Neppel - Spatial Extension of Extreme Raifall

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

ATMOSPHERIC

RESEARCH
ELSEVIER Atmospheric Research 45 (1997) 183-199

Spatial extension of extreme rainfall events: return


period of isohyets area and influence of rain gauges
network evolution
L. Neppel *, M. Desbordes, J.M. Masson
ISTEEM, Laboratoire G£ofluide Bassin Eau UMR 5569, Equipe Hydrologie Case Courrier 056, Unic,ersit~ de
Montpellier II, Place Eugene Bataillon, Montpellier Cedex 5, 34095, France
Received 22 April 1997; accepted 21 July 1997

Abstract

Those last 10 yrs, many cities in southern Europe have been affected by heavy rainfall events
leading to severe runoffs. Thus, the assessment of rainfall risk needs a better knowledge of the
climate hazards and particularly the rainfall hazards. A spatial stochastic analysis is proposed here,
based on rainfall events selected during the 1958-1993 period in the South of France, using the
French weather service daily rainfall database. The isohyet areas defined at various rainfall
thresholds ~- are estimated. Because of the sampling constraints, two different time steps have
been considered: 24 and 48 h. A two parameters gamma distribution law has been fitted to each
sample of isohyet areas, and the two parameters are related to the threshold ~-. For those 36 yrs,
the evolution of the network of rain gauges in space and time has been described. The empirical
expression of the probability to observe a given rainy area with a given network configuration,
using simulation, has been proposed. The bias introduced in sampling events has been estimated
and it is shown that its influence on isohyet areas quantiles could be negligible, for studies based
on daily rainfall gauges. © 1997 Elsevier Science B.V.

Keywords: Rainfall events; Isohyet areas; Rainfall gauges

I. Introduction

For the last 10 yrs, mainly in southern Europe, many cities have been faced with
heavy rainfalls leading to severe runoff floodings. In the South of France, the flooding
of N~mes in October 1988 (420 m m in 6 to 7 h) (Desbordes et al., 1989) or

* Corresponding author. Tel.: +33 04 67144262; fax: +33 04 67524861; e-mail: neppel@dstu.univ-
montp2.fr.

0169-8095/97/$17.00 © 1997 Elsevier Science B.V. All rights reserved.


PII S0 169-8095(97)0003 8-0
184 L. Neppel et al. /Atmospheric Research 45 (1997) 183-199

Vaison-la-Romaine in September, 1992 (300 mm in 3 to 4 h) (Benech et al., 1993) are


the best known and best analysed ones. Cities in this region have now to assess the
rainfall risks, not only when planing their development, but also in order to estimate the
consequences of such rainfall events on existing urbanised areas, and to imagine how
such a risk could be managed in these areas.
The most usual rainfall risk assessment is based on a stochastic approach, mainly
because we are not able, at present, to define precisely if rainfall processes are or are not
fully decisive, particularly when extreme rainfalls are considered. The space-time scales
of convective cell dynamic, leading to the highest intensity, is difficult to represent
because of the number of parameters to estimate and of the lack of knowledge of the
physical processes which govern rainfalls at those required scales (Creutin et al., 1994;
Juvanon du Vachat, 1994).
The point rainfall frequency analysis, for mean rainfalls over a given duration, 6 t,
remains the most used method. There is a very wide choice of stochastic models, never
supported by some theoretical consideration (Bobee et al., 1993): parametric ones like
the TCEV law for extreme rainfall (Rossi et al., 1984) or the nonparametric ones like the
Villasenor method which seems to perform better than ~classical>> laws (Nguhyen et al.,
1993). It leads to a wide range of return periods according to the distribution function
used, the recording period, and the sampling fluctuation due to partial series (Desbordes
et al., 1989; Masson, 1991). It is well known that accurate quantiles estimations need the
largest record period, as well as the best historical information (Hosking and Wallis,
1986; Duband, 1994). Over a large period, the most comprehensive information consist
in rain gauges, but the evolution of the gauges network in time and space leads to a
biased sampling.
To fill the gap in rainfall time series, other approaches are based on simulating data
series, using stochastic models. Their structure depends on the variables and time scale
used for the description of the events. Some of them operate on a continuous basis,
using the Neyman-Scott processes (Waymire and Gupta, 1981; Cowpertwait, 1991) or
the Barthley-Lewis ones (Rodriguez-Iturbe et al., 1987; Bartual-Garcia and Marco,
1990). They are based on the assumption that storms could be considered as aggregating
group hierarchy. Disaggregating methods enable to simulate series at smaller duration,
from daily to hourly rainfalls (Hershenhorn and Woolhiser, 1987; Econopouly et al.,
1990) or at urban hydrology scales from hourly to 5-rain rainfalls (Cowpertwait et al.,
1996). However, they need the assumption of a spatial stationary time structure. Croley
et al. (1978), Tourasse (1981) and Cernesson and Lavabre (1994) developed hourly
hyetographs models, describing time structure events using nine independent variables.
The strongest limitation of all these models is that they are based on point rainfall
analysis, even if the development of the regionalisation allows their transposition
(Cernesson et al., 1996; Cowpertwait et al., 1996). However, for the Mediterranean
region, great spatial variations of rainfall depths frequencies can be observed during the
same event. During the N~mes flooding event, for example, the maximal depth for 6 h
varies from 120 yrs to more than 400 yrs for 3 gauges less than 10 km apart (Desbordes
and Masson, 1992). Moreover, and especially for aggregating models, many parameters
are required; their estimations sometimes need radar data, inconsistent with large periods
of observation.
L. Neppel et al. /Atmospheric Research 45 (1997) 183-199 185

The transposition method (Foufoula-Georgiou, 1989; Fontaine and Potter, 1989)


enables a spatial risk assessment by the estimation of the spatial average depth
distribution function over a given catchment, integrated over the storm event duration. It
relies on the assumption that the catchment belongs to a climatological homogeneous
region, where an event observed at a given site could theoretically occur elsewhere in
the region. The description of the space-time structure of the rainfield is the crucial
problem: Franchini et al. (1996), using DAD curves (Depth Areal Duration) coupled
with a disaggregation model, estimate the maximal runoff density function. Shah et al.
(1996) propose another space-time description of the rainfield to build a space-time
hourly rainfall model. It is based on the association of an autoregressive Arima
time-structure model with a turning band method for spatial-structure. However, the
authors underline the limitation to small basins where the spatial rainfield variance is
low.
In this paper, the extreme rainfall risk has been analysed on the Languedoc-Roussil-
Ion scale (LR), a region in the south of France, along the Mediterranean sea. It is based
on a 36-yr observation period of the regional rain gauge network. After describing the
selection of rainfall events and the evolution of the network, a spatial approach is
proposed using a regional frequency estimation of isohyet areas. The influence of the
network evolution on the estimation of the quantiles is presented in the last part of the
study.

2. The selection of rainfall events and the network evolution

2.1. The selection o f rainfall events

The LR region covers an area of about 28,000 km 2 in the South of France (Fig, I a).
The rainfall data are belonging to the <<Pluvio>> data base, managed by Mrt~o-France,
the French Weather Service, and the rainfall events has been selected by that service
(Jacq, 1995). The observation series covers the 1958-1993 period. The daily rainfall
gauges have been considered. The rainfall events have been selected with a 190
m m / d a y threshold at one gauge at least. Such a threshold has a punctual return period
varying from 1.1 to more than 500 yrs when considering 30 rain gauges with the same
period of observation.
In the data base, a given <<meteorological>> day begins at 6 h UTC the j - 1 day and
finished at 6 h UTC the j day. In order to select an event running over the 6 h UTC
limit, rainfall depths with more than 190 mm for two consecutive days have also been
extracted from the data base. In fact, the true duration of a selected event is unknown,
and according to the chosen threshold it may vary from some hours to 48 h. A total of
93 events have thus been selected. Two samples will be considered: the first ($48)
includes the 93 events with a maximum duration 6 t of 48 h and the second ($24)
includes 100 rainfields (not necessarily independent if they come from the same
meteorological situation) integrated over a maximum duration of 24 h.
186 L. Neppel et al. /Atmospheric Research 45 (1997) 183-199

Z"n //4~/

) ~$" 4

= / ~ ~.~,,,.?,

Spain ~--~ MediterraneanSea ]


(b)
300 (c)
o °o o°O o o o ~=

/
~T + o Oo %
25O o ~ ,~ ~ + ° r

200!
o /~++ , ..--~. o,
-~ + ÷+¢ + /
1001
I o, ++ ~
50 ~° 8 oo+ ++ + ~e n

1958
r~o ~o ~o 7~0 800 550 600 650 700 750 800
(km) (kin)

Fig. 1. Region Languedoc-Roussillon (a) and illustration of the evolution of the rain gauges network: location
of the rain gauges in 1958 (b), and 1972 (c). * represent gauges operating each year during the 36-yr
observation period of the region, o represent the available surrounding gauges in the neighbouring region.

2.2. The evolution of the daily rain gauges network during the 1958-1993 period

During the recording period, the density of the rain gauge network has been
changing. From 254 stations in 1958, it reached 353 in 1972 (Fig. lb,c). Since 1976, the
number of gauges over the region has known little evolution and has shown a tendency
to decrease (Fig. 2). In 1993, only 298 rain gauges were used, which leads to the same
regional density as in 1963 (10.6 gauges/1000 km2).
The distances between the gauges are estimated as following. For a given year, a disc
of diameter R k is centred on each gauge in service. The proportion of gauges, Q(R,),
which have at least one neighbour inside the disc border is estimated. For each year,
Q(R k) is studied according to R k.
L. Neppel et al. / Atmospheric Research 45 (1997) 183-199 187

36O
.............................................. • . . . . i .................................................................
::~ ............... i ................. ...'.++.o...++........~..!
- ....o. i ........... i ..... ....... i .........
34o ........... +.............. e+ .......... ] .......... ~ ............... ] ....... ::.......... i ........... :+.....
33O IO +
320 ......................................................... ~............................................... ~.........
310 • + [
3oo
290
........... i-.- .... i............ L ........ i+++...o., +~..+....o%,.,,~
..........................................................................................................................
.,...!+...%
280 . . . . . . 1 ° ........... !.............. ,.............. ~. . . . . . . . . . ]. . . . . . . . . . . +......... ~ ..... ~ .........
J 270

26O
........ 41------+ ............. + .........

-e .......... i ......................................................
+............ i ............. i .......... ...........

i....................................
........... i

i .....
.....

240 I I I t I I I t I b---I I I I -I ~ I I L ~ I i I I I I I i i ! I

1958 1962 1966 1970 1974 1978 1982 1986 19£0

Fig. 2. N u m b e r of g a u g e s b r o u g h t into service over the L a n g u e d o c - R o u s s i l l o n region, e a c h years between


1958 a n d 1993.

When R~ tends to zero, Q(R k) estimates the proportion of clustered gauges. The R k
value, named d, corresponding to the 100% proportion, measures the largest intergauges
distance. The higher Q(R k) and d, the less uniform the spatial location of the rain
gauges over the LR region is.
Between 1958 and 1993, d varied from 18 km to 13 kin, while Q(1 km) varied from
6 to 10% between 1958 and 1975 against 1 to 3% between 1976 and 1993 (Fig. 3). It
reveals the importance of clustered gauges between 1958 and 1976, separated by less
than 1 krn, whereas large distances up to 20 km existed: the network was not
homogeneous over the LR region. But, since 1976, a reorganisation of the network has
been operated, the clustered gauges have been reduced so have been the largest
intergauge distances: the network has become more homogeneous over the region.

100 T ............................................................................. Z ............ ~ "c ........ .~______-~ --

Q(Rk) m
(*/.)
60

50

40

30

20

lO

0 t t I t I i I i I I I q- I - I ~- I I I --~
2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

Rk in km.

Fig. 3. Evolution, a c c o r d i n g to the radius, Rk, o f the proportion o f gauges, Q(Rk), w h i c h h a v e at least one
other g a u g e in the n e i g h b o u r h o o d Rk, for e a c h y e a r between 1958 a n d 1993. The years not represented are
located b e t w e e n the 1969 a n d 1972 c u r v e a n d the dotted curve.
188 L. Neppel et al. /Atmospheric Research 45 (1997) 183-199

3. Estimation of isohyets area frequency

3.1. Isohyets area estimation

Since rain gauges data has been used, rainfalls have only been known at sampled
points. Spatial extension studies of heavy rainfall events needs to know the rainfall depth
at all points of the domain covered by the selected events. So for each event, the domain
area where it rains above a given rainfall threshold is estimated:

S~(z) = {u ~ Dk/Hk(u ) >_ ~}, (1)


where: index k is referring to the kth event; r is a rainfall threshold in mm; D k is the
domain included in the LR region covered by the kth event, i.e., the points group
belonging to LR where rainfall intensity is over 0 ram; Sk(r) is the area of the points
group belonging to D k where rainfall intensity exceed r mm; in the following text, it is
called ~dsohyet a r e a ) ~ ; Hk(u i) is the point rainfall, measured at a u i = ( x i , y i) point, with
i = 1 . . . n k.
According to the available data, Eq. (1) must be solved by the way of an interpolation
method. The choice of the spatial interpolator (Neppel et al., 1997) has led to a
climatological kriging with two major direction of anisotropy (Journel and Huijbregts,
1978).
For each event, the areas of isohyets Sk(r) are estimated by Eq. (2), where the
interpolated rainfall H * is calculated at the m k nodes of a grid system with a p * p km 2
mesh, over the D~ domain:

mk
= fo {u//-/; (u) >_ }du = p 2 E ( u # n ; ( u i ) >_7 , (2)
k i=1

where m k is the number of nodes of the grid system describing the D k domain,
associated to the kth event.

3.2. Isohyets area frequency

For each sample $24 and $48, the isohyets area defined by ~-= 50(50)250 mm and
z = 50(10)300 mm, respectively, are estimated. The notation a(b)c means that values
from a to c are considered using an interval of amplitude b. For thresholds below 50
mm, the S~(z) domains are imprecisely defined due to edge effects as a result of the
chosen region (sea side in the south and administrative limits in the north). For
thresholds above 250 and 300 mm for a 24- and 48-h duration, respectively, the number
of Sk(z) becomes smaller than 30 and the estimates of spatial frequency are too sensitive
to sampling fluctuations.
A probabilistic model is then applied to each S(~-) sample. Theoretical probability
functions are tried using a X2-test at a 10% confident limit. The gamma function with 2
L. Neppel et al. / Atmospheric Research 45 (1997) 183-199 189

Table 1
Gamma function parameters, A1 and A2, fitted on the samples of surfaces in function of the isohyets
threshold. AI* and A2 * represent the estimations of A1 and A2 with the regression model, trAl* and
o'A2 * are standard deviations of Al* and A2 *, in percentage of A1 and A2, respectively
r (mm) Duration: 24 h Duration: 48 h
A1 AI* orAl* (%) A2 A2* o'A2* (%) A1 AI* o'Al*(%) A2 A2* o'A2* (%)
50 2946 2966 3.8 2.238 2.153 3.2 4059 4108 2.2 2.205 2.158 2.8
60 2759 2797 4.1 1.972 1.932 3.7 4207 4150 2.1 1.858 1.884 3.4
70 2620 2630 4.3 1.709 1.724 4.3 4263 4161 2.1 1.589 1.640 4.0
80 2487 2465 4.5 1.478 1.530 5.0 4203 4143 2.1 1.394 1.425 4.5
90 2345 2304 4.7 1.284 1.349 5.7 4099 4098 2.2 1.229 1.237 5.2
100 2201 2146 5.0 1.124 1.180 6.5 3969 4028 2.3 1.092 1.074 5.9
110 2041 1994 5.4 0.986 1.025 7.5 3828 3936 2.4 0.975 0.935 6.7
120 1898 1848 5.8 0.851 0.883 8.6 3723 3824 2.5 0.862 0.817 7.6
130 1736 1708 6.3 0.739 0.754 9.9 3619 3692 2.6 0.758 0.720 8.8
140 1520 1576 7.2 0.666 0.638 11.0 3496 3545 2.7 0.672 0.640 10.0
150 1364 1452 8.1 0.584 0.535 12.5 3361 3382 2.9 0.601 0.577 11.4
160 1251 1337 8.9 0.496 0.445 14.7 3232 3208 3.0 0.535 0.529 13.0
170 1160 1232 9.6 0.406 0.368 17.9 3109 3023 3.2 0.473 0.494 14.9
180 1094 1138 10.2 0.323 0.304 22.6 2964 2830 3.4 0.42 0.469 17.0
190 1067 1056 10.4 0.242 0.254 30.2 2834 2630 3.6 0.375 0.454 19.3
200 1047 986 10.6 0.216 0.216 34.1 2535 2427 4.1 0.42 0.447 17.5
210 1003 930 11.1 0.212 0.192 35.1 2239 2221 4.7 0.467 0.446 16.0
220 962 888 11.7 0.192 0.180 39.5 1999 2015 5.3 0.487 0.449 15.5
230 917 860 12.8 0.175 0.182 44.3 1810 1811 6.0 0.467 0.454 16.4
240 847 849 15.0 0.181 0.197 44.1 1552 1611 7.1 0.492 0.460 15.8
250 754 854 18.8 0.191 0.224 43.3 1323 1418 8.4 0.507 0.465 15.6
260 1134 1232 10.0 0.515 0.467 15.6
270 992 1057 11.6 0.498 0.464 16.4
280 899 893 13.1 0.458 0.455 18.3
290 811 744 14.9 0.432 0.438 19.8
300 732 612 17.0 0.414 0.410 21.3

parameters has the best fits for each of the samples, whatever the duration and the
threshold. Its density function is:
X

1
f ( x, A1, a 2 ) x ~a2- ~)eA1, (3a)
IAIIA2F( a2)
where A1 and A2 are related to the two first moments, i.e:
A1 = or2/m, (3b)
and:
A2 = ( m / ~ r ) 2. (3c)
The S(r) probability is then estimated by:

p = prob [S < S ( r ) ] --- fs(~)f( x,A1,A2)dx.


"O
(4)

A1 and A2 parameters are estimated by the maximum likelihood method (see Table
1). Fig. 4a,b shows A1 and A2 versus r. It can be observed that: (i) for sample $48, A1
190 L. Neppel et al. /Atmospheric Research 45 (1997) 183-199

- 48h - - 24h -

2.5
4500 2.5
4000 I =2
35O0 2
3OOO 2000 ~
1.5 1.5
2500
< 2O00 1
1
1500 1000
1000 0.5 500 0.5
500
0 0 0
50 100 150 200 250 300 5O 100 150 200 250
x(mm)

(4a) (4b)

Fig. 4. A1 and A2 evolution in function of the threshold r for duration 24 h (b) and 48 h (a).

has a maximum around 7-= 70 mm and then decreases, for $24, it decreases for the
whole range of r; and (ii) for each sample $24 and $48, A2 is decreasing as far as
7-= 180 m m and should be considered as constant for greater r values.
The heterogeneity of the sample, from a meteorological point of view, may explain
those last observations. Tourasse (1981), Llasat and Puigcerver (1992), Benech et al.
(1993) have pointed out various groups of meteorological conditions: purely convective
ones with one storm cell or more; frontal and (or) orographic ones inside which one or
more convective cells may appear. Isolated summer storms, and mesoscale convective
systems (MCS), belonging to the first group, present quite stationary cells with heavy
rainfalls. Their trails over the rain gauge network, when integrated over l- or 2-day
duration have a smaller extent than the events from the second group (known as
~Cevenols~r events in the region). Moreover, according to the model of Austin and
Houze (1972), the areas including the storm cells are evolving more slowly than the
convective cells. As a consequence, the S(r) areas related to the first group should be
less scattered. Thus, one could explain why the coefficient of variation (CV) of the
samples S(7-) is increasing with z, while the A2 pattern should result from Eq. 4c.
However, the A 1 pattern cannot be easily justified by the Eq. 4b, and it appears to be
the product of an increasing term, i.e., the variance, and a decreasing term, i.e., the
inverse of the mean value of the samples S(7-).
A generalised model of Eq. (3a) can be proposed for thresholds r varying from 50
m m to 300 mm, with the A l ( r ) and A2(7-) parameters given by:
AI(~-) = 3.42 × 10-4"/-3 - - 0.216z 2 + 24.87-+ 3365, (5a)
A 2 ( z ) = 2.93 × 10-77- 3 + 2.053 × 10-47-2 - - 4.735 × 10-27- + 4.04, (5b)
for 6t = 48 h, and:
Al(7-) = 1.3810-47- 3 - 1.66 × 10-27- 2 - 16.327-+ 3806, (5c)
A2(7-) = 6.54 × 10-57- 2 + 2.92 × 1 0 - 2 r - 3.45, (5d)
for 6 t = 24 h.
The mean relative error when estimating A1 and A2 from Eq. (5a) to Eq. (5d) stays
L. Neppel et al. /Atmospheric Research 45 (1997) 183-199 191

Table 2
Arrangement of the fitting error for the surfaces quantiles estimation with the ~generalised~ method
AI A1 A1 A I + ~A1 A I + ~A1 A I + ~AI AI-~A1 AI-~AI AI-~AI
A2 A2-~A2 A 2 + ~A2 A2 A2-~A2 A 2 + ~A2 A2 A2-~A2 A 2 + ~A2

below 4 and 5.7%, respectively. If one assumes a normal distribution for these errors, at
a 5% confident limit, they remain under 11% for both parameters, whatever the duration.
However, the gamma law being sensitive to A2, one could wonder how these errors
modify the frequency estimates.
Such a sensitivity has been analysed. It is presented only for sample $48, for 6
isohyets 50(50)300 mm, but the methods and results still are the same for sample $24.
Area quantiles are estimated for various return periods, T, following two methods. First,
directly from the fitting of a Gamma function to the S(~') sample, using Eqs. (3a) and
(4). The confident intervals of the quantiles are also estimated. It is called the ~direct>>
method. Second, from Eqs. (5a) and (5b), taking into account the fitting errors called
orAl and trA2. Return periods are then estimated from Eq. (4). It is called the
~generalised>> method.
The generalised method gives, for each return period, 9 quantiles according to the 9
arrangements of the fitting errors related to parameters A1 and A2, as shown in Table 2.
Fig. 5 shows the results for ~- equal to 50 and 300 ram. For the ~generalised>> method,
and for each return period the median quantile is presented. It is related with case 1 in
Table 2, the minimum and maximum quantiles resulting from arrangements 2 to 9.
Whatever the threshold, it can be observed that the median quantiles as estimated by
the ~generalised>> method are close to the ones given by the direct method (differences
between both of them are less than 7%). Moreover, the fluctuations of the quantiles
resulting from the ~generalised>> method are included in the confident limits of those
given by the ~direct>> one. For z equal to, and greater than 250 mm the differences
between the median quantiles of both methods becomes larger (5 to 16%) than for lower
thresholds but those of the ~generalised>> method are still inside the confident limits of
the direct one, for each return period. So it can be assumed that the <~generalised>> model

40000 50 m m 300 mm
3500 1 I ."
35000
300o! ~ .... ~ __ _ ...T,,
30000
2500
~. 25000
E20000 ~ 2000

15000 1500
10000 1 1000
5000 500
0 I 0
1 10 100 1 10 100
T (years) T (years)
r Direct ~ Generalized, I Direct X Generalizedll

Fig. 5. Isohyets area quantiles and their 5% confident limit estimated from the ~direct>~ and ~,generalised>>
method, for 2 thresholds 50 mm and 300 ram. Duration: 48 h.
192 L. Neppel et al. /Atmospheric Research 45 (1997) 183-199

is not significantly sensitive to estimation errors of parameters AI(~-) and A2(~-), even
if some uncertainties appear for r greater than 250 mm, that may be related to the
sample size at such thresholds.
However, it must be noticed that as the AI(~-) and A2(r) models are based on an
empirical choice, they may not represent the evolution of the parameters beyond the
studied range of the thresholds. Thus, the ~generalised>> method should not be used for
an extrapolation below 50 mm threshold and for thresholds greater than 250 m m / 2 4 h
and 300 m m / 4 8 h.

4. Influence of the evolution of the rain gauge network on the isohyets area
quantiles

The influence of the rain gauge density studies mainly deal with the spatial rainfield
average accuracy. Paturel et al. (1986) estimate a required density to obtain a given
accuracy of the spatial average, depending on the basin area, Morrissey et al. (1995) test
the influence on the spatial averages of the network geometry. At the scale of urban
hydrology, Niemczynowicz (1990) and Lei and Schilling (1993) proposed an optimal
density to guaranty a given accuracy level on the maximal runoff estimations. However,
to our knowledge, no study has tried to estimate the bias introduced in sampling by
space-time rain gauge density instationary, particularly when high selection thresholds
are considered.
4.1. Bias estimation

A simulation approach has been chosen: the rain gauge network has been considered
each year t between 1958 and 1993, and 10 4 rainy surfaces, for each year, t, and each
area, S, are simulated over the LR region. The number Nobs (S,t) of rainy surfaces
intercepted by the network, i.e., which affect at least one gauge, have been counted. The
interception percentage, K(S,t), is defined as follows:
Nobs(S,t)
K(S,t) = 104 , (6)
where ~'(S, At) denotes the K(S,t) average over period At, for a given rainy surface S.
Austin and Houze (1972), Hobbs and Locatelli (1978) and Gupta and Waymire
(1979) have identified the different rainy surfaces and their characteristics, related to an
observation scale. The elementary ones are the convective cells, governed by a strong
dynamic, and which generate the highest intensity (Felgate and Read, 1975; Niem-
czynowicz, 1987; Berndtsson et al., 1994). Nevertheless, when observed through a basin
window, the rainfield just represents the space-time integration of the different rainy
surfaces dynamics: all this organised structure disappears and is giving place to a chaotic
appearance. So, over a 24-h or 48-h duration, the assumption of a circular rainy area is
made, its radius varying from 1, 3, 5(5)30, 40, 50 and 60 km. As the spatial probability
distribution function of those events according to their spatial extent is unknown, the
strong assumption of a uniform distribution over the LR region is used for simulation.
Characteristics of the K(S,t) distribution over 36 yrs are shown in Table 3: K(S,t)
decreases exponentially with S, for a radius of 5 km (near 80 km 2) on average over
L. Neppel et al. /Atmospheric Research 45 (1997) 183-199 193

Table 3
Distribution of K(S,t) over a 36-yr observation period according to the rainy surfaces
Radius (km) 1 3 5 10 15 20 25 30 40 50 60
Area (km2) 3.1 28.3 78.5 314 706 1256 1962 2826 5024 7850 11304
Min. 1.9 17.2 40.3 60.5 81.2 93.3 96.9 98.6 99.5 100 100
Max. 3.1 23.9 52.9 75.4 90.5 98 99.5 100 100 100 100
Mean 2.5 21.3 49.1 71.5 88.3 96.6 98.9 99.8 99.9 100 100
Median 2.6 21.5 50.1 73 89,2 97.1 99.2 99.9 100 100 100
Quartile 1 2.4 21.0 49.2 71.7 88.7 96.7 99.1 99.9 100 100 100
Quartile 3 2.7 22.3 51.2 73.6 89.6 97.3 99.3 99.9 100 100 100
CV 0.11 0.08 0.06 0.05 0.03 0.01 0.01 0.004 0.001 0 0

1958-1993, only 50% of the rainy surfaces affecting the region are intercepted by the
network. On average, 90% of the rainy surfaces are intercepted if their area is at least
900 km 2.
For a given radius, the distribution of K(S,t) over the 36 yrs simulation is not very
scattered: the mean K(S, At) is representative of the interception percentage for a given
S rainy surface, during 1958-1993. A spherical exponential model Eq. (7) is used to fit
K(S, At), according to S (Fig. 6):
h'*(S,At) = ( 1 - e - S / ° ) , (7)
where h" *(S, At) is a ~'(S, P1) estimator and b the single parameter.
One could assume that with 10 4 simulations, the sample fluctuations are negligible,
so F,(S, At) could be interpreted as the probability of observing a rainy surface S, over
the period 1958-1993, at one gauge of the network at least.
4.2. Bias influence on the isohyets area return period
At the ~" threshold, the definition of the return period of an isohyet area S(~-) equal or
over A, estimated over a D observation period, is given by:
D
T [S(~'),A] - naobs(,r ) , (8)

1
//o~--Co-O- o- o- . . . . ~ . . . . . o
0.9

I~
0.8
0.7
0.6
0,5
0.4
0.3-
( --
0

-- --
Ks simulated

Spherical model
0.2
0,1

0
0 200 400 600 800 1000 1200 1400 1600 1800 2000

S (knm).

Fig. 6. Fitting of K(S, At) with a spherical model. Parameter b = 122.8.


194 L. Neppel et al. /Atmospheric Research 45 (1997) 183-199

where nAObS(~') is the number of isohyets area more than A knl 2, which are observed
with rain gauges.
But nAObS(T) is only a fraction of the ~reab~ number, nareal(r), of those surfaces
which occurred over the region. T[S(r), A] is biased, the debiased expression is given by
Eq. (9):

D
T'[S(r),A] nAreal(~.) . (9)

Table 4
Comparison between biased (S o) and debiased (Sun0) isohyets area quantiles (kin2), for duration of 24 h (b)
and 48 h (a). IC(95%) represent the quantiles 5% confident limit, expressed in % of them. E is the relative
difference between biased and debiased quantiles, expressed in % of the biased ones
T (yrs) A tmax : 48 h
250 300
mln mill

So IC Sunb E So IC Sun b E
(95%) (95%)
1 90 41 120 -25.8
2 566 41 576 - 1.7 90 > 100 127 - 29.9
3 930 54 937 - 1.0 230 > 100 249 -9.6
5 1430 57 1440 -0.7 450 60 466 - 3.6
10 2160 52 2174 -0.5 800 64 820 - 2.1
15 2600 57 2623 -0.5 1030 65 1042 - 1.5
20 2930 51 2944 -0.4 1200 67 1205 - 1.4
25 3180 52 3186 -0.2 1320 64 1329 -0.8
30 3400 52 3408 -0.4 1430 66 1439 -0.4
50 400 54 4027 -0.1 1740 63 1756 - 1.0
100 4800 65 4851 -0.2 2170 62 2189 - 0.9

T (yrs) Atmax: 24 h
150 200 250
mm mm mm

Sb IC Sun b E Sb IC Su. o E Sb IC Sun b E


(95%) (95%) (95%)
1 710 23 721 - 1.5 40 > 100 75 - 46.8
2 1428 33 1438 -0.7 246 84 262 -6.0 10 > 100 36 -72.6
3 1880 30 1888 -0.4 440 60 447 - 1.5 40 > 100 80 -50.1
5 2460 37 2479 - 0.7 730 63 740 - 1.4 150 > 100 178 - 15.9
10 3290 34 3301 -0.3 1200 67 1207 -0.6 390 > 100403 -3.2
15 3780 30 3793 -0.3 1500 67 1505 -0.3 560 > 100575 -2.5
20 4130 32 4145 -0.4 1720 66 1725 -0.3 690 > 100732 -5.8
25 4400 33 4370 0.7 1890 63 1919 - 1.5 800 > 100838 -4.5
30 4630 33 4646 -0.4 2040 63 2065 - 1.2 900 > 100928 -3.0
50 5270 36 5151 2.3 2450 62 2482 - 1.4 1170 > 1001194 -2.0
100 6140 39 6023 1.9 3040 70 2981 2.0 1560 > 100 1587 - 1.7
L. Neppel et al. / AtmosphericResearch 45 (1997) 183-199 195

Let k(A,~-) define the ratio between naobs(z) and nareal(~-). It has been shown that
the percentage of interception, here k(A,'r), is related to S, so it comes:

k(a,'r) = fA
oo

K( s ) L ( s / s > a ) d s , (10)

where f ~ ( s / s > A) is the gamma density probability of the S(~-) isohyets area truncated
at threshold A.
Remembering Eq. (7) and developing Eq. (10) with Eqs. (8) and (9) yields Eq. (11):

T'(A,T)=T(A,7) l-Cl-~A) , (11)

where c is an expression of the gamma law parameters fitted on S(~-) sample and of
parameter b:

c= b +AI(I-) ' (12)

and F'(is a gamma law repartition function with parameters (AI'(~-), A2(~-)) where:
bal( )
al'(~-) b+al(T) " (13)

Debiased isohyets area quantiles are estimated with Eq. (11), at the 24-h duration for
thresholds of 150(50)250 ram, and at 6t equal to 48 h for thresholds of 250 and 300
mm. Table 4 shows that for a 24-h duration, the bias only influence quantiles
corresponding to 1-,2- and 3-yr return periods: the relative error between biased and
debiased quantiles goes up to 73% of the debiased one. For a 48-h duration, the relative
error does not exceed 30% of the debiased quantile, and for a return period equal to or
over 2 yrs, it represents less than 10%.
Whatever the threshold and duration, the bias influence has little weight for return
periods equal to or over 10 yrs. Moreover, it appears that the bias should be negligible
compared with the quantiles 5% confident limits, which for the highest thresholds
exceed 100%.

5. Conclusion

The frequency of isohyets area has been studied from a sample of 93 extreme events
observed during a 36-yr period at the scale of the Languedoc-Roussillon region. A
two-parameter Gamma function can be fitted to each isohyets area sample, for thresh-
olds varying from 50 to 300 nun, and for a 24- or 48-h duration. However, several
points must be noticed.
The areal quantile, corresponding to a high value of isohyets, seems to be quite
important. For example, according to Fig. 7, the area of 250 and 300 mm isohyets
corresponding to a 10-yr return period are 2160 and 800 km 2, respectively, for 6t equal
48 h. On the one hand, for a given k event, the area Sk(1-) can be organised in several
196 L. Neppel et al. /Atmospheric Research 45 (1997) 183-199

- 24h -
25000

.'O"
~- 15000 . . . . . . . ;/

~" ~oooo : . . ~ ........ 2 ~ . ~ . . . ~ 0 ............

0 ~ - --m m ~" ~ 1" ~ ' m ' l , , + 1 t- I I [~

10 100

T (years)
- 48h -

- 2oo0o1 . . . . . . . . . . . . .
4,"
,~
i
. . . . .
OO ,~
o-=:~ ~ --
.1oo,-,,.,[
15000 ~ .... . z 4'" : . , ~ . . . . . . .~ -- -~ - 150 m m I
1 I jl'" ~ F ..,..~ ~ " - 200 mm[
10000 ~ ......... : . ; . ' . " ~ ..... A = , , " ~ ~ ~._ =x __.

1 10 100

T (years)

Fig. 7. Isohyets area quantiles for thresholds 50(50)250 m m and duration 24 h, and for thresholds 50(50)300
m m and duration 48 h.

distinct z threshold surfaces, but in this study, only the total area Sk(z) has been
considered. On the other hand, the selection method may explain such high values. As
shown above, the events durations are unknown, and the selected events may have lasted
2 days or less. On October 2nd and 3rd of 1988, at N~mes, the event duration was 6 to 7
h if we consider the N~mes hydrological system (about 45 km2), whereas the 6th and 7th
of November, 1982 event affected the whole region and lasted about 40 h. Now for the
30 events observed during more than one <<meteorological>> day (among 93), the
durations are just unknown at the present moment of the study. The maximal areas
correspond to 2-day selected events; their trails over the rain gauge network have been
made during an unknown duration, but less than 48 h. The ratio of quantiles isohyets
area of a 48- and 24-h duration varies from 1.3 to 20, it increases with the return period
and with the ~- threshold. This could be explained by the increasing development of the
convective cells in time, which are characterised by high dynamics and which generate
the highest intensities. On the contrary, rainy areas corresponding to lower intensities
evolve much more slowly, so their trails on the rain gauge network vary less with
duration than trails corresponding to the highest rainfall thresholds.
Over the 1958-1993 period, the rain gauges network was instationary. The bias
introduced on the isohyets area quantiles has been estimated. It has been shown that it is
negligible when compared to the quantiles 5% confident limits. Only the areas corre-
sponding to a less than a 5-yr return period are significantly affected. However, this
L. Neppel et al. / Atmospheric Research 45 (1997) 183-199 197

influence may be underestimated: as it was said above, for a given event k and a ~-
threshold, the area Sk(r) can be organised in several distinct rainy surfaces. Here, the
bias influence is tested on the whole area Sk(~-), or the proportion of rainy surfaces
intercepted by the rain gauges network decreases exponentially with the rainy surface
area. So if the distinct (thresholded surfaces are considered, the bias effects would
probably be higher than with the total area Sk(T).
It would be worthwhile making such a study on other areas in order to see if the
gamma law and the parameters evolution according to the rainfall thresholds are
invariable. Such an approach could be easily transposed to other sites provided that one
could estimate the rainy surfaces area either with a rain gauges network using a spatial
interpolator, or with a remote sensor like a meteorological radar. Nevertheless, one
should take care of the spatial interpolator choice when dealing with rain gauges data at
a smallest time step, for example hourly data: the rain field intermittency is no more
negligible and other interpolators like indicator kriging (Barancourt et al., 1992) may be
more suitable for delineating rainy surfaces.
Nevertheless, two main points still need to be investigated. The isohyets area
confident limit is unknown. However, kriging gives a variance of estimation of punctual
interpolations. Such an information should be used to assess the isohyet area accuracy.
The influences of rain gauge instationary is probably higher when the confident limits of
the isohyets area are conspired. On the other hand, an extension of rainfall risk
assessment to the other French Mediterranean regions, supported by the whole data
included in the (<pluvio>>database (since 1870) is under development. It should enable to
test the time step influences on the gamma law parameters according to ~-. From a
practitioners point of view, lower selection thresholds should allow to perform the
analysis of extreme rainfall lasting a few hours only, that may generate severe runoff
flooding on small urbanised catchments.

Acknowledgements

This research is supproted by the French National Scientific Research Center (CNRS)
in the frame of the National Program on the Natural Hazards (PNRN).
The authors would like to thanks the Direction Interr6gionale Sud-Est de M&6o-France
for allowing them to use the Pluvio data base and for its help in selecting the events, and
Marc Montgaillard from SIEE (Montpellier) for his assistance in the software manage-
ment.

References

Austin, P.M., Houze, R.A., 1972. Analysis of the structure of precipitation patterns in New England. J. Appl.
Meteorol. 11,926-935.
Barancourt, C., Creutin, J.D., Rivoirard, J., 1992. A method for delineating and estimating rainfall fields.
Water Resour. Res. 28 (4), 1133-1144.
Bartual-Garcia, R., Marco, J., 1990. A stochastic model of internal structure of convective precipitation in time
at a rain gauge site. J. Hydrol. 118, 129-142.
198 L. Neppel et aL /Atmospheric Research 45 (1997) 183-199

Benech, J.C., Brunet, H., Jacq, V., Payen, M., Riverain, J.C., Santurette, P., 1993. La catastrophe de
Vaison-la-Romaine et les violentes prcipitations de Septembre 1992: aspects m&~orologiques. La
M&~orologie 8 (1), 72-90.
Berndtsson, R., Jinno, K., Kawamura, A., Larson, M., Niemczynowicz, J., 1994. Some Eulerian and
Lagrangian statistical properties of rainfall at small space-time scales. J. Hydrol. 153, 339-355.
Bobee, B., Cavadias, G., Ashkar, F., Bernier, J., Rasmussen, P., 1993. Towards a systematic approach to
comparing distributions used in flood frequency analysis. J. Hydrol. 142, 121-136.
Cernesson, F., Lavabre, J., 1994. Recurrence des 6pisodes pluvieux extrSmes--approche par modrlisation.
CNFSH, 2 Juin 1994, Aix-en-Provence, 9 pp.
Cernesson, F., Lavabre, J., Masson, J.M., 1996. Stochastic model for generating hourly hyetographs. Atmos.
Res. 42, 149-161.
Cowpertwait, P.S.P., 1991. Further development of the Neyman-Scott clustered point process for modeling
rainfall. Water Resour. Res. 27 (7), 1431-1438.
Cowpertwait, P.S.P., OConnel, P.E., Metcaife, A.V., Mawdsley, J.A., 1996. Stochastic point process modeling
of rainfall: 2. Regionalisation and disaggregation. J. Hydrol. 175, 47-65.
Creutin, J.D., Andrieu, H., Krajewski, W.F., 1994. Prrvisions h trbs courte 6chance des intensit& pluvieuses.
Utilisation des informations des radars m&6orologiques. Congrbs de la SHF N~mes, 23~me J. Hydraulique,
tome 1: 595-599.
Croley, T.E., Eli, R.N., Cryer, J.D., 1978. Ralstom Creek hourly precipitation model. Water Resour. Res. 14
(3), 485-490.
Desbordes, M., Durepaire, P., Gilly, J.C., Masson, J.M., Maurin, Y., 1989. 3 Octobre 1988, inondations sur
N3mes et sa rrgion: Manifestation, Causes et Consrquences. Lacour, N'/mes, 95 pp.
Desbordes, M., Masson, J.M., 1992. Prrcipitations extrSmes dans le sud de l'Europe. VIII~me J. Hydrol.
Orstom, pp. 153-164.
Duband, D., 1994. Pour une meilleur prise en compte de l'information hydrom&6orologique historique, sur les
crues importantes des bassins suprrieurs de certaines rivi~res h risques. Congrbs de la SHF Nimes, 23~me
J. Hydraulique, tome 1: 137-144.
Econopouly, T.W., Davis, D.R., Woolisher, D.A., 1990. Parameter transferability for a daily rainfalls
disaggregation model. J. Hydrol. 118, 209-228.
Felgate, D.G., Read, D.D., 1975. Correlation analysis of the cellular structure of storms observed by rain
gauges. J. Hydrol. 24, 191-200.
Fontaine, T.A., Potter, K.W., 1989. Estimating probabilities of extreme rainfalls. J. Hydraulic Eng. 115 (11),
1562-1575.
Foufoula-Georgiou, E., 1989. A probabilistic storm transposition approach for estimating exceedance probabil-
ities of extreme precipitation depths. Water Resour. Res. 25 (5), 799-815.
Franchini, M., Helmlinger, K.R., Foufoula-Georgiou, E., Todini, E., 1996. Stochastic storm transposition
coupled with rainfall-runoff modeling for estimation of exceedance probabilities of design floods. J.
Hydrol. 175, 511-532.
Gupta, V.K., Waymire, E., 1979. A stochastic kinematic study of subsynoptic space-time rainfall. Water
Resour. Res. 15 (3), 637-644.
Hershenhorn, J., Woolhiser, D.A., 1987. Disaggregation of daily rainfalls. J. Hydrol. 95, 299-322.
Hobbs, P.V., Locatelli, J.D., 1978. Rainbands, precipitation cores and generating cells in a cyclonic storm. J.
Atmos. Sci. 35, 230-241.
Hosking, J.R.M., Wallis, J.R., 1986. The value of historical data in flood frequency analysis. Water Resour.
Res. 15 (3), 1606-1612.
Jacq, V., 1995. Inventaire des situations h pr6cipitations diluviennes en Languedoc-Roussillon/Paca/Corse.
Ministbre de l'Environnement, Service Central d'Exploitation de la Mrtrorologie, Direction Interrrgionale
Sud-Est, 190 pp.
Journel, A.G., Huijbregts, Ch.J., 1978. Mining Geostatistics. Academic Press, New York, 600 pp.
Juvanon du Vachat, R., 1994. Panorama de la pr6vision num~rique h ~chelle fine ou ~t domaine limitC La
M6t6orologie 8 (6), 31-47.
Lei, J., Schilling, W., 1993. Requirements of spatial raindata resolution in urban rainfall runoff simulation. 6th
Int. Conf. Urban Storm Drainage, IARH-IAWQ, Niagara Falls, Sept. 12-17, 1993, 447-452.
Llasat, M.C., Puigcerver, M., 1992. Pluies extrSmes en catalogne. Hydrol. Continentale 7 (2), 99-115.
L. Neppel et al./ Atmospheric Research 45 (1997) 183-199 199

Masson, J.M., 1991. Un probl~me parmi d'autres dans l'analyse des distributions des variables hydrologiques:
les Horsains (Outliers). Statistique appliquEe, Seminfor. Orstom, Paris, 10 pp.
Morrissey, L.M., Maliekal, J.A., Greene, J.S., Wang, J., 1995. The uncertainty of simple spatial averages using
rain gauge networks. Water Resour. Res. 31 (8), 2011-2017.
Neppel, L., Desbordes, M., Masson, J.M., 1997. CaractErisation de l'alEa climatique pluvieux en region
mEditerranEnne: Etude statistique de surfaces pluvieuses. Rev. Sci. Eau (accepted for publication).
Nguhyen, T.P.T., Bois, P., Villasenor, J.A., 1993. Simulation in order to choose a fitting method for extreme
rainfall data. Atmos. Res. 30, 13-36.
Niemczynowicz, J., 1987. Storm tracking using rain gauge data. J. Hydrol. 93, 135-152.
Niemczynowicz, J., 1990. Necessary level of accuracy in rainfall input for runoff modeling. 5th Int. Conf.
Urban Storm Drainage, IARH-IAWQ, Osaka, July 23-27, 1990, pp. 593-601.
Paturel, E., Desbordes, M., Masson, J.M., 1986. Evaluation de l'influence de la densitE des r6seaux
pluviomtEriques sur la d&ermination des lames pr6cipitEes. Laboratoire d'Hydrologie Math6matique,
Universit6 Montpellier II, note interne LHM 14-1986, 44 pp.
Rodriguez-lturbe, I., Cox, D.R., Isham, F.R.S., Isham, V., 1987. Some models for rainfall based on stochastic
point processes. R. Soc. London A410, 269-288.
Rossi, F., Fiorentino, M., Versace, P., 1984. Two-component extreme value distribution for flood frequency
analysis. Water Resour. Res. 20 (7), 847-856.
Shah, S.M.S., OConnel, P.E., Hosking, J.R.M., 1996. Modeling the effects of spatial variability in rainfall on
catchement response: 1. Formulation and calibration of a stochastic rainfall field model. J. Hydrol. 175.
67-88.
Tourasse, P., 1981. Analyses spatiales et temporelles de pr6cipitations et utilisation op6rationnelle dans un
syst~me de pr~vision de crues--application aux r6gions C~venoles. Th6se de Docteur Ing~nieur, USMG-
INPG, Grenoble, 189 pp.
Waymire, E., Gupta, V.K., 1981. The mathematical structure of rainfall representations: a review of the
stochastic rainfall models. Water Resour. Res 17 (5), 1261-1272.

You might also like