journal of Hydrology, 240(1,2), 45 61.
(1999) Seasonality of flood processes in Austria. In Schuurmans J.M., Troch P.A., Veldhuizen A.A., Bastiaanssen
Hydrological Extremes: Understanding, Predicting, Mitigating, W.G.M. and Bierkens M.F.P. (2003) Assimilation of remotely
Gottschalk L., Olivry J.C., Reed D. and Rosbjerg D. (Eds.), sensed latent heat flux in a distributed hydrological model.
Proceedings of the Birmingham Symposium, IAHS Publication Advances in Water Resources, 26, 151 159.
No. 255, IAHS: pp. 273 278, July 1990. SCS Soil Conservation Service (1973) A Method for Estimating
Mishra S.K. and Singh V.P. (2003) Soil Conservation Service Volume and Rate of Runoff in Small Watersheds, Technical paper
Curve Number (SCS-CN) Methodology, Book Series: Water 149, U.S. Department of Agriculture: Washington.
Science And Technology Library, Volume 42 , Kluwer Academic Sefton C.E.M. and Howarth S.M. (1998) Relationships between
Publishers. dynamic response characteristics and physical descriptors of
Moore I.D., Grayson R.B. and Ladson A.R. (1991) Digital catchments in England and Wales. Journal of Hydrology, 211,
terrain modelling: a review of hydrological, geomorphological, 1 16.
and biological applications. Hydrological Processes, 5, Seibert J. (1999) Regionalisation of parameters for a conceptual
3 30. rainfall-runoff model. Agricultural and Forest Meteorology,
Mroczkowski M., Raper G.P. and Kuczera G. (1997) The quest 98 99, 279 293.
for more powerful validation of conceptual catchment models. Seibert J. and McDonnell J.J. (2002) On the dialog between
Water Resources Research, 33(10), 2325 2335. experimentalist and modeler in catchment hydrology: use of
Nash I.E. and Sutcliffe I.V. (1970) River flow forecasting through soft data for multicriteria model calibration. Water Resources
conceptual models, part I. Journal of Hydrology, 10, 282 290. Research, 38(11), 1241, doi:10.1029/2001WR000978.
Nathan R.J. and McMahon T.A. (1990) Identification of Sepaskhah A.R. and Bondar H. (2002) Estimation of manning
homogeneous regions for the purpose of regionalization. roughness coefficient for bare and vegetated furrow irrigation.
Journal of Hydrology, 121, 217 238. Biosystems Engineering, 82(3), 351 357.
RAINFALL-RUNOFF MODELING OF UNGAUGED CATCHMENTS 19

Sivapalan M., Beven K.J. and Wood E.F. (1987) On hydrologic Walker J.P., Willgoose G.R. and Kalma J.D. (2001) One-
similarity. 2. A scaled model of storm runoff production. Water dimensional soil moisture profile retrieval by assimilation of
Resources Research, 23(12), 2266 2278. near-surface observations: a comparison of retrieval algorithms.
Stewart J.B., Engman E.T., Feddes R.A. and Kerr Y. (1996) Advances in Water Resources, 24(6), 631 650.
Scaling up in Hydrology Using Remote Sensing, John Wiley: Warrick A.W., Zhang R., Moody M.M. and Myers D.E. (1990)
p. 255. Kriging versus alternative interpolators: errors and sensitivity
Szolgay J., Hlavkova K., Kohnova S. and Kubes R. (2002) to model inputs. In Field-scale Water and Solute Flux in Soils,
Regional calibration of a water balance model for estimating Roth K. etal (Eds.), Birkhauser Verlag: Basel, pp. 157 164.
mean monthly flow in small ungauged catchment, ERB Western A.W. and Bloschl G. (1999) On the spatial scaling of
and Northern European FRIEND Project 5 Conference, soil moisture. Journal of Hydrology, 217, 203 224.
Demanovska dolina. Western A.W., Grayson R.B., Bloschl G., Willgoose G.R. and
Tucker C.J., Grant D.M. and Dykstra J.D. (2004) NASAs global McMahon T.A. (1999) Observed spatial organisation of soil
orthorectified landsat data set. Photogrammetric Engineering moisture and its relation to terrain indices. Water Resources
and Remote Sensing, 70(3), 313 322. Research, 35(3), 797 810.
United States Department of Agriculture (USDA) (1991) State Western A.W., Grayson R.B. and Bloschl G. (2002) Scaling of
Soil Geographic Data Base (STATSGO) Data Users Guide, soil moisture a hydrological perspective. Annual Review of
Miscellaneous Publication No. 1492, USDA-Soil Conservation Earth and Planetary Sciences, 30, 149 180.
Service: Washington, p. 88. Western A.W., Zhou S.-L., Grayson R.B., McMahon T.A.,
USACE (1994) Engineering and Design Flood-Runoff Analysis, Bloschl G. and Wilson D.J. (2004) Spatial correlation of soil
Publication Number EM 1110 2-1417, U.S. Army Corps of moisture in small catchments and its relationship to dominant
Engineers: Washington. spatial hydrological processes. Journal of Hydrology, 286(1 4),
Vieux B.E. (2001) Distributed Hydrologic Modeling Using GIS: 113 134.
Water Science And Technology Library, Vol. 38, Kluwer WMO (1995) INFOHYDRO Manual, Second Edition, Operational
Academic Publishers. Hydrology Report, No. 28, WMO Report No. 683, World
Vogel R.M. and Kroll C.N. (1991) The value of streamflow record Meteorological Organisation, Geneva.
augmentation procedures in low-flow and flood-flow frequency Woods R.A. (2003) The role of catchment classification in
analysis. Journal of Hydrology, 125(3 4), 259 276. Hydrology science. Geophysical Research Abstracts, Vol. 5,
Wagner W., Scipal K., Pathe C., Gerten D., Lucht W. and 08 192, European Geophysical Society.
Rudolf B. (2003) Evaluation of the agreement between the Wosten J.H.M., Pachepsky Y.A. and Rawls W.J. (2001)
first global remotely sensed soil moisture data with model Pedotransfer functions: bridging the gap between available
and precipitation data. Journal of Geophysical Research- basic soil data and missing soil hydraulic characteristics.
Atmospheres, 108(D19), 4611, doi: 10.1029/2003JD003663. Journal of Hydrology, 251(3 4), 123 150.

You might also like