Todini1996 ARNO PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 44

Journal

of
Hydrology
ELSEVIER Journal of Hydrology 175 (1996) 339-382

The ARNO rainfall-runoff model


E. Todini
Institute for Hydraulic Construction, University of Bologna, Via/e Risorgimento 2,40136 Bologna, Italy
Received 2 March 1995; accepted 30 March 1995

Abstract

This paper describes in detail a semi-distributed conceptual rainfall-runoff model known as


the ARNO model, which is now in widespread use both in land-surface-atmosphere process
research and as an operational flood forecasting tool on several catchments in different parts of
the world. The model, which derives its name from its first application to the Arno River,
incorporates the concepts of a spatial probability distribution of soil moisture capacity and of
dynamically varying saturated contributing areas.
The ARNO model is characterized by two main components: the tirst and most important
component represents the soil moisture balance, and the second describes the transfer of runoff
to the outlet of the basin. The relevance of the soil component emerges from the highly non-
linear mechanism with which the soil moisture content and its distribution controls the
dynamically varying size of the saturated areas mainly responsible for a direct conversion of
rainfall into runoff. The second component describes the way in which runoff is routed and
transferred along the hillslopes to the drainage channels and along the channel network to the
outlet of the basin. Additional components, such as the evapotranspiration, snowmelt and
groundwater modules, are also described. A discussion on the advantages of the model, cali-
bration requirements and techniques is also presented, together with the physical interpretation
of model parameters.
Finally, after describing the original calibration of the ARNO model on the Arno basin, and
a comparison with several conceptual models, recent applications of the ARNO model, as part
of a real-time flood forecasting system, as a tool for investigating land use changes and as an
interesting approach to the evaluation of land-surface-atmosphere interactions at general
circulation model (GCM) scale, are illustrated.

1. Introduction

The literature contains many works that summarize the level of understanding of
the complex physics governing the transformation of rainfall into runoff (Dunne,
0022-1694/96/$15.00 0 1996 - Elsevier Science B.V. All rights reserved
SSDI 0022-1694(95)02853-6
340 E. Todini / Journal of Hydrology 175 (1996) 339-382

1978; Freeze, 1980). Many efforts have been made to schematize the whole process so
as to develop mathematical models (Dooge, 1957, 1973; Amorocho and Hart, 1964;
Freeze and Harlan, 1969; Todini, 1989). These range from the simple lumped
calculation of design discharge to the distributed representation of the various
processes based on the conservation of mass, energy and momentum (Bathurst,
1986; Abbott et al., 1986a,b; Beven et al., 1987; Binley et al., 1989). Taken together,
such models form the broad category of distributed differential models (Todini,
1986); they are frequently referred to as ‘physically based models’ to highlight the
fact that their respective parameters are (or should be) reflected in the field
measurements (Beven, 1989). Given their nature, such models are appropriate for
studying the effects of land use changes, soil erosion, surface groundwater inter-
actions, etc., but are less suitable for conventional rainfall-runoff applications at
catchment scale.
Another category of models, which was developed mainly for operational
purposes, is that of the complex ‘conceptual models’. From the early 1960s a large
number of these models were described, from the Stanford Watershed Model IV
(Crawford and Linsley, 1966) to the SSARR (Rockwood et al., 1972), the Sacramento
(Burnash et al., 1973) and the Tank (Sugawara et al., 1976), which represented in
different ways the response mechanisms of the various phenomena, but mostly by
means of non-linear reservoirs and thresholds, directly connected or linked either by
means of linear transfer functions or by linear or non-linear hydrologic or hydraulic
type routing.
The underlying reason for developing these models was to represent the hydrologic
cycle by linking together process components which described physical concepts, on
the presumption that the model parameters would also bear physical meaning, so that
they could be assigned values without reference to the observed data. In other words,
it was assumed that most of their parameters (such as storage coefficients, roughness
coefficients or thresholds present in the various sub-components) could be defined
from the physiographic characteristics of the basins. In reality, the parameters needed
to be estimated by minimizing objective functions *(e.g. the sum of squared
deviations), which generally led to groups of unrealistic parameters incorporating
both data measurement errors and the errors presemin the structure of the model
itself; in addition, parameter observability conditions could not always be guaranteed
(Sorooshian and Gupta, 1983).
Although they are in widespread use throughout the world, it is now understood
that the basic failure of these models to represent catchment response with a small
number of parameters is essentially due to their inability to reproduce the dynamic
variation of the saturated areas within the catchment (Beven et al., 1983). Indeed, in
recent years, a general consensus has been reached on the fact that it is this dynamic
variation, a function of the accumulation and horizontal movement of water in the
upper soil layers (see Todini (1995) and Franchini et al. (1996)), which is mainly
responsible for the highly non-linear nature of catchment response to storm ,events.
Most of the conceptual modellers tried to compensate for the inadequacy of their
models by adding more and more process components as well as parameters, but they
failed to reproduce the actual phenomena and reduced the models to extremely
E. Todini / Journal of Hydrology I75 (1996) 339-382 341

complex black boxes with an exceedingly high number of parameters (frequently


larger than 20) to be estimated from the available records. This emerged clearly
from the WMO intercomparison of conceptual models (World Meteorological
Organization (WMO), 1975) where the results of all the different models did not
appear to be significantly better than those produced by the constrained linear systems
(CLS) model, a simple piecewise linear black-box model (Natale and Todini, 1977).
More recently, newly developed conceptual models have regarded the soil moisture
replenishment, depletion and redistribution mechanism as directly responsible for the
dynamic variation of the areas contributing to direct runoff. From this concept, a
number of models originated which use a probability distribution of the soil moisture
content, as described by Zhao (1977) and Moore and Clarke (1981) or the distribu-
tion of a topographic index, as in TOPMODEL (Beven and Kirkby, 1979; Beven et
al., 1984). The advantage of these models lies in their capability to reproduce catch-
ment response with a smaller number of physically meaningful parameters than the
more traditional lumped conceptual models (see Franchini and Pacciani, 1991). The
ARNO model falls in this class of models, as it derives its soil moisture accounting
module directly from the distribution function approach of Zhao (1977) augmented
by the introduction of drainage and percolation losses in the soil moisture balance.
The major advantage of the ARNO model is the fact that it is entirely driven by the
total catchment soil moisture storage, which is functionally related, by means of
simple analytical expressions, to the dynamic contributing areas, and to the drainage
and the percolation amounts.
The soil moisture module of the ARNO model has been extensively used in hydro-
logical practice; in particular, it has become the kernel of a real-time operational flood
forecasting system developed on behalf of the Commission of the European Com-
munities (the European Flood Forecasting Operational Real-Time System
(EFFORTS); ET&P, 1992) which is already operational on several rivers in many
countries: the Fuchun in China; the Danube in Germany; the PO, the Arno, the Tiber,
the Adda and the Oglio in Italy. Recently, it has been tested in joint meteorological-
hydrological experiments, such as the Spatial Variability of Land Surface Processes
(SLAPS) project (Dooge et al., 1994) and by a number of meteorologists (Rowntree
and Lean, 1994; Polcher et al., 1995) and, given the simplicity of its formulation, it
was successfully included in the Hamburg climate model (Dtimenil and Todini, 1992).
In addition to the crucial soil moisture component, the ARNO model includes
evapotranspiration, snowmelt and groundwater components, all of which are
parsimonious descriptions in line with the overall philosophy. Moreover, once a
satisfactory description of the so-called ‘runoff production function’ has been
achieved, the runoff contributions of the area units considered must then be trans-
ferred downstream, and aggregated as they move along the slopes and the drainage
network.
A wide choice of runoff transfer or routing methods is available, and one may find
the folllowing in the literature: numerical integration of the non-linear parabolic flow
equations as in SHE (Abbott et al., 1986b); numerical integration of the non-linear
kinematic flow equation (Kibler and Woolhiser, 1970); methods based on the
Muskingum-Cunge algorithm (Cunge, 1969; Zhao, 1977); methods based on a
342 E. Todini / Journal of Hydrology 175 (1996) 339-382

gamma distribution (Nash, 1958; Kalinin and Miljukov, 1958); methods based on the
derivation of single transfer functions (Nalbantis et al., 1988a,b) or multiple transfer
functions (Natale and Todini, 1977); derivation of the geomorphological unit hydro-
graph (GIUH) (Rodriguez Iturbe and Valdes, 1979; Rodriguez Iturbe et al., 1982;
Rosso, 1983, 1984); methods based on a linear parabolic representation leading to an
inverse Gaussian distribution (Todini, 1988a; Franchini and Pacciani, 1991; Naden,
1992).
However, backed by numerous experimental results (e.g. Franchini and Pacciani,
1991) and by various authoritative opinions (Cordoba and Rodriguez-Iturbe, 1983),
it may safely be stated that the routing component, which is essential for linking
together the various area units, is not the real kernel of the problem, and,
whatever the representation used, the results will not be very different, except
perhaps for more or less physical significance to be attributed to the parameters.
In the ARNO model, the linear parabolic approach has been successfully used
with parameter values that can be established according to physical reasoning,
without the need for extensive trial and error or highly non-linear optimization
procedures.

2. Basic concepts in the ARNO model

The ARNO model is a semi-distributed conceptual model in integral form (Todini,


1988b), which is based on the schematic representation of catchment hydrology
shown in Fig. 1. The catchment is divided into a series of sub-basins to each of
which the rainfall-runoff model is applied; this division takes place according to

LATERAL OR
lKFLOWS

IJTFIDW DLSCHARGIZS

Fig. 1. Schematic representation of processes within a catchment unit.


E. Todini / Journal of Hydrology 175 (1996) 339-382

Fig. 2. Example of a catchment sub-division.

the natural sub-basin boundaries so that the sub-basin closing sections coincide with
the cross-sections of interest along the river and its tributaries (Fig. 2). These sections
are chosen according to the presence of hydrometric measuring stations, either
because they are of interest for flood forecasting or for reasons connected with the
morphology of the basin. The sub-catchments can then be represented as a tree (Fig.
3) which defines the order in which the computations are to be carried out.
The following operations are performed on each sub-basin for each time interval:
calculation of the evapotranspiration and runoff production for the ith generic sub-
basin; transfer of runoff, by means of a simplified hydraulic routing model, along the
slopes and in the channel reach inside the ith sub-basin to the closing section; transfer
of the input hydrograph from the upstream closing section to that of the sub-basin

w G F

H
E

Fig. 3. Schematic tree representing the catchment of Fig. 2.


344 E. Todini / Journal of Hydrology I75 (1996) 339-382

immediately downstream using the same form of routing model; summation at the
downstream closing section of the input hydrograph transferred from the upstream
sub-basin with the hydrograph produced by the ith sub-basin.
The main physical phenomena represented in the ARNO model are the following
(Fig. 1): water balance in the soil, on the basis of present soil moisture content,
rainfall, runoff, evapotranspiration, drainage and percolation; water losses through
evapotranspiration, evaluated on the basis of the air temperature data and the soil
moisture content; snow accumulation and/or melt, based upon the energy balance of

p(t) precipitation
e(t) evapo-transpiration
r(t) runoff
QMi inflow discharges

Qv outflow discharges
N snowmelt
runoff
:: interflow
I percolation
B baseflow

Fig. 4. Processes and quantities represented in the ARNO model.


E. Todini / Journal of Hydrology 175 (1996) 339-382 345

the snow cover (when present) as a function of the air temperature and precipitation
data; groundwater flow represented by means of a multiple linear reservoir-type
model; overland and channel flow routing represented with linear parabolic models.
Accordingly, as mentioned above, several modules have been developed to
represent the individual aspects of the overall rainfall-runoff process, as follows:
soil moisture balance module; evapotranspiration module; snowmelt module;
groundwater module; parabolic transfer module. Fig. 4 shows, for each sub-
catchment, a schematic flow diagram connecting the different modules considered
in the ARNO model.
Geomorphological data such as average catchment elevation, catchment surface
area and length of stream, as well as parameters including celerity and diffusivity,
thermal gradient, and parameters characterizing the spatial distribution of soil
moisture storage must be provided to apply the model to a catchment. Rainfall
and air temperature inputs are provided to the model as area averages by means of
weights generally based upon Thiessen polygons. In the case of temperature, the
calculation of the average temperature over a sub-basin takes into account the effects
of a thermal gradient with elevation.
Once the runoff is obtained from the soil moisture balance, the routing of runoff on
the hillslopes of the catchment is simulated by applying a distributed inflow linear

Schematic representation of channel routing


with distributed lateral inflow

Fig. 5. Hillslope and channel routing schemes, where q(t) is the lateral distributed inflow discharge, Q(r) is
the outflow discharge and r(t) is the distributed runoff inflow.
346 E. Todini / Journal of Hydrology 175 (1996) 339-382

parabolic model to an ‘open book’ representation of the hillslope elements (Fig. 5).
The channel routing is also performed by means of a distributed input linear
parabolic model, and the contribution from upstream sub-catchments is routed
downstream by means of a linear concentrated-input parabolic model (Fig. 5). The
sub-catchment outflows are then routed downstream according to the tree scheme
(Fig. 3) which represents the overall catchment.

3. The soil moisture balance module

The soil moisture balance module of the ARNO model derives from the Xinanjiang
model developed by Zhao (1977, 1984), who expressed the spatial distribution of the
soil moisture capacity in the form of a probability distribution function, similar to
that advocated by Moore and Clarke (1981) and Moore (1985). Successively, to
account more effectively for soil depletion owing to drainage, the original Xinanjiang
model scheme was modified by Todini (1988a), who originated the ARNO model
within the frame of the hydrological study of the river Arno.
The basic assumptions expressed in the soil moisture balance module of the ARNO
model are as follows: (1) the precipitation input to the soil is considered uniform over
the catchment (or sub-catchment) area; (2) the catchment is composed of an infinite
number of elementary areas (each with a different soil moisture capacity and a
different soil moisture content), for each of which the continuity of mass can be
written and simulated over time; (3) all the precipitation falling over the soil infiltrates
unless the soil is either impervious or it has already reached saturation; (4) the
proportion of elementary areas which are saturated is described by a spatial
distribution function; (5) the spatial distribution function describes the dynamics of
contributing areas which generate surface runoff, (6) the total runoff is the spatial
integral of the infinitesimal contributions deriving from the different elementary
areas; (7) the soil moisture storage is depleted by the evapotranspiration as well as
by lateral sub-surface flow (drainage) towards the drainage network and the percola-
tion to deeper layers; (8) both drainage and percolation are expressed by simple
empirical expressions.
A sub-basin of given surface area ST (excluding the surface extent of water bodies
such as reservoirs or lakes) is in general formed by a mixture of pervious and less
pervious terrains, the response to precipitation of which will be substantially different.
For this reason, the total area ST is divided into the impervious area S, and the
pervious area SP:
ST = s, + sp (1)
To derive the expressions needed for the continuous updating of the soil moisture
balance, let us first deal with the amount of precipitation that falls over the pervious
area. Given that, from Eq. (1) the entire pervious area is
s, = S* - s, (2)
E. Todini / Journal of Hydrology 175 (1996) 339-382 341

if one denotes by (S - S,) the generic surface area at saturation, then x, defined as
s - s,
(3)
x=sT.
will indicate the percentage of pervious area at saturation. Zhao (1977) demonstrated
that the following relation holds reasonably well between the area at saturation
and the local proportion of maximum soil moisture content W/W,, where w is the
elementary area soil moisture at saturation and w, is the maximum possible soil
moisture in any elementary area of the catchment:
b
(4)

This is similar to defining the cumulative distribution for the elementary area soil
moisture at saturation, shown in Fig. 6 and defined as
w = Wm[l- (1 -X)$ (5)
In the ARNO model, an interception component (Rutter et al., 1971, 1975) is not
explicitly included. Nevertheless, to allow for a substantially larger evapotranspira-
tion when the canopies are wet (without obviously explicitly accounting for the
disappearance of the stomata1 resistance (Shuttleworth, 1979)), the following succes-
sion of operations is followed. If the precipitation P is larger than the potential
evapotranspiration ETp, the actual evapotranspiration, for the reasons expressed
above, is assumed to coincide with the potential, i.e.
ET, = ETP (6)
and so an ‘effective’ meteorological input M,, defined as the difference between
precipitation and potential evapotranspiration, becomes
M,=P-ET,=P-ET,>0 (7)

Fig. 6. Cumulative distribution for the elementary area soil moisture at saturation.
348 E. Todini / Journal of Hydrology 175 (1996) 339-382

With reference to Fig. 7, the surface runoff R generated by the entire catchment is
obtained as the sum of two terms; the first one is the product of the meteorological
effective input and the percentage of impervious area, and the second one is the
average runoff produced by the pervious area, which is obtained by integrating the
soil moisture capacity curve, which gives
M,+W
sT - SI
R=zM,+- 40 d5 if M,+w< w, (8)
ST
w

or

1 if M,+waw,

Eqs. (8) and (9) can also be expressed in terms of the catchment average soil moisture
(9)

content ( W)and that at saturation ( W,), and after integration they become

1I
bfl
R=M,+
- (b +: w,

forO<M,<(b+l)W,,,

R=M,+ V(W,- W) forM,Z(b+l)w,,,

Fig. 7. Runoff R generated by an effective meteorological input M, > 0.


E. Todini / Journal of Hydrology 175 (19%) 339-382 349

If the precipitation P is smaller than the potential evapotranspiration ET,, the


effective meteorological input it4, becomes negative:

M,=P-ET,<0 (11)

which implies that the runoff R is zero. The actual evapotranspiration is then
computed as the precipitation P plus a quantity which depends upon A& reduced
by the average degree of saturation of the soil (see Appendix A), which gives

(1+Ek)-(1-E)”
ET, = P + (ETp - P) (STs; ‘I) (12)
(1+$ - (l-E)&

These equations, which represent the average surface runoff produced in the sub-
catchment, must be associated with an equation of state to update the mean water
content in the soil. This equation takes the form

W(t + At) = W(t) + P(t, t + At) - ET,(t, t + At) - R(t, t + At)


- D(t, t + At) - Z(t, t + At) (13)

where P(t, t + At) is the area precipitation between t and t + At; ET,(t, t + At) is the
loss through evapotranspiration between t and t + At; R(t, t + At) is the total runoff
between t and t + At; D(t, t + At) is the loss through drainage between t and t + At;
Z(t, t + At) is the percolation loss to groundwater between t and t + At; W(t,+ At) is
the soil moisture content at time t + At; W( t)‘is the soil moisture content at time t. All
the quantities representing averages over the sub-basin are expressed in millimetres.
The non-linear response of the unsaturated soil to precipitation, represented by the
shape of the distribution curve given by Eq. (5), is strongly affected by the horizontal
drainage and vertical percolation losses. The drainage loss D is an important quantity
to be reproduced in a hydrological model, because on the one hand it affects the soil
moisture storage and on the other hand it controls the hydrograph recession.
Experience derived from applications suggested the use of the following empirical
relationship to express also the drainage loss D as a non-linear function of the soil
moisture content:

D = D,,,$- for W < W,


m
(14)
D = Dmin5 + (Dmax-Dtin)(EIFJ for W>WW,
m

where c is the exponent of the variation law, Dti,, and D,,, are drainage parameters
at saturation and W, is .moisture content threshold value.
The percolation loss Z, which feeds the groundwater module, which will control the
base flow in the model, varies less significantly over time if compared with the other
350 E. Todini / Journal of Hydrology 175 (19%) 339-382

terms; nevertheless, a non-linear behaviour is also assumed as follows:


Z= 0 for W < Wi
(15)
Z=U(W- Wi) for Wa Wi
where Wi represents the moisture content threshold value below which the
percolation may be considered negligible, and a is an empirical coefficient.
The total runoff per unit area produced by the precipitation P is finally expressed as
Rtot = R i- D + B (16)
where B is the base flow generated by the presence of a groundwater table fed by the
percolation, and computed by the groundwater module.

4. The evapotranspiration module

Although the Penman-Monteith equation is the most rigorous theoretical descrip-


tion for this component, in practice many simplifications are necessary because in
most countries the required historical data for its estimation are not extensively
available, and, in addition, apart from a few meteorological stations, almost nowhere
are real-time data available for flood forecasting applications. Moreover, it should be
clearly understood that evapotranspiration plays a major role in the rainfall-runoff
process not in terms of its instantaneous impact, but in terms of its cumulative
temporal effect on the soil moisture volume depletion; this reduces the need for an
extremely accurate expression, provided that its integral effect be well preserved. In
the ARNO model, the effects of the vapour pressure and wind speed are explicitly
ignored and evapotranspiration is calculated starting from a simplified equation
known as the radiation method (Doorembos et al., 1984):

ETod = C, W,R, = C, W, (0.25 + 0.50 ;;> R, (17)

where ETod is the reference evapotranspiration, i.e. evapotranspiration in soil satura-


tion conditions caused by a reference crop (mm day-‘); C, is an adjustment factor
obtainable from tables as a function of the mean wind speed; W,, is a compensation
factor that depends on the temperature and altitude; R, is the short-wave radiation
measured or expressed as a function of R, in equivalent evaporation (mm day-‘); R,
is the extraterrestrial radiation expressed in equivalent evaporation (mm day-‘); n/N
is the ratio of actual hours of sunshine to maximum hours of sunshine (values
measured or estimated from mean monthly values as described in Appendix B).
Hence the calculation of R, requires both knowledge of R,, obtainable from tables
as a function of latitude, and knowledge of actual n/N values, which may not be
available. In the absence of the measured short-wave radiation values R, or of the
actual number of sunshine hours otherwise needed to calculate R, as a function of R,
(see Eq. (17)), an empirical equation was developed that relates the reference potential
evapotranspiration ETod,computed on a monthly basis using one of the available
simplified expressions, to the compensation factor W,, the mean recorded
E. Todini 1 Journal of Hydrology 175 (1996) 339-382 351

temperature of the month T and the maximum number of hours of sunshine N. The
developed relationship is linear in temperature (and hence additive), and permits the
dissaggregation of the monthly results on a daily or even on an hourly basis, whereas
most other empirical equations are ill-suited for time intervals shorter than 1 month.
The relation used, which is structurally similar to the radiation method formula in
which the air temperature is taken as an index of radiation, is
ET0 = cx + PNW,,T,,, (18)
where ET0 is the reference evapotranspiration for a specified time step At (in mm
(At)-‘); cxand ,L3are regression coefficients to be estimated for each sub-basin; T,,, is
the area mean air temperature averaged over At; N is the monthly mean of the
maximum number of daily hours of sunshine (tabulated as a function of latitude).
W,, for a given sub-basin can be either obtained from tables or approximated by a
fitted parabola:

W,,=AT2+BT+C (19)
where A, B and C are coefficients to be estimated; T is the long-term mean
monthly sub-basin temperature (“C). Further details on the estimation of the
evapotranspration parameters are given in Appendix B.

5. The snowmelt module

Again, for reasons of limited data availability, the snowmelt module is driven by a
radiation estimate based upon the air temperature measurements; in practice, the
inputs to the module are the precipitation, the temperature, and the same radiation
approximation which is used in the evapotranspiration module.
Given the role that altitude may play in combination with the thermal gradient, the
sub-catchment area is subdivided into a number of equi-elevation zones (snow-bands)
according to the hypsometric curve, and for each zone simplified mass and energy
budgets are continuously updated. For each snowband the following steps, similar to
those adopted in SHE (Abbott et al., 1986a,b), are followed: (1) estimation of
radiation at the average elevation of the snowband; (2) decision on whether precipita-
tion is solid or liquid; (3) estimation of the water mass budget based on the hypothesis
of zero snowmelt; (4) estimation of the energy budget based on the hypothesis of zero
snowmelt; (5) comparison of the total available energy with that sustained as ice by
the total available mass at 273°K; (6) computation of the snowmelt produced by the
excess energy; (7) updating of the water mass budget; (8) updating of the energy
budget.

5.1. Estimation of radiation at the average elevation of the snowband

The estimation of the radiation is performed by re-converting the latent heat (which
has already been computed as the reference evapotranspiration ETo) back into
radiation, by means of a conversion factor C,, (kcal kg-‘), which can be found in
352 E. Todini 1 Journal of Hydrology 175 (1996) 339-382

any thermodynamics textbook as


C,, = 606.5 - 0.695(T - T,,) (20)
where TOis the temperature of fusion of snow (273°K).
In addition, to account for albedo, which plays an extremely important role in
snowmelt, it is necessary to apply an efficiency factor, which will be assumed approxi-
mately as r] = 0.6 for clear sky and n = 0.8 for overcast conditions; this leads to the
following estimate for the driving radiation term:

Rad = ~[606.5 - 0.695( T - TO)]ETO (21)


Given the lack of information concerning the status of the sky when simulating with
historical data, for practical purposes it is generally assumed in the ARNO model that
the sky is clear if there is no precipitation and overcast if precipitation is being
measured.

5.2. Decision on whether the precipitation is solid or liquid

Information concerning the status of precipitation (solid or liquid) is rarely avail-


able as a continuous record; therefore it is necesary to define a mechanism, mainly
based upon the air temperature measurements and the historical precipitation. If one
plots the frequency of the usually scattered observations with which precipitation was
observed to be liquid or snow as a function of the air temperature, a Gaussian
distribution is generally obtained, with a mean value T, which very seldom will
coincide with TO. For this reason, the following rules are adopted: precipitation is
taken as liquid if the air temperature T > T,; precipitation is taken as snow otherwise.
The value of T, (which generally ranges between 271 and 275°K) must be derived, as
mentioned above, by plotting the frequency of the status of historically recorded
precipitation as a function of the air temperature.

5.3. Estimating the water mass

The water equivalent mass is estimated with the following simple mass balance
equation, where all quantities are expressed in millimetres of water:

Zr;& = 2, + P (22)
The water equivalent at the end of the time step is designated with a star because it is a
tentative value which does not yet account for the eventual snowmelt.

5.4. Estimating the energy

Similarly to the mass, the energy is estimated in the following way, by computing
the increase (or decrease) of total energy E: if the precipitation is zero,

E&at = El + Rad (23)


E. Todini / Journal of Hydrology 175 (1996) 339-382 353

if the precipitation is non-zero and the precipitation is in the solid phase (T < T,),

E& = Et + Rad + CSgToP (24)


if the precipitation is non-zero and the precipitation is in the liquid phase (T > T,),
E*t+~t = El + Rad + [CsiTo + Cy + C,,(T - To)]f’ (25)
Again, the total energy at the end of a time step is designated with a star to
denote a tentative value; in the previous equations CSi is the specific heat of ice
(0.5 kcal “K-’ kg-‘), C, is the heat of fusion of water (79.6 kcal kg-‘) and C,, is the
specific heat of water (1 kcal “K-’ kg-‘).

5.5. Estimation of snowmelt and updating of mass and energy state variables

If the total available energy is smaller than or equal to that required to maintain the
total mass in the solid phase at the temperature To, i.e. CsiZ:+AtTo 2 ,!?:+A~,it means
that the available energy is not sufficient to melt part of the accumulated water, and
therefore
R, =0

&+At = z:+At (26)

E t+At - E,*,A~

where R, is the snowmelt expressed in millimetres. If the total available energy is


larger than that required to maintain the total mass in the solid phase at the tem-
perature To, it means that part of the accumulated water will melt, and therefore the
following energy balance equation holds:

Csi(Z,*,At - Km) To = G+A~ - (~J~To + C/f&n (27)


from which the snowmelt and the mass and energy state variables can be computed as
E’
&,, = t+At - C~iT&+A,
C/f

Z r+At - zt+At - &m (28)


E t+Ar - G+A~ - (CsiTo + Clf)&m

6. The groundwater module

The groundwater module represents the overall response of its storage by means of
a cascade of linear reservoirs, characterized by two parameters: the number of
reservoirs n and their time constant k, a model which is well known in hydrology
as the Nash model (Nash, 1958). For practical reasons, instead of using the gamma
354 E. Todini / Journal of Hydrology 175 (1996) 339-382

distribution function to express the impulse response, a numerical procedure has been
adopted. The expression to be used can easily be derived from the assumption of a
cascade of linear reservoirs where the volume of the ith reservoir is proportional to its
outflow, i.e. S, = kBf; thus, the continuity equation written for the generic ith
reservoir becomes

which, after discretization in time with a finite difference scheme centred at time
t + At/2, becomes
&i-, + Bf-1 Bi
;(BaAr_ Bf) = 2 _ ‘+A;+B’ (30)

and the required expression is then obtained by making B:+A~explicit:

2k - At Bi +
B’t+b =2k+At f & (B&A, + Bf’) (31)

where B’ is the outflow from the ith reservoir; B” = Z is the percolation given by the
soil moisture balance equation; B” = B is the resulting base flow. Eq. (31) is then
recursively solved n times at each time step.

7. The parabolic routing module

As outlined in Section 2, the hillslope routing and channel routing of distributed


inflows are both performed using a distributed inflow linear parabolic model, whereas
channel routing of upstream inflows to a sub-basin is performed by means of a
concentrated input parabolic model. The parameters of the two linear parabolic
transfer functions can be estimated as a function of the slope, the length of the
drainage system and the roughness.

7.1. Routing of upstream inflow

The propagation of inflows from upstream is carried out by means of a linear model
consisting of a parabolic unit hydrograph deriving from the analytical integration of
the unsteady flow equations when the inertia effects are ignored and the coefficients
are linearized around mean outflow values. In this case the following differential
equation is obtained:

ae=.a’e_,ae (32)
at ax2 ax
where D and C are the diffusivity and the convectivity coefficient, respectively. The
discrete time solution of Eq. (23) for mean values of the coefficients C and D is (Todini
E. Todini / JournaI of Hydrology 175 (1996) 339-382 355

and Bossi, 1986)

VA&, t + At) = -&[ZF(t + At) - 2ZF(t) + ZF(t - At)] (33)

where

IF(t) = jF(O)dO= jlr&)didS (34)


0 00

with uax, the impulse response relevant to Eq. (34), which can be written as

(35)

Integrating Eq. (34), after substitution from Eq. (35) one obtains

ZF(t)=i{(Ct-Ax)iV[-Giz] +(Ct+Ax)exp(y)

(36)

where N(*) is the value of the standard normal probability distribution in (*), and Ax
is the length of the channel reach.

7.2. D&se lateral inflow routing

The runoff generated in each sub-basin moves first along the slopes and then
reaches the drainage network as diffuse inflow, which, for simplicity, may be
considered uniformly distributed. In this case, the following differential equation
applies:

-- cdQ
__-= aQ
DaZQ
ax2 dx at --(A (37)

where q is the lateral inflow per unit length. The discrete time solution of Eq. (37) for
mean values of the coefficients C and D is (Franchini and Todini, 1989)

m&t, t + At) = -$[ZG(t + At) - 2ZG(t) + ZG(t - At)] (38)

where

ZG(t) = ; [t2 - ZZF(t)] (39)

with

ZZF(t) = jZF(E)d[= jjP(O)dt9d< = jjj~&)drdRdf (40)


0 00 000
356 E. Todini / Journal of Hydrology I75 (1996) 339-382

Integration of Eq. (40), after substitution of uAx from Eq. (35), yields

&&lZ) ++$C2$ezp(%F)]}

-+$)t$wo)]}
1 (41)

The results provided by Eq. (38) when substituted for IG(t) given by Eq. (41), differ
from those obtained by Naden (1992), in that they represent the response of a discrete
time system to a time discretized input (see Todini and Bossi, 1986).
As mentioned above, the parameters of each of the models are naturally composed
of the two convectivity (C)and diffusivity (D) coefficients for each response function
which are related to the dimensions, slopes and lengths of the individual sub-basins.
In general, very small values for D are assumed for the hillslopes, given the marked
kinematic nature of the phenomenon; generally, D increases with the size of the river
and with the inverse of the bottom slope. Table 1 gives an indication of the orders of
magnitude of C and D.

8. Model calibration requirements

To adapt the ARNO model to a specific basin, the following data are needed:
(1) an orographic map of the basin, at an appropriate ‘manageable’ size (1:25 OOO-
1:lOO000) depending on the actual size of the catchment showing the hydrographic
network.
(2) Continuous historical records of precipitation, temperature and river levels at
several measurement stations within the catchment, sampled at the appropriate time
steps (1 h for sub-catchment sizes of 200-300 km2 or 3 h for sub-catchments of the
order of 1000-2000 km2 as an upper limit). The length, of these records must be
sufficient to include periods both of dry and wet soil moisture conditions. When
the model is used for simulating historical data, it is advisable to use at least 2 or 3
years of data as a calibration period. Nevertheless, when the model is part of an

Table 1
Approximate initial guess for parameters C and D

c D

Hillslopes l-2 l-100


Brooks l-3 100-1000
Rivers l-3 looO-loo00
Large rivers l-3 10000-100000
E. Todini / Journal of Hydrology 175 (1996) 339-382 357

on-line real-time flood forecasting system, 1 year of data may be suIlicient to start,
and additional calibrations can be performed after 1 or 2 years of real-time
operations. This is generally advisable because the historical observation network
rarely coincides with the real-time observation network.
(3) Historical records, not necessarily continuous in time, of monthly average
temperature, for at least 10 or 20 years, should be acquired for estimating the long-
term potential evapotranspiration as described in Appendix B. Whenever possible,
data should be collected for the same stations as referred to in (2), although a sub-set
can eventually be used.
(4) Rating curves for all the hydrometric stations where the discharge is to be
simulated.
(5) Geographical coordinates and elevation above mean sea-level, for all the
measurement stations.
(6) Whenever possible, a soil map and a land use map (a coarse scale would be
enough) should be used to obtain an impression of the order of magnitude for the
various sub-catchments’ soil moisture capacity parameter values W,,,.
The calibration phase starts by resealing the measured temperatures at the average
sub-catchment elevation, by means of a thermal gradient, and then proceeding to the
estimation of the potential ’ evapotranspiration using the technique described in
Section 4. After dividing the basin into a number of sub-catchment units following
the river geomorphology, a connection tree can be identified and its topological
characteristics used to generate the correct succession of operations and routing
from the uppermost sub-basins to the downstream closing section. For each sub-
catchment, one has then to determine the hypsographic curve from the available
orographic information, and all the precipitation and temperature data must be
averaged in space (for instance, according to Thiessen polygons), over each of the
sub-catchments.
After these preliminary activities, initial values are assigned to the model
parameters. From the geometrical characteristics of all the sub-catchments, the
average length of slopes can be determined, on the assumption of the ‘open book’
schematization, as L, = A/2L, with L the length of the main channel in the specific
sub-catchment. An initial guess for the values of the parameters C and D can be
roughly given as C M 2 and D x l/So, where S0 is the average terrain slope or the
average bed slope, depending on whether the hillslopes or the river are considered.
When all the routing parameters have been assigned, all the necessary transfer func-
tions are evaluated and are then provided to the ARNO model in the form of unit
graphs. An initial guess for all the other model parameters must also be provided.
Initially, in the absence of soil and vegetation maps, the same value is used for all the
sub-catchments; it is not easy to give good starting values, but one could start by
considering that the catchment average soil water storage W,,, generally ranges
between 50 and 300 mm and the other soil moisture curve parameter 6, which
expressses the degree of homogeneity of soil characteristics, should generally be
taken between 0.1 and 0.01. The initial soil moisture condition W is not of great
relevance, because a wrong assumption for this will be evident after the first calibra-
tion run. Unfortunately, there are no fixed rules for initializing the drainage curve
358 E. Todini / Journal of Hydrology 17s (1996) 339-382

parameter values, except that c should be taken between one and two, but after one or
two runs one will be able to reproduce the recession of hydrographs by adjusting the
initial guess for the drainage parameter values. Finally, after a few calibration trials,
both the long-term water balance and the shape of the baseflow will be easily matched
by acting on the percolation and on the groundwater model parameter values.
Calibration is not performed automatically; the reason is that, in most situations,
the available information does not allow for resolving the parameter values, unless
additional knowledge of the system as well as hydrological understanding is
introduced in the estimation phase. For instance, in the case of two upstream sub-
catchments, both contribuing to the outflow in a downstream section which is to be
used for calibration, it is very unlikely that an automatic procedure can correctly
estimate the differences between the parameter values for the two sub-catchments. In
this case, one has to find one or more precipitation events when precipitation was
occurring on one of the two and use that event (or those events) for adjusting the
parameter values of the specific sub-catchment.
Calibration proceeds from the upstream sub-catchments to the downstream ones.
The first step requires the assessment of the routing parameters. This is done by
routing individually, to the downstream closing section, the flood waves produced
by the upstream ungauged sub-catchments (i.e. without combining their effects) and
by comparing the results with the observed discharges. It is relatively easy to see how
the different flood waves add together and to identify those with substantially wrong
parameters in terms of travel time or subsidence.
After this first assessment of the routing parameters, attention is focused on the
calibration of the two sets of most important model parameters. The first set refers to
the shape parameter b of the soil moisture storage curve and the maximum soil
moisture storage W,,,, and the second set relates to the parameters of the drainage
curve. Although interdependence exists among the parameters, it is generally not
difficult to reach, after a few trial-and-error iterations, a good combination of para-
meters, given that their effects are easily detectable. A very large value for the soil
moisture curve shape parameter (i.e. b > 1) will produce peaky runoff responses even
when the soil is mostly dry, and a large value for W,,, can reduce the outflow by a
factor of 100%. Furthermore, the drainage parameters are mainly associated with the
recession curves of the observed hydrographs, and again, even if some interaction
with the soil moisture storage exists, it is not too difficult to reproduce the observed
effects.
With respect to the estimation of the other parameters relevant to the percolation
and the groundwater storage, they tend to be easily identified on long recession
periods and by matching the overall mass balance. Needless to say, the most
important aspect in calibrating a continuous time rainfall-runoff model must be its
capacity to reproduce the ‘entire’ record, including minor flood events and long dry
spells, and not just a few large flood events. This will give a better guarantee of good
operational performance, because the model learns the different ways of reacting to
rainfall events as a function of its initial soil moisture conditions. The effect of the
initial conditions on runoff production can only be disregarded in urban or in mostly
impervious catchments. In fact, if one compares the amount of available soil moisture
E. Todini / Journal of Hydrology 175 (1996) 339-382 359

storage (50-300 mm) with maximum rainfall intensities (in most cases smaller than
100 mm h-‘) one can determine the different responses obtained by 100 mm of rainfall
over a dry soil or over an entirely saturated soil.
After calibration of the upstream sub-catchments, one can proceed downstream,
following the river network connection tree, until it reaches the downstream closing
section. Calibration is repeated, checking that the water balance is matched between
observed and computed discharges, until a reasonable value for the explained
variance is obtained. In general, if the available hydrological data are of good quality
as well as representative of the spatial variation, this reasonable agreement ranges
between 85% and 90% of explained variance for upstream small sub-catchments and
between 90% and 98% for the downstream larger ones, with very small biases.

1’ 0 N ‘43
“, ;

MA . R
“irr(“’
j
CANA
IR RE

Fig. 8. The Amo river catchment.


360 E. Todini / Journal of Hydrology I75 (1996) 339-382

Fig. 9. The Arno river catchment closed at Nave di Rosano and its division into sub-catchments.
E. Todini / Journal of Hydrology I75 (1996) 339-382 361

9. Examples of application of the ARNO model

The ARNO model has been applied to a number of river basins both as a catch-
ment model and as the basis for a real-time flood forecasting system. The calibration
results, although strongly affected by the quality of the available data, are generally
more than adequate for the simulation of the catchment responses to rainfall, with
explained variances ranging from 90% to 98% without the need for an automatic
parameter estimation procedure based upon the minimization of a quadratic function
of residuals. As mentioned above, this is done on purpose to preserve the physical
meaning of the parameters (Todini, 1988a).

SIEVE

Fig. 10. Schematic representation of the Amo river sub-catchments in the ARNO model.
362 E. Todini/ Journal of Hydrology 175 (1996) 339-382

2500
Obaawed discharges
--------- Computed discharges

2000

1500

1000

500

0
1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31

Fig. 11. Amo at Nave di Rosano. Calibration: December 1959.

1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31

Fig. 12. Amo at Nave di Rosano. Validation: January 1960.


E. Todini / Journal of Hydrology I75 (1996) 339-382 363

1000

--___--
500

-l-7-7--T-l--r--7--r-1-1-l-T-
0
1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31

Fig. 13. Amo at Nave di Rosano. Validation: February 1960.

--
2500r--------
Observed discharger
---.---- G,qut.d discharges

-.-----
2000

1500 -~.-_.-.___ -

1000 --_-_-

1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31

Fig. 14. Amo at Nave di Rosano. Validation: March 1960.


364 E. Todini / Journal of Hydrology 175 (1996) 339-382

9.1. The ARNO application

The ARNO model was originally developed for the Arno river in Tuscany (Fig. 8)
closed at Nave di Rosano, a few kilometres upstream of Florence. The catchment
area, of approximately 4000 km*, was divided into a number of sub-catchments as
shown in Figs. 9 and 10. The calibration of the model was performed using the few
continuous records of rainfall, temperature and water levels (sampled at 1 h intervals)
which were available in computer compatible form at the time, and rating curves for
five stations. The calibration period considered was the month of December 1959
(Fig. 1l), when a succession of medium-sized floods was observed. Three successive
months (January, February and March 1960) were then used as the validation period
(Figs. 12-14). Although the calibration and the validation periods were limited,
according to the previously stated requirements, the results were more than adequate,
presumably as a consequence of the exceptional quality of the data. The statistical
analysis of the results for the calibration and validation periods is summarized in
Table 2, where EV stands for explained variance and DC for determination coefficient
(see Franchini et al. (1996) for definitions). The quality of calibration can easily be
seen, both from Table 2 and the graphical comparison of simulated and observed
flows in Fig. 11. The statistics in Table 2 and the graphical comparisons in Figs. 12-14
for the validation period indicate a comparable level of agreement to that obtained
for the calibration period. Moreover, an additional validation test shows that the
validity of the calibration remained unchanged after 20 years. Fig. 15 demonstrates
good agreement between simulated and observed flows, and also shows the
reconstruction of the flows during a flooding period in 1981, when the recording
level station was not operational for a number of hours (steady discharge measures
during the latest flood event).
After calibration on the Arno river, the model was compared with other existing
conceptual models by Franchini and Pacciani (1991) and Franchini et al. (1996) on
the Sieve river, a tributary of the Arno (note that, for the comparison, the Sieve
catchment was not divided into sub-basins, which explains a small difference in the

Table 2
Statistical analysis of ARNO model residuals

Catchment Amo model

Calibration Validation

EV DC EV DC

Amo-Nave di Rosano’ 0.922 0.922 0.932 0,922


Amo-Nave di Rosanob 0.943 0.943 0.949 0.939
Sieve-Fomacina 0.886 0.878 0.866 0.865
Amo-Subbiano 0.901 0.885 0.901 0.898
Chiana-P. Ferrovia 0.891 0.886 0.822 0.805

’ Rainfall over the entire catchment.


b Rainfall over Valdamo, Ambra and Sieve only.
.----_--- ---__--- E. Todini / Journal of Hydrology 175 (1996) 339-382 365

2500

r --------
0bsarve.d
Computad
discharges
disqhargas

--- __ ---.--- -----.---__----


2000

1500 I- __--_-_----_ ------A..----___

4
II
1I
I
r.--._-- ..-..- _____,_(_
1I
1000
leve.1km@e stuck

500

0 -iv-- I 111-T-T-m
i 3 5 ? 9 11 13 15 17 19 21 23 25 27 29 31

Fig. 15. Amo at Nave di Rosano. Split sample test: December 1981.

ARNO model results when compared with Table 2). From Table 3, TOPMODEL
(Franchini et al., 1996) and the ARNO model clearly perform best, and, although
TOPMODEL performs better in the calibration period, the ARNO model shows a
smaller degradation in performance in the validation period, indicating good
predictive stability.

Table 3
Statistical analysis of different model residuals

Catchment Sieve Fomacina results

Calibration Validation

EV DC EV DC

ARNO 0.888 0.880 0.853 0.851


TOPMODEL 0.914 0.912 0.852 0.846
Xinanjiang 0.880 0.840 0.822 0.821
Stanford IV 0.843 0.830 0.845 0.844
Sacramento 0.836 0.821 0.835 0.833
Tank 0.875 0.856 0.847 0.845
APIC 0.776 0.751 0.820 0.779
SSARR 0.867 0.829 0.834 0.824
366 E. Todini / Journal of Hydrology 175 (1996) 339-382

9.2. Real-time forecasting applications

Following its derivation and application to the Arno river data, the model has
become the basic component of the European Flood Forecasting Operational
Real-Time System (EFFORTS), a fully automatic on-line computer package for
real-time flood forecasting developed under a CEC-funded R&D project for the
study and the implementation of a real-time forecasting scheme for the Fuchun
river (ET&P, 1992), whose catchment lies in the Zhejiang province in mainland
China, SW of Nanjing and Shanghai. The application of the ARNO model can be
considered fairly successful, given that, although the number of raingauges (23) was
limited when compared with the size of the cat&n-rent area to be modelled
(18230 km2), in, the calibration phase the model explained 96% of the total observed
runoff variance, and Fig. 16 also shows the good quality of the results. A real valida-
tion period is not available, as, in practical applications, all the available data are
generally used to improve calibration; nevertheless, the following comments were
reported by Professor Liu Gu Chong (personal communication, 1994), the Chinese
Project Leader, after the second year of on-line running: “From a flood forecasting
point of view, the system had experienced several floodings occurred during the
reported period. Undoubtedly the flood forecasting softwares, EFFORTS and
HFS, were put into use for those floodings and satisfactory forecasts were obtained.
For example, for the flooding of June 19, 1993 a forthcoming discharge of 11 100
cubic meters per second at Lanxi station was forecast, compared with an actually

- obwrvad
_____ compiled

Fig. 16. Fuchun at Lan Xi. Calibration period: 24 April 1992-1 August 1992.
E. Todini / Journal of Hydrology 175 (1996) 339-382 367

measured value of 11200 cubic meters per second. Such a forecast not only helped the
action of flood mitigation measures within the basin, but increased the power
production by regulating the Fuchun hydropower plant.” It should be noted that
both the EFFORTS package and the HFS (Hydrological Forecasting System),
developed under sub-contract by the Laboratoire d’Hydraulique de France, operate
with the ARNO model.
In 1993, the ARNO model was applied to three rivers in Italy: the upper Tiber in
Central Italy, and the Adda and the Oglio rivers in Northern Italy. The application to
the Tiber was performed on behalf of the Italian National Electric Company (ENEL),
to control better the reservoir of Corbara which closes a catchment area of approxi-
mately 5250 km*. In Fig. 17 the results of the model are compared with data for the
station of Ponte Nuovo (4179 km*), where the explained variance is 90%. The
relatively low value of the variance is mainly associated with the lack of spatial
coverage of the rain gauges; the number has recently been increased and a new
calibration is under way. Application of the model to the Adda and Oglio rivers
was also aimed at controlling the levels of the Lake Como and of Lake Iseo,
respectively. In this case, not only were very few flow data available for calibration,
but also all the natural low flows were considerably modified by the large number of
interconnected upstream reservoirs, which made it impossible to provide a
meaningful measure of the agreement between observed and computed values. Never-
theless, although no information is available for the Iseo river, the reported

Fig. 17. Tevere at Ponte Nuovo. Calibration period: 1 April 1991-31 December 1991.
368 E. Todini / Journal of Hydrology I75 (1996) 339-382

operational performances of the model on the Adda river are rather impressive. If one
considers that the operational range of Lake Como is 1.3 m, it is possible to appreci-
ate the following performances described by Eng. Luigi Bertoli (personal communi-
cation, 1994), the Technical Director of the Adda Consortium, who is in charge of the
regulation of the Lake Como gates: “In the only available flood, good forecasts were
obtained particularly for the Lake Como levels; for instance the level (in cm) in
Malgrate above the hydrographic zero was:
Time of forecast issue: Sept. 13, 1994, 01 p.m.; present level +53. 12 hours in
advance forecast (Sept. 14,01 a.m.): Forecast +69, Observed +68.
Time of forecast issue: Sept. 14, 1994, 03 p.m.; present level +82. 12 hours in
advance forecast (Sept. 15,03 a.m.): Forecast +106, Observed +104.”
In 1994 the ARNO model was also applied to the upper Danube (4037 km*) on
behalf of the Baden-Wurttemberg flood forecasting centre. The model performances
are adequate in this case, with 94% of explained variance, and the results for the
measurement station of Berg are shown in Fig. 18. Again, in 1994, a new calibration
was performed on the Arno river using the data collected by the new real-time
telemetering data acquisition system recently installed by the Tuscany Regional
Government. The new calibration substantially improved the previous one, with an
explained variance of 96%. Fig. 19 shows the entire calibration record, and Fig. 20
enlarges the rainy season to give a better appreciation of the results. One can observe
that most of the errors produced by the model are a consequence of the upstream
reservoir operations, which are evident in the medium or low flow periods and

Fig. 18. Donau at Berg. Calibration period: 1 March 1993-31 December 1993.
E. Todini / Journal of Hydrology 175 (1996) 339-382 369

Fig. 19. Amo at Nave di Rosano. Calibration period: 1 January 1992-31 December 1992.

‘.
t b

- obsmed
-- compihd

Fig. 20. Amo at Nave di Rosano. Calibration period: 1 September 1992-31 December 1992.
370 E. Todini/ Journal of Hydrology I75 (1996) 339-382

disappear during flood conditions; the model could not account for the reservoir
operations because of the lack of knowledge of the operating rule.
At present, several other operational applications of the ARNO model are under
way in Italy, such as the model of the Reno river, the model of the Ticino river, and
the extension of the Tiber model from Corbara to the river mouth, as part of a flood
warning system developed for the City of Rome.

9.3. Land use change application

Within the framework of the NERC-ESRC Land Use Programme (NELUP) of


research being undertaken at Newcastle University, the ARNO model has been fully
integrated into a decision support system (DSS) designed to assess the impacts of
future land use changes on the hydrology, ecology and economics of large river basin
systems. Its role within the DSS is to screen rapidly alternative future land use change
scenarios for significant shifts in hydrological behaviour; significant changes are then
explored in more detail using the distributed flow and transport modelling system
SHETRAN. Alternative futures can include both a redistribution of land cover types
and new land and water management strategies.
As a test of the ARNO model, and the DSS as a whole, the system has been applied
to two river basins in the UK: the Tyne River basin in NE England and the Cam River
basin in East Anglia. The two basins are hydrologically very different. The Tyne
basin, which covers an area of 2000 km2, is primarily an upland catchment possessing
little groundwater and a rapid surface response to rainfall and snow melt events.
Water management is restricted to controls at a small number of reservoirs. The
smaller Cam basin covers an area of approximately 800 km2 and includes both
groundwater and surface water dominated sub-catchments. Water management
and agricultural activity are both intensive.
For the adequate representation of both catchments, the ARNO model has been
adapted to include more complete representations of groundwater storage and move-
ment, land cover distribution, and water abstractions and effluent returns. It has been
very successfully applied to the surface water dominated sub-cat&rents, and tests on
the groundwater-dominated sub-catchments have shown acceptable but not complete
agreement between the simulated and observed runoff for a range of meteorological
conditions including dry, wet and average years. For further details, the reader is
referred to Adams et al. (1995).

9.4. Land-surface-atmosphere applications

Interesting research results have been obtained with the ARNO model in represent-
ing the land-surface-atmosphere exchanges within the framework of global climate
change studies. A noticeable improvement in the behaviour of the ECHAM general
circulation model (GCM) was obtained by substituting the ARNO soil moisture
balance component (Diimenil and Todini, 1992) for the Manabe bucket model
(Manabe, 1969). In this case, the availability of a major rivers data bank (the Global
Runoff Data Centre-GRDC) provided the opportunity to compare not only the
E. Todini 1 Journal of Hydrology 175 (1996) 339-382 371

precipitation, evapotranspiration and temperature fields produced by the ECHAM-


GCM with those available from climatological atlases, but also the predicted monthly
river discharges with those observed, a step in the calibration of GCMs which is now
considered of the utmost importance. Successively, the ARNO model was compared
with a number of different land surface-atmosphere schemes used in GCMs, within
the frame of the EC-funded research project SLAPS (Soil-Land-Atmosphere
Process Simulation) and, while preserving the overall meteorological exchanges, it
was found to produce more realistic outflow discharges than the schemes generally
used by meteorologists (Polcher et al., 1995); this effect was also confirmed in a
comparison of the ARNO scheme with the UK Meteorological Office soil-
atmosphere column model on the basis of the River Thames data (Rowntree and
Lean, 1994).

10. Conclusions and future perspectives

As a result of the increased collaboration between meteorologists, hydrologists and


soil scientists, several studies are now under way aimed at developing a new
parameterization of soil-atmosphere exchanges. Several international programmes,
such as the Global Energy and Water Cycle Experiment (GEWEX) or the Inter-
national Geosphere-Biosphere Programme (IGBP) with its Core Project ‘Biospheric
Aspects of the Hydrological Cycle’ (BAHC), deal with the problem. The possibility of
using topographical information to parameterize as well as to evaluate the lumped
model parameter values, a concept originally introduced by Beven and Kirkby (1979)
with TOPMODEL, is attracting the interest of scientists, given the practical impos-
sibility of extending to the required scales the distributed differential models such as
the SHE or the IHDM model. Although improvements in the rainfall-runoff process
representation can be sought mostly through lumping in space at the different scales
the model parameters, as a function of the distributed topographic, pedologic and
topologic information, nevertheless, the TOPMODEL approach still retains a
number of physical inconsistencies (Franchini et al., 1996), which may prevent its
generalized use. From a detailed analysis of the effects of lumping in space and a
comparison of the physical behaviour of both the ARNO model and TOPMODEL,
Todini (1995) recently proposed the TOPKAPI approach, which transforms the
rainfall-runoff process into two non-linear reservoir differential equations. Both
derive from the integration in space of the non-linear kinematic wave model; the
first represents the drainage in the soil and the second represents the overland flow
on saturated or impervious soils. The parameter values of the model are shown to be
scale dependent and obtainable from digital elevation maps and soil maps in terms of
slopes, soil permeabilities and topology (Todini, 1995).
In the mean time, when choosing among the recently available models, and in
particular between the ARNO model and TOPMODEL, in spite of the enthusiasm
of TOPMODEL proponents for its immediate matching with geographical
information systems as well as with rasterized radar images, which may give the
illusion of operating in distributed system form, one has to recognize that the
372 E. Todini / Journal of Hydrology 175 (1996) 339-382

ARNO model allows for a more comprehensive approach to rainfall-runoff


modelling. In addition, the integration of. the ARNO model components into a
standard and well-documented package not only made possible the application of
the model to a wide variety of medium and large size catchments, but also allowed for
its operational use. Operational packages must in fact not only perform well, but also
be simple, well documented and require extremely simplified calibration operations,
possibly in line with those traditionally applied by hydrologists. Last but not least, the
ARNO model, which has shown high flexibility by achieving fairly good calibrations
in a variety of different climate and geomorphological conditions, has shown a notice-
able stability of results, in that the drop in performance between calibration and
validation periods is always very small.
Until the time comes, and it could be tomorrow, that a new scheme is demonstrated
to improve either the physical interpretation of the phenomena or to simplify and
improve the calibration requirements, it is the belief of the author that the ARNO
model constitutes a robust operational tool also providing, in an extremely simplified
form, a reasonable interpretation of the overall rainfall-runoff process at the
catchment scale.

Appendix A. Derivation of the formulae for the soil moisture balance module in the
ARNO model

A sub-basin of given surface area ST (excluding the surface extent of water bodies
such as reservoirs or lakes) is, in general, formed by a mixture of pervious and less
pervious areas, the response to precipitation of which will be substantially different.
For this reason, the total area ST is divided into impervious area S1 and pervious area
s,:

ST = s, i- s, (Al)

To derive the expressions needed for the continuous updating of the soil moisture
balance, let us first deal with the amount of precipitation that falls over the pervious
area. Given that, from Eq. (Al), the entire pervious area is

s, = ST - s, (AZ)

if one denotes by (S - S,) the generic surface area at saturation, then x, defined
as

s - s,
x=gq (A3)

will indicate the percentage of pervious area at saturation. Zhao (1977) demonstrated
that the following relationholds reasonably well between the area at saturation and
the local proportion of maximum soil moisture content w/w,, where w is the
elementary area soil moisture at saturation and w, is the maximum possible soil
E. Todini / Journal of Hydrology I75 (1996) 339-382 373

moisture content in any elementary area of the catchment:


b
(A4)

This is similar to defining a cumulative distribution for the local soil moisture at
saturation, as shown in Fig. 6, defined as

w = Wm[l- (1 - x)“] (A5)


Following the description in Section 3, an analysis was first carried out for the case
where the precipitation P is larger than the potential evapotranspiration ET,, and
therefore the actual evapotranspiration ET, is assumed to coincide with the potential,
i.e.
ET, = ETP 646)
In this case, the ‘effective’ meteorological input M,, defined as the difference between
precipitation and potential evapotranspiration, is larger than zero:
M,=P-ET,=P-ET,>0 047)
With reference to Fig. 7, the surface runoff generated by the entire catchment is
obtained as the sum of two terms: the first is the product of the effective rainfall
times the percentage of impervious area, and the second is the average runoff
produced by the pervious area, which is obtained by integrating the soil moisture
capacity curve, to give
h4,+tV

40 dt if M,+w<w, 648)

or

649)

Substituting x from Eq. (A4) into Eq. (A8) gives

(Al01
374 E. Todini / Journal of Hydrology 175 (1996) 339-382

Eq. (AlO) can be integrated by substitution as follows:

,=1-L C=wm(l -q), dc= -w,dn


WOl’
c=o, q=l_W
Wi?l

t = w, p,l_M,.tw
i wm
to give
~-~w/Y?l)
R=ikf,+-w - SI ST
m nbdn = 44, +
ST
'-(‘+fe+w%n

EM +sT-%
STb+l wm [(l --K)b+l-(l -y)b+‘]
e (Al 1)

The catchment average amount of soil moisture content can then be evaluated by
integrating Eq. (A5), to give

W=jw(<)dt+(l-x)w= w”j,l - (1 -#]d5+(1 -.x)w,,,[l - (1 -x)“]


0 0

=w,+w,/(l-#dc+w,-w,x-w,(l-x)?

=wm[l-(l:#-t(l-&d[] (Al2)

Integration of Eq. (A12) can also be performed by substitution as follows:


rl=l-J, <=1-q, d‘$=-dn

c=o, r]=l

{ E = x7 7/=1-x
to give

= w, 1-(1-x)9- j&+&(l-x)y] =5[1-(1-x)Y]


[

(Al3)
E. Todini / Journal of Hydrology I75 (1996) 339-382

If all the pervious area has reached saturation, x = 1, Eq. (A13) becomes

w= w,=w”
b+l
which can be rewritten as

W, = (b+ l)Wm (Al5)

Substitution of Eq. (A15) in Eq. (A13) gives

w= W,[l -(l -x)Y] (Al(j)


Therefore the percentage of pervious surface area that has reached saturation as well
as the percentage of unsaturated area become respectively

(A17)

Substituting for x from Eq. (A4) into the first equation of (A17) gives the
relationship between the average and the elementary area soil moisture storage
quantities:
b+l

Eq. (A18) can also be written as

(A19)

from which one can obtain

and finally, bearing in mind Eq. (Al 5) one finds the value of the elementary area soil
moisture at saturation, corresponding to a specific percentage x of the pervious
surface, with x written in terms of the catchment average amount of soil moisture
content as in Eq. (A17):

w=(b+l)W,,,[l- (l-E)&] 6421)

Returning to Eq. (Al 1) and substituting for w and w, from Eqs. (A21) and (Al5),
376 E. Todini / Journal of Hydrology I75 (19%) 339-382

respectively, one obtains


sT - sI

-I
R-M,+- Wi?l
ST
b+l

1
M,+(b+l)wJl- cl-a”]
x
(1-F > m
l-
(b+ l)W,

=M,+ y

1 (Wm- W)-

(b+l)W,,,-Me--(b+l)W,+(b+l)W,
w,

X
(b+ 1)Wm

(A221
Eq. (A22), after simple algebraic manipulations, allows for estimating the runoff
when M, + w < w,, or in terms of the catchment average quantities when

M,<(b+l)W,,, (A23)

as

R=M,+

6424)

In the other case, when M, + w 2 w,, or in terms of the average quantities when

6425)

the integral appearing in Eq. (A9) is easily evaluated in terms of the catchment
average quantities, given that the integral appearing in Eq. (A9) can be written as
W,,, - W, which gives

R= s’M,+ VIM,-(W,- W)] W6)


ST
E. Todini 1 JournaL of Hydrology 175 (19%) 339-382 377

Eq. (A26) can be finally rewritten as

R=M,-qqw,- W) (A271

An analysis is now carried out of the case where the precipitation P is smaller than the
potential evapotranspiration ETP and the effective meteorological input M,, becomes
negative, i.e.
M,=P-ET,<0 (A281

implying that the runoff R is zero. The actual evapotranspiration ET, is now
computed as the precipitation P plus a quantity which depends upon M,, which is
taken to be unchanged for the surface area at saturation, whereas, for the remaining
unsaturated surface area, it is reduced by a factor which depends upon the ratio
between the actual moisture content and the total:

1
s - s, s,-s w-w,
ET,=P+M, - -
[ ST + ST wm-ws

SW,-SW,-SIW,+SIW,+STW-STW,-SW+SW,
=P+M,
sT(wm- ws)

=P+M (S-S,)W,+(ST-S)W-(ST-SI)W,
e
sT(wm - w,)

= P+M,-
ST

=p+M~T-SIXW,+(~-~)W-w~
(~29)
e ST wl?l-ws

Substituting for x from Eq. (A15) into Eq. (A29), one obtains

s _,[1-(1-~)fi]w,.+(1+)~w-w~
ET, = P+MeL

sT;;sI,l _ (1 K)';;i
= P+M,- (A30)

With reference to Fig. 7, the soil moisture content W, of the unsaturated area (1 - x),
can be computed as

w,= w-w(l-x)= w-(b+1)w~[1-(1-x)~](l-.x)

= w-(b+1)W,(1-x)+(b+l)W,(1-x)~ (A311
378 E. Todini / Journal of Hydrology 175 (19%) 339-382

and by substituting into Eq. (A30) for x, from Eq. (A4), one obtains

= W~[l+b(l-~)-(b+l)(l-$-)m] 6432)

The desired expression for the actual evapotranspiration when the precipitation is
smaller than the potential evapotranspiration is finally found by substituting for W,
from Eq. (A32) into Eq. (A30), to give

1
(1 -E)“(l -$K)
ET,=P+M,p
STs,s1 ’ -(b+ l)(l _?-)‘-b(1 -g)

=p+ (A33)

Appendix B. Estimation of parameters for the evapotranspiration module in the ARNO


model

For each of the sub-basins, the estimation of the evapotranspiration expression


(Eq. (18)) is based upon the following steps: (1) computation of the long-term
monthly average temperatures at the measurement stations; (2) computation of the
thermal gradient; (3) computation of the spatially averaged, long-term monthly mean
temperature for each sub-basin; (4) estimation of the average monthly potential
evapotranspiration according to Thornthwaite; (5) computation of the regression
coefficients cx and a for each sub-basin.

Bl. Computation of long-term monthly temperature averages

Long-term monthly temperature averages are calculated at the available


temperature measurement stations, for the calendar months over a long period of
time (a minimum of 10 years is required), and constitute the basis for the estimation of
the evapotranspiration.
E. Todini / Journal of Hydrology I75 (1996) 339-382 379

B2. Computation of the thermal gradient

The influence of the temperature gradient may be significant if the stations are
located at different heights or if the catchment shows large topographical variation.
The temperature gradient varies according to place and time of the year. For the
purpose of the ARNO model an average value of the gradient is used and estimated
by means of long-term averages. If this is not available a table such as the ‘US
Standard Atmosphere’ or ‘NACA Standard Atmosphere’ can be used (see Hess,
1959, p. 85), which is based on the following assumptions: (1) the air temperature
at sea-level is 15°C and the pressure is 1013.25 hPa; (2) the air is dry and follows
the perfect gas law; (3) the gravity acceleration is assumed constant and equal to
980.665 cm se2.

B3. Computation of the long-term monthly temperature for a sub-basin

All the long-term mean temperatures computed for the various temperature
stations are reduced to a common level (for instance, mean sea-level) by means of
the thermal gradient described above. By using Thiessen polygons, a spatial average
temperature is computed for each sub-basin, which is then reduced to the mean
elevation of the sub-catchment, again using the thermal gradient.

B4. Estimation of the average monthly potential evapotranspiration according to


Thornthwaite

As was mentioned in Section 4, evapotranspiration plays a major role in the


rainfall-runoff process not really in terms of its instantaneous value, but in terms
of its accumulated effect over time on the soil moisture volume depletion. Therefore it
is extremely important that the long-term mean values be accurately reproduced, even
if the expression used, which is generally different from the Penman-Monteith
equation owing to the lack of data, may not be very precise on a short time
interval such as 1 or 3 h. The long-term values of the average daily reference potential
evapotranspiration to be used for the calibration of the evapotranspiration
parameters in the ARNO model (although any alternative formula could be used)
are generally computed for a given sub-catchment according to Thornthwaite and
Mather (1955), by means of the following formula:

44 W T(i) k2
Edi) = 1630~
1o _

(Bl)
[ kl 1

where Eth(i) is the average monthly potential evapotranspiration (mm day-‘) in


month i, T(i) is long-term average temperature (“C) in month i,
a(i) = [n(i)/30][N(i)/12] is a sunshine percentage index, n(i) is number of days in
month i, N(i) is mean daily duration of maximum possible sunshine hours
(Doorembos et al., 1984), kl is the thermal index defined as Cjz1[T(i)/5]1.514, and
k2 = 0.49239 + (1792 x 10-5)b - (771 x 10-7)b2 + (675 x 10-9)b3.
380 E. Todini / Journal of Hydrology 175 (1996) 339-382

B5. Computation of the regression coejkients a and b for each sub-basin

The expression actually used in the ARNO model for the estimation of the
reference evapotranspiration at each time step is a linear expression in T, in such a
way that El,,(i) can be disaggregated to short time steps without losing the long-term
balance. It has been found that the radiation formula presented in Section 4, rewritten
in a more convenient parametric form, closely approaches the Thornthwaite
estimates, i.e. one can write

l&(i) = (Y + ,L?N(i)W,,(i)‘T(i) 032)


where a and p are two model parameters. W,, is a compensation factor dependent
upon the monthly average temperature and elevation above sea-level, which can be
determined by means of a table given by Doorembos et al. (1984) that can be well
approximated by means of the following parabola:

W,,(i) = AT(i)2 + BT(i) + C (B3)


where T(i) is the monthly long-term temperature (in “C) and A, B and C are three
fitting parameters. Eq. (B2) very closely approximates Eq. (Bl), generally providing a
determination coefficient of R2 = 0.999 or larger.
Given the linearity of the equation with respect to temperature, once the
parameters a and /3 have been estimated, the following expression is finally used at
a more appropriate time step for the rainfall-runoff model:
ET0 = h +,&V(i) W,(i)T 034)
where ET,, is the reference evapotranspiration for a specified time step (mm per time
step), & and fi are the estimated values for the two model parameters, and T is the
average temperature over a specified time step (in “C).

References

Abbott, M.B., Bathurst, J.C., Cunge, J.A., O’Connell, P.E. and Rasmussen, J., 1986a. An introduction to
the European Hydrological System-Systeme Hydrologique Europeen, ‘SHE’, 1: History and
philosophy of a physically-based, distributed modelling system. J. Hydrol., 87: 45-59.
Abbott, M.B., Bathurst, J.C., Cunge, J.A., O’Connell, P.E. and Rasmussen, J., 1986b. An introduction to
the European Hydrological System-Systeme Hydrologique Europeexr, ‘SHE’, 2: Structure of
physically-based, distributed modelling system. J. Hydrol., 87: 61-77.
Adams, R., Dunn, S.M., Lunn, R., Mackay, R. and O’Callaghan, J.R., 1995. Validating the NELUP
hydrological models for river basin planning. J. Environ. Plann. Manage., 38(l): 53-76.
Amorocho, J. and Hart, G.T., 1964. Nonlinear analysis of hydrologic systems. Univ. Calif. Water Resour.
Centre Contrib. 40.
Bathurst, J.C., 1986. Physically based distributed modelling of an upland catchment using the Systeme
Hydrologique Europeen. J. Hydrol., 87: 79-102.
Beven, K.J., 1989. Changing ideas in Hydrology-the case of’physically based models. J. Hydrol., 105:
157-172.
Beven, K.J. and Kirkby, M.J., 1979. A physically-based variable contributing area model of basin
hydrology. Hydrol. Sci. Bull., 24(l): 43-69.
E. Todini / Journal of Hydrology 175 (1996) 339-382 381

Beven, K.J. and Wood, E., 1983. Catchment geomorphology and the dynamics of contributing areas. J.
Hydrol., 65: 139-158.
Beven, K.J., Kirkly, M.J., Schofield, N. and Tagg, A.F., 1984. Testing a physically-based flood forecasting
model (TOPMODEL) for three UK catchments. J. Hydrol., 69: 119-143.
Beven, K.J., Calver, A. and Morris, E.M., 1987. The Institute of Hydrology Distributed Model (IHDM).
Rep. 98, Institute of Hydrology, Wallingford.
Binley, A., Elgy, J. and Beven, K.J., 1989. A physically based model of heterogeneous hillslopes 1. Runoff
production, Water Resour. Res., 25(6): 1219-1226.
Bumash, R.J.C., Ferral, R.L. and McGuire, R.A. 1973. A general streamflow simulation system-
conceptual modelling for digital computers. Report by the Joint Federal State River Forecasts Center,
Sacramento, CA.
Cordoba, J.R. and Rodriguez-Iturbe, I., 1983. Geomorphologic estimation of extreme flow probabilities. J.
Hydrol., 65: 159-173.
Crawford, N.H. and Linsley, R.K., 1966. Digital simulation in hydrology, Stanford Watershed Model IV.
Stanford Univ. Dep. Civil Eng. Tech. Rep. 39.
Cunge, J.A., 1969. On the subject of a flood propagation method. J. Hydraul. Res. IAHR, 7: 205-230.
Dooge, J.C.I., 1957. The rational method for estimating flood peak. Engineering (London), 184: 31 l-374.
Dooge, J.C.I., 1973. The linear theory of hydrologic systems. US Dep. Agric. Tech. Bull., 1468.
Dooge, J.C.I., Bruen, M. and Dowley, A., 1994. Spatial variability of land surface processes. Final Report
of CEC Research Contract PL 890016 EPOCH.
Doorembos, J., Pruitt, W.O., Aboukhaled, A., Damagnez, J., Dastane, N.G., van den Berg, C., Rijtema,
P.E., Ashford, O.M. and Frere, M., 1984. Guidelines for predicting crop water requirements. FAO
Brig. Drainage Pap., 24.
Dtlmenil, L. and Todini, E., 1992. A rainfall-runoff scheme for use in the Hamburg climate model. In: J.P.
O’Kane (Editor), Advances in Theoretical Hydrology-A Tribute to James Dooge. Elsevier,
Amsterdam.
Dunne, T., 1978. Field studies of hillslope flow processes. In: M.J. Kirkby (Editor), Hillslope Processes.
Wiley, New York, pp. 227-293.
ET&P, 1992. The Fuchun River project-a computer based real time system-EC-China cooperation.
Final Report Research Contract CI13-0004-I (A), Bologna.
Franchini, M. and Pacciani, M., 1991. Comparative analysis of several conceptual rainfall runoff models. J.
Hydrol., 122: 161-219.
Franchini, M. and Todini, E., 1989. PABL: a parabolic and backwater scheme with lateral inflow and
outflow. Fifth IAHR Int. Symp. on Stochastic Hydraulics, Report No. 10, Institute for Hydraulic
Construction, University of Bologna.
Franchini, M., Wendling, J., Obled, Ch. and Todini, E. 1996. Physical interpretation and sensitivity
analysis of the TOPMODEL. J. Hydrol., 175: 293-338.
Freeze, R.A., 1980. A stochastic-conceptual analysis of rainfall-runoff processes on a hillslope. Water
Resour. Res., 16(8): 1272-1283.
Freeze, R.A. and Harlan, R.L., 1969. Blueprint for a physically based digitally simulated hydrologic
response model. J: Hydrol., 9: 237-258.
Hess, S.L., 1959. Introduction to Theoretical Meteorology. Henry Holt, New York.
Kalinin, G.P. and Miljukov, P.I., 1958. Approximate methods of computing unsteady movement of water
masses. Trans. Central Forecasting Inst., 68.
Kibler, D.F. and Woolhiser, D.A., 1970. The kinematic cascade as a hydrologic model. Colo. State Univ.
(Fort Collins) Hydrol. Pap. 39.
Manabe, S., 1969. Climate and ocean circulation: I. The atmospheric circulation and the hydrology of the
Earth’s surface. Mon. Weather Rev., 97: 739-774.
Moore, R.J., 1985. The probability-distributed principle and runoff production at point and basin scales.
Hydrol. Sci. J., 30(2): 273-297.
Moore, R.J. and Clarke, R.T., 1981. A distribution function approach to rainfall-runoff modelling. Water
Resour. Res., 17(5): 1367-1382.
Naden, P.S., 1992. Spatial variability in flood estimation for large catchments: the exploitation of channel
network structure. J. Hydrol. Sci., 37(1-2): 53-71.
382 E. Todini 1 Journal of Hydrology 175 (1996) 339-382

Nalbantis, I., Obled, C. and Rodriguez, J.Y., 1988a. Modelisation pluie-debit: validation par simulation de
la methode DPFT. Houille Blanche, 5-6: 415-424.
Nalbantis, I., Obled, C. and Rodriguez, J.Y., 1988b. Rainfall-runoff modeling by the FDTF method:
testing and validation by generated data. IVth Int. Symp., Analyse des Systernes dam la Gestion des
Ressources en Eau, Rabat, Vol. I, pp. 35-48.
Nash, J.E., 1958. The form of the instantaneous unit hydrograph. IAHS Publ., 45: 114-121.
Natale, L. and Todini, E., 1977. A constrained parameter estimation technique for linear models in
hydrology. In: Mathematical Models for Surface Water Hydrology. Wiley, Chichester.
Polcher, J., Laval, K., Diimenil, L., Lean, J. and Rowntree, P.R., 1995. Comparing three land surface
schemes used in GCMs. J. Hydrol., submitted.
Rockwood, D.M., Davis, E.D. and Anderson, J.A., 1972. User Manual for COSSARR Model. US Army
Engineering Division North Pacific, Portland, OR.
Rodriguez-Iturbe, I. and Valdes, J.B., 1979. The geomorphologic structure of the hydrologic response.
Water Resour. Res., 15(6): 1409-1420.
Rodriguez-Iturbe, I., Gonzales-Sanabria, M. and Bras, R.L., 1982. A geomorphocliiatic theory of the
instantaneous unit hydrograph. Water Resour. Res., 18(4): 877-886.
Rosso, R., 1983. Sulla taratura della risposta idrologica in base ai caratteri morfologici della rete
idrografica. Idrotecnica, 1: 3-25.
Rosso, R., 1984. Nash model relation to Horton order ratios. Water Resour. Res., 20(7): 914-920.
Rowntree, P.R. and Lean, J., 1994. Validation of hydrological schemes for climate models against
catchment data. J. Hydrol., 155: 301-323.
Rutter, A.J., Kershaw, K.A., Robins, P.C. and Marton, A.J., 1971. A predictive model of rainfall
interception in forests. I: Derivation of the model from observations in a plantation of Corsican
Pine. Agric. Meteorol., 9: 367.
Rutter, A.J., Marton, A.J. and Robins, PC., 1975. A predictive model of rainfall interception in forests. II:
Generalisation of the model and comparison with observations in some coniferous and hardwood
stands. J. Appl. Ecol., 12: 367.
Shuttleworth, W.J., 1979. Evaporation. Rep. 56, Institute of Hydrology, Wallingford.
Sorooshian, S. and Gupta, V.K., 1983. Automatic calibration of conceptual rainfall-runoff models: the
question of parameter observability and uniqueness. Water Resour. Res., 19(l): 260-268.
Sugawara, M., Ozaki, E., Watanabe, I. and Katsuyama, Y., 1976. Tank Model and its application to Bird
Creek, Wollombi Brook, Bihin River, Sanaga River, and Nam Mune. Res. Note 11, National Centre for
Disaster Prevention, Tokyo.
Thomthwaite, C.W. and Mather, J.R., 1955. The water balance. Publications in Climatology, 8(l).
Laboratory of Climatology, Centerton, NJ.
Todini, E., 1988a. 11modello afllussi deflussi de1 fiume Amo. Relazione Generale dello studio per conto
della Regione Toscana, Tech. Report, Bologna.
Todini, E., 1988b. Rainfall runoff modelling: past, present and future. J. Hydrol., 100: 341-352.
Todini, E., 1989. Flood forecasting models, EXCERPTA, 4: 117-162.
Todini, E., 1995. New trends in modelling soil processes from hillslope to GCM scales. In: H.R. Oliver and
S.A. Oliver (Editors), The Role of Water and the Hydrological Cycle in Global Change, Nato AS1
Series, Springer Verlag, Berlin.
Todini, E. and Bossi, A., 1986. PAB (Parabolic and Backwater), an unconditionally stable flood routing
scheme particularly suited for real time forecasting and control. J. Hydraul. Res., 24(5): 405-424.
World Meteorological Organization (WMO), 1975. Intercomparison of conceptual models used in
operational hydrological forecasting. Operational Hydrology Rep. 7. WMO 429. WMO, Geneva.
Zhao, R.J., 1977. Flood forecasting method for humid regions of China. East China College of Hydraulic
Engineering, Nanjing.
Zhao, R.J., 1984. Watershed Hydrological Modelling. Water Resources and Electric Power Press, Beijing.

You might also like