Sunteți pe pagina 1din 15

HYDROLOGICAL PROCESSES Hydrol. Process. 15, 2281 2295 (2001) DOI: 10.1002/hyp.

257

Application of a data-based mechanistic modelling (DBM) approach for predicting runoff generation in semi-arid regions
S. Mwakalila,1 * P. Campling,1 J. Feyen,1 G. Wyseure2 and K. Beven3
2 1 Institute for Land and Water Management (ILWM), K.U.Leuven, Belgium Faculty of Agricultural and Applied Biological Sciences, K.U.Leuven, Belgium 3 Institute of Environmental and Natural Sciences, Lancaster University, UK

Abstract:
This paper addresses the application of a data-based mechanistic (DBM) modelling approach using transfer function models (TFMs) with non-linear rainfall ltering to predict runoff generation from a semi-arid catchment (795 km2 ) in Tanzania. With DBM modelling, time series of rainfall and streamow were allowed to suggest an appropriate model structure compatible with the data available. The model structures were evaluated by looking at how well the model tted the data, and how well the parameters of the model were estimated. The results indicated that a parallel model structure is appropriate with a proportion of the runoff being routed through a fast ow pathway and the remainder through a slow ow pathway. Finally, the study employed a Generalized Likelihood Uncertainty Estimation (GLUE) methodology to evaluate the parameter sensitivity and predictive uncertainty based on the feasible parameter ranges chosen from the initial analysis of recession curves and calibration of the TFM. Results showed that parameters that control the slow ow pathway are relatively more sensitive than those that control the fast ow pathway of the hydrograph. Within the GLUE framework, it was found that multiple acceptable parameter sets give a range of predictions. This was found to be an advantage, since it allows the possibility of assessing the uncertainty in predictions as conditioned on the calibration data and then using that uncertainty as part of the decision-making process arising from any rainfall-runoff modelling project. Copyright 2001 John Wiley & Sons, Ltd.
KEY WORDS

data-based mechanistic modelling approach; transfer function models; Generalized Likelihood Uncertainty Estimation; parameter sensitivity and predictive uncertainty

INTRODUCTION Rainfall-runoff models are important tools for water resource planning, development and management. However, limitations of both model structures and the data available on parameter values, initial conditions and boundary conditions, generally make it difcult to apply any hydrological model without some form of calibration or conditioning of model parameter sets. The data-based mechanistic (DBM) modelling addressed in this study is based on the concepts of letting the data suggest an appropriate model structure compatible with the inputoutput data available, but then evaluating the resulting model to see if there is a mechanistic interpretation that might lead to insights not otherwise gained from modelling based on theoretical reasoning. These concepts have been applied successfully in humid climates, e.g. see Young and Beven (1994), Young (1998), and Beven (2001). Similarly, in the application of an IHACRES model the data are also allowed to suggest the form of the transfer function used for ow routing rather than specifying a xed structure beforehand (see, Jakeman et al., 1990, 1993; Jakeman and Hornberger, 1993; Schreider and Jakeman, 1996; Post and Jakeman, 1996; Sefton and Howarth).
* Correspondence to: Dr. S. Mwakalila, Institute for Land and Water Management Katholieke Universiteit Leuven, Vital Decosterstraat 102, B-3000 Leuven, Belgium. E-mail: mwakalila@hotmail.com

Copyright 2001 John Wiley & Sons, Ltd.

Received 28 June 2000 Accepted 16 January 2001

2282

S. MWAKALILA ET AL.

The results generally show that the parallel transfer function is a suitable structure for rainfall-runoff simulation at the catchment scale, with one fast ow pathway and one slower ow pathway. The fast ow pathway provides the major part of the predicted storm hydrograph, and the slower pathway the major part of the recession discharge between storm periods. This study focuses on extending the application of these concepts to semi-arid catchments and also to assess the parameter sensitivity and discharge prediction limits of the identied model structure. Compared with humid climate regions, in semi-arid regions that cover a large part of the world there are a lot of problems in predicting runoff generation. In general, the two interrelated underpinning problems in rainfall-runoff modelling in such regions are due to scarcity of data and data uncertainties. The success of any catchment rainfall-runoff model depends critically on the data available to set it up and drive it. Nearly all rainfall-runoff models require data on some combination of rainfall, discharge, evapotranspiration and catchment morphology. The basic sources of error and uncertainty in each of these data types and methods to account for these uncertainties are discussed in detail in Melching (1995). He noted that the primary uncertainties in using point-rainfall measurements to describe the true rainfall input for a catchment are: depth measurement error in the gauge due to equipment malfunction; representativeness of groundlevel precipitation at the gauging point; gauge location; the areal-mean rainfall; gauge-network areal-mean rainfall versus true areal-mean rainfall; rainfall spatial variability; rainfall temporal variability; lack of synchronization between time clocks for rain and stream gauges; and lack of synchronization between time clocks for the various rain gauges in the catchment. The availability of discharge data is important for the model calibration process. Discharge data are, however, generally available at only a small number of sites in semi-arid regions. It is also an integrated measure, in that the measured hydrograph will reect all the complexity of ow processes occurring in the catchment. However, the discharge data available are generally point data and may not be error free. If any model is calibrated using data that are in error, the effective parameter values will be affected and the predictions for other periods that depend on the calibrated parameter values will be affected. This will be an additional source of uncertainty in the modelling. The main objective of this paper therefore, is to address the application of a DBM modelling approach to predict runoff generation from a semi-arid catchment. Furthermore, the paper employs a Generalised Likelihood Uncertainty Estimation (GLUE) methodology using Monte Carlo simulation (MCS) as key component to assess the model parameter sensitivity and predictive uncertainty. This type of technique has been used on several occasions (e.g. Beven and Binley, 1992; Beven, 1993; Romanowicz et al., 1994; Freer et al., 1997; Franks et al., 1998; Cameron et al., 1999), but not in the context of semi-arid regions.

METHODOLOGY The methodology in this study has three distinct stages. At the rst stage a DBM approach was used to suggest an appropriate transfer function model (TFM) structure compatible with the available set of daily rainfallrunoff data. At this stage it is the idea that the data should be allowed to indicate which model structure is most appropriate rather than specifying a structure beforehand. Therefore model structures were evaluated by looking at how well the model tted the data, and how well the parameters of the model were estimated. The second stage involves the separation of the slow ow component from the total discharge in order to get a hydrologically more realistic model structure. At the third stage the GLUE methodology approach using MCS as a key component was employed for performance evaluation of the identied model structure for the purpose of parameter sensitivity analysis and uncertain estimation. Data used The data used in this study were rainfall and streamow records available on a daily basis for 197080 period for the Great Ruaha catchment at Salimwani gauging station (IKA 8A) in Tanzania. The study area
Copyright 2001 John Wiley & Sons, Ltd. Hydrol. Process. 15, 2281 2295 (2001)

PREDICTING RUNOFF GENERATION

2283

covers an area of about 795 km2 . The gauging station is located at 34 07 E longitude and 8 54 S latitude. It is a sub-basin of the Great Ruaha River Basin, which is draining an area of about 68 000 km2 . The annual mean temperature varies from about 18 C at the higher altitudes to about 28 C at the lower and drier part of the basin. The rainfall regime in the basin is typical of the unimodal type, with a single rainy season from November through May and no rainfall during the rest of the year. As the main process causing rainfall is convection, the rainfall is highly localized and with the incoming wind direction and the topography being the main determining factors of spatial patterns. The mean annual rainfall ranges from 800 mm in the lower part of the catchment to 1000 mm in the highlands. Owing to the generally high potential evaporation rates (14001800 mm year 1 ) compared with the yearly rainfall, most of the catchment has an annual rainfall decit. The runoff patterns in the basin correspond closely to the unimodal rainfall patterns prevailing in the whole catchment. The streams start rising in NovemberDecember, experience a maximum ow in MarchApril and have their recession period from May to October. As rainfed cultivation is very unreliable in the area due to the low and unreliable rainfall, farmers in the area depend very much on irrigation. The major part of irrigation takes place downstream of the gauging station. In all types of irrigation, rice is the dominant crop, which is grown during the rainy season (DANIDA, 1995). Initial evaluation of general linear TFM In this stage of the methodology, the general linear TFM was evaluated to identify an appropriate model structure. Within the DBM approach a non-linear rainfall lter function [Equation (1)] was used to produce an effective rainfall, which was then related to discharge using a generalized linear transfer function [Equation (2)]. U t / Rt Q t n 1 where Ut is the effective rainfall, Rt is the rainfall input at time t, Qt is the discharge; parameter n controls the non-linearity of fast runoff generation (if n D 0, the rainfall lter is linear as for subsurface recharge). Here, the current discharge was used as an index of the contribution of antecedent moisture status of the catchment on runoff generation. Therefore, the interpretation of Rt Qt n can be referred to as a contributing area function. The general form of the discrete time linear transfer function that forms the basis of the TFM package can be presented as follows: b0 C b1 z 1 C C bM z M Qt D Ut 2 1 a1 z 1 a2 z 2 C C aN z N where Qt is a simulated discharge, Ut is the effective rainfall, z is a backward difference operator, a and b are coefcients treated as parameters, such that parameter a is related to the mean residence time of fast ow pathway and parameter b is a gain parameter that scale the differences in total volumes of input and output; M and N represent the total number of b and a parameters respectively, and is the time delay. The derivation of Equation (2) can be found in Beven (2001). Parameter estimation is carried out using a simplied rened instrumental variable (SRIV) technique which has been shown to be robust to data errors. For more detail see Young (1984), Young and Beven (1994) and Beven (2001). The higher the model order (i.e. the greater the number of a and b coefcients), the larger the standard errors of the estimated parameter values will tend to be. Very large standard errors on the coefcients are an indication that a model is over-parameterized. One way of checking for over-parameterization is by looking at how well the coefcients of the model are estimated. The aim in this study, is to nd a model structure that gives a good t but is parsimonious in having a small number of coefcients. Therefore, using a daily rainfall-runoff data set an appropriate TFM structure was identied, by nding the best values of (N, M, ) and the corresponding coefcients.
Copyright 2001 John Wiley & Sons, Ltd. Hydrol. Process. 15, 2281 2295 (2001)

2284

S. MWAKALILA ET AL.

Performance evaluation of TFM structures Underlying the model identication step is the idea that the data should be allowed to indicate which model structure is most appropriate, rather than specifying a structure beforehand. There may be many models giving an acceptable t to the data. In this study, therefore, model structures were evaluated in terms of the following criteria: NashSutcliffe efciency R2 , which indicates how well the model ts the data; Young Information Criterion (YIC), which combines a measure of goodness of t with a measure of how well the parameters are estimated. R2 , introduced as the efciency measure for rainfall-runoff modelling by Nash and Sutcliffe (1970), is statistically known as the coefcient of determination indicating the proportion of the variance in the data explained by the model, and can be dened as: R2 D 1
e o 2 2

where o 2 is the variance of the observed output data calculated over all time steps used in tting the model and e 2 is the variance of the differences between observed and simulated outputs at each time step. As the model t improves, the value of R2 will approach unity. If the model is no better than tting the mean of the observed outputs e 2 D o 2 , the value of R2 will be zero or less. The YIC is dened as: 2 1 e 2 Pii YIC D ln C ln 4 e 2 k k i o where i , i D 1 . . . k, are the model coefcients, and Pii is the ith diagonal element of a scaled parameter covariance matrix. From its denition, the rst term of YIC is simply a relative logarithmic measure of how well the model explains the data: the smaller the model residuals the more negative the term becomes. The second logarithmic term, on the other hand, tends to become large (less negative) when the model is over-parameterized and the parameter estimates are poorly dened. Consequently, as whole, the criterion attempts to identify a model that explains the data well but with the minimum of statistically well-dened (low-variance) parameters. Separating a slow ow component from total discharge Since it was learnt that the general linear TFM is underestimating the long-term recession ows, therefore, the slow ow component was separated off from total discharge in order to get a hydrologically more realistic model structure. Therefore, to allow for non-linearity the linear storageoutow relationship was generalized by an exponential reservoir model as presented by Beven et al. (1995): Qb D Qo e
s/m

where Qb is the outow from the saturated zone considered as baseow, Qo is the discharge when recession of streamow commences, s is the local storage decit and m is a model parameter considered as the scaling depth of the effective catchment soil prole. Solution of Equation (5) for a pure recession in which inputs are assumed to be zero shows that discharge has an inverse or rst-order hyperbolic relationship to time as: 1 1 t D C Qb Qo m 6

A plot of 1/Qt against time t should plot as straight line with a slope of 1/m, where Qt is the recession ow at time t. Therefore, Equation (5) was used for separating the slow component based on the following major steps.
Copyright 2001 John Wiley & Sons, Ltd. Hydrol. Process. 15, 2281 2295 (2001)

PREDICTING RUNOFF GENERATION

2285

(1) The catchment average storage decit before each time step was updated by subtracting the unsaturated zone recharge and adding the baseow (slow ow component) calculated for the previous time step, thus: StC1 D St C Qo e
St /m

cRt

where Rt is the observed rainfall at time t and parameter c represents the proportion of rainfall that percolates as recharge to the groundwater. (2) If an initial discharge Q tD0 is known and assumed to be only the result of drainage from the saturated zone, Equation (5) can be inverted to give a value for S at time t D 0 as: SD m ln QtD0 Qo 8

Equations (7) and (8), therefore, form the basis for separating the baseow from the total discharge during storm events. Then, given the time series of discharge Qt and rainfall Rt at time t, the time series of direct runoff (fast ow component) Qd was calculated as: Qd D Qt Performance evaluation of identied model structure The performance evaluation of the identied model structure was based on the framework of the GLUE methodology for the purpose of parameter sensitivity analysis and predictive uncertainty estimation. The main objective here was to examine those parameters to which the model simulation results were most sensitive and to assess the probability of simulated discharge being within a certain prediction interval. The prediction interval was dened by 5 and 95% limits (i.e. a 90% probability that the predicted discharge value lies within the interval). The main procedure was based on the following major steps: (1) determining a feasible parameter range from initial analysis of recession curves and calibration of TFM; (2) using a MCS method to choose parameter values from uniform distributions spanning specied ranges of each parameter; (3) using a likelihood value to divide acceptable simulations from unacceptable simulations; (4) normalizing likelihood values for the parameter sets that yield acceptable simulations, such that the sum of the normalized likelihood values equals unity; (5) using the same generated parameter sets and associated likelihood value and model outputs to assess parameter sensitivity and predictive uncertainty. The modelling efciency R2 in Equation (3) was used as a likelihood measure for determining acceptable simulations. With the MCS technique, 5000 runs of the model were made with different randomly chosen parameter sets. The feasible range of parameters to be sampled was based on an initial evaluation of the TFM and recession ow analysis. Those simulations with R2 > 0 5 were retained, those that did not meet this performance criterion were deleted. Finally, 4000 simulations were retained as behavioural. In this study, therefore, R2 D 0 5 was considered as a behavioural threshold indicating the proportion of the variance in the data explained by the model. This value was considered as a minimum value to accept the simulations (Beven, 2001). In the GLUE procedures a prior likelihood estimate can be updated with a new (posterior) likelihood measure calculated from the prediction of a new set of observations. The idea here is to check if the model, which performs well during the conditioning or calibration period, will continue to perform well in other periods. If this is not the case then its combined likelihood measure will be reduced. The Bayes equation
Copyright 2001 John Wiley & Sons, Ltd. Hydrol. Process. 15, 2281 2295 (2001)

Qo e

St /m

2286

S. MWAKALILA ET AL.

was used here to combine likelihood measures, the type of combination using the Bayes equation may be expressed in the form: Lo i L i jY 10 Lp i jY D C where Lo i is the prior likelihood of the ith parameter set, L i jY is the likelihood calculated for the current evaluation given the set of observations Y, Lp i jY is the posterior likelihood and C is a scaling constant to ensure that the cumulative posterior likelihood is unity. The special characteristic of the Bayes equation is its multiplicative operation, if any evaluation results in a zero likelihood, the posterior likelihood will be zero regardless of how well the model has performed previously. This may be considered as an important way of rejecting non-behavioural models: it may cause a re-evaluation of the data for that period or variable; it may lead to the rejection of all models. The GLUE methodology used here is one technique for conditioning of model parameter sets in the face of rejecting the idea that there is an optimum parameter set in favour of the concept of the equinality of different models and parameter sets. Parameter interactions and non-linearity in the model responses are handled implicitly. Errors in input data and the observation data are also handled implicitly. Thus the likelihood measure reects the ability of a particular model to predict a particular series of observations (which may not be error free) given a particular set of inputs (which may not be error free). There is thus an implicit assumption that, in prediction, error structures will be similar in some broad sense to those in the evaluation period.

RESULTS AND DISCUSSIONS Initially, models were calibrated for ve periods of 2 years in length within the decade (197080). The second sub-period, 197273, was found to give the best calibrations. Therefore, the time series of daily rainfall and discharge (for the 197273 observation period) were allowed to suggest an appropriate model structure within the general transfer function modelling framework. The results in Table I indicate that a rst-order (1, 1, 0) transfer function is an appropriate model structure with parameter n in Equation (1) ranging between 05 and
Table I. DBM model results of total observed rainfall and discharge in 197273 period Model ordera 1, 1, 1, 1, 1, 1, 1, 2, 0 1 2 2 R2 079 078 077 078 079 078 079 079 YIC 873 843 841 602 473 383 261 170 a1: a1: a1: a1: a parameters 09592(00017) 09542(00022) 09561(00021) 09610(00020) b0: b0: b0: b0: b1: b0: b1: b0: b1: b0: b1: b0: b parameters 00202(00008) 00223(00010) 00213(00009) 00473(00043) 00281(00044) 00350(00042) 00133(00044) 00096(00042) 00129(00044) 00080(00043) 00199(00061) 00239(00035)

1, 2, 1 1, 2, 0 2, 2, 0 2, 1, 1

a1: 09554(00026) a1: 09335(00024) a1: a2: a1: a2: 06958(02429) 02471(02316) 08573(01549) 00937(01482)

Model order D N, M, (N is the number of a parameters, M is the number of b parameters, is a delay).

Copyright 2001 John Wiley & Sons, Ltd.

Hydrol. Process. 15, 2281 2295 (2001)

PREDICTING RUNOFF GENERATION

2287

80 Rainfall (mm) 60 40 20 0 0 200 400 Time (days) Discharge (mm/day) 12.00 Qo 10.00 Qs 6.00 4.00 2.00 0.00 0 200 400 Time (days) 10.00 Error 0.00 0 10.00 Time (days)
Figure 1. Rainfall, observed discharge Qo and simulated discharge Qs and error

600

800

600

800

100

200

300

400

500

600

700

800

10 for a non-linear rainfall lter function. However, this model structure is underestimating the long-term recession ows, as depicted in Figure 1. As can be noted from Table I, it seems a second-order transfer function model (2, 2, 0 and 2, 1, 1) does better in terms of R2 and YIC, presumably because errors are dominated by peaks during the wet season and are relatively small in the dry period recession (see Figure 1). This indicates that, in order to get a hydrologically more realistic model, the slow ow component should be separated off rst before evaluating the TFM structures. The DBM model results after separating the slow ow component are presented in Figure 2 and Table II, and the synthesized total and baseow hydrograph is given in Figure 3. A graphical comparison between the observed and simulated fast ow time series shows a good t. This justies that the slow ow component should be separated off rst before calibrating the TFM. In connection with Figure 3, this implies that, from the general linear TFM [Equation (2)], the effective rainfall can be linearly related to the fast ow component of the hydrograph as presented in the Equation (11): Qs t D aQs t
Copyright 2001 John Wiley & Sons, Ltd.
1

C bRt Qd t

11
Hydrol. Process. 15, 2281 2295 (2001)

2288

S. MWAKALILA ET AL.

Table II. DBM results after separating the slow ow Qb component Model ordera 1, 1, 1, 1, 1, 1, 1, 2, 0 1 2 2 R2 092 091 090 090 091 091 YIC 1096 1074 1046 565 523 377 a1: a1: a1: a1: a parameters 09580(00009) 09559(00010) 09550(00011) 09565(00012) b0: b0: b0: b0: b1: b0: b1: b0: b1: b parameters 00257(00005) 00268(00006) 00271(00006) 00378(00034) 00115(00035) 00342(00032) 00080(00033) 00227(00031) 00036(00032)

1, 2, 1 1, 2, 0

a1: 09568(00011) a1: 09570(00010)

a Model

order D N, M, (N is the number of a parameters, M is the number of b parameters, is a

delay).

Discharge (mm/day)

10.00 8.00 6.00 4.00 2.00 0.00 0 200 400 Time (days)
0 Figure 2. Observed Qd and simulated Qd fast ow component

Qd (mm) Qd(mm)

600

800

Discharge (mm/day)

12.00 10.00 8.00 6.00 4.00 2.00 0.00 0 200 400 Time (days) 600

Qt(mm) Qb(mm)

800

Figure 3. Total discharge Qt and slow ow component (baseow) Qb

Copyright 2001 John Wiley & Sons, Ltd.

Hydrol. Process. 15, 2281 2295 (2001)

PREDICTING RUNOFF GENERATION

2289

where Qs t is a simulated fast ow component, Rt is the observed rainfall, Qd t is the estimated (from Equation (9)) fast ow component, t is time, and a, b, and n are optimized model parameters. The slow ow component can be represented by Equation (12) as: Qb t D Qo e
St /m

12

where Qb t is a simulated slow ow component, St is the catchment storage decit estimated from Equations (7) and (8), and Qo , m and c are optimized model parameters. The identied nal model structure can be presented as shown in Figure 4, which indicates that a parallel model structure is appropriate with a proportion of the runoff being routed through a fast ow pathway and the remainder through a slow ow pathway. As an example, the synthesized model performance of behavioural simulations for feasible parameter ranges using the framework of the GLUE approach is shown in Table III. Table III, shows that there may be many representations of a catchment that may be equally valid in terms of their ability to produce acceptable simulations of the available data. Multiple acceptable parameter sets give a range of predictions. This may actually be an advantage, since it allows the possibility of assessing the uncertainty in predictions
Effective rainfall Fast Flow Pathway (Surface Flow) Qs(t) = aQs(t1) + bRt (Qd(t))n

Rainfall Rt

Nonlinear rainfall filter function

Discharge Q t

Recharge

Slow Flow Pathwya (Base Flow) Qb(t) = Qoe-st/m

Figure 4. Final block diagram of the identied model structure

Copyright 2001 John Wiley & Sons, Ltd.

Hydrol. Process. 15, 2281 2295 (2001)

2290

S. MWAKALILA ET AL.

Table III. A sample of identied model performance after using a GLUE approach Set of parameters c 040 033 033 040 026 028 033 034 042 048 041 009 036 054 020 051 042 040 014 035 029 026 020 018 009 047 053 033 042 045 020 049 044 026 039 011 050 044 042 m 5804 4989 5014 6233 5130 5705 5447 4917 4527 4970 5870 6038 5491 5393 6358 5044 5349 6189 5375 5686 4920 5712 4870 4891 6032 4616 6064 5032 6306 5688 6055 4746 4745 4566 4716 5633 4684 5338 4628 Qo 462 440 446 715 461 567 741 626 677 405 523 451 488 521 416 747 606 477 797 429 629 718 793 627 489 786 775 663 710 697 510 549 619 460 718 767 675 724 671 n 096 084 085 084 077 093 073 069 082 099 063 055 091 088 075 086 063 082 065 075 090 060 064 088 056 071 085 075 057 073 092 061 086 072 100 055 095 060 077 a 0742 0770 0777 0768 0781 0686 0739 0855 0796 0575 0807 0840 0807 0767 0767 0846 0877 0815 0869 0809 0653 0826 0780 0782 0801 0708 0848 0803 0894 0733 0675 0846 0526 0752 0883 0880 0815 0886 0670 b 0043 0044 0040 0041 0046 0049 0049 0034 0031 0047 0041 0049 0027 0037 0044 0029 0031 0028 0032 0032 0046 0037 0044 0033 0048 0041 0026 0030 0028 0037 0043 0030 0049 0039 0014 0030 0019 0023 0039 1971 091 091 090 091 090 085 089 094 092 090 093 090 086 094 083 095 096 088 088 088 082 088 084 080 084 092 094 086 096 089 077 095 087 083 086 086 090 094 087 Model performance (likelihood measure) 1972 092 092 093 092 091 093 093 089 091 089 090 088 091 083 092 083 085 090 090 092 092 092 092 091 092 088 081 091 084 088 091 083 089 092 088 091 085 085 090 1973 096 096 096 095 095 095 094 094 093 093 093 094 092 094 092 094 093 092 093 091 092 091 090 092 090 090 092 090 091 089 090 091 089 089 090 089 090 090 088 1974 088 087 087 086 086 088 085 083 083 084 082 082 083 081 084 080 079 082 081 082 085 081 081 083 081 079 078 080 077 079 084 076 079 080 079 078 078 076 078 1975 084 085 084 084 084 082 083 084 082 082 083 084 080 085 080 084 084 080 081 079 079 079 078 079 079 079 083 078 082 077 077 080 076 076 077 077 078 078 076 1976 084 084 083 083 083 082 083 083 082 081 083 082 081 083 081 083 082 081 081 081 079 081 080 079 080 080 082 079 081 079 077 081 078 078 078 080 079 080 078

as conditioned on the calibration data and then using that uncertainty as part of the decision-making process arising from any rainfall-runoff modelling project. The dotty plots in Figure 5 represent a projection of a sample of points on the goodness of t response surface onto individual parameter dimensions. Each dot represents one run of the model from an MCS out of 4000 simulations with different randomly chosen parameter values. It will be seen that for each parameter there are good simulations across a wide range of parameters. The good models are those that plot near the top. This indicates that for each parameter set there are good simulations across a wide range of parameters. There are generally also poor simulations across the whole range of each parameter sampled. Whether a model gives good or poor results is not a function, therefore, of individual parameters, but of the whole set
Copyright 2001 John Wiley & Sons, Ltd. Hydrol. Process. 15, 2281 2295 (2001)

PREDICTING RUNOFF GENERATION

2291

Likelihood measure

Likelihood measure

1 0.8 0.6 0.4 0.2 0 0

* Simulations

1 0.8 0.6 0.4 0.2 0 0.5

* Simulations

0.5 c-Parameter value

0.7

0.9

n-Parameter value

Likelihood measure

0.8 0.6 0.4 0.2 0 45 65 85 m-Parameter value

Likelihood measure

* Simulations

1 0.8 0.6 0.4 0.2 0 0.5

* Simulations

0.6

0.7

0.8

0.9

a-Parameter value

Likelihood measure

Likelihood measure

1 0.8 0.6 0.4 0.2 0 4

* Simulations

1 0.8 0.6 0.4 0.2 0 0.01

* Simulations

0.02

0.03

0.04

0.05

Qo-Parameter value

b-Parameter value

Figure 5. Parameter distributions of behavioural simulations after being conditioned with the 1970 71 period data of G. Ruaha catchment at IKA 8A station

of parameter values and the interactions between parameters. As a projection of the response surface, the dotty plots cannot show the full structure of the complex parameter interactions that shape that surface. In one sense, however, that does not matter too much, since it primarily allows us to identify where the good parameter sets are as a set. An impression of the parameter sensitivity can be gained from Figure 6, which depicts the cumulative likelihood weighted distributions for each parameter after evaluating each behavioural parameter set on data from the 197075 simulation period. Lacking any prior information about the covariation of the individual parameters, each was sampled independently from uniform distributions across the feasible range identied
Copyright 2001 John Wiley & Sons, Ltd. Hydrol. Process. 15, 2281 2295 (2001)

2292

S. MWAKALILA ET AL.

Figure 6. Rescaled likelihood weighted distributions for parameters after being conditioned with the 1970 71 period data at IKA 8A station

from initial evaluation of the TFM and recession ow analysis. A straight diagonal line on each plot would represent the prior distributions. Parameter c and parameter m show the strongest departures from the prior distributions. The other four parameters (Qo , n, a and b) are much less well conditioned by the observations. The possible reasons for parameter sensitivity can be related to the physical interpretation of parameters controlling the slow ow and fast ow pathways of the hydrograph: parameter n controls the non-linearity of fast runoff generation, parameter a control the residence time of the fast ow pathway; parameter b controls the differences in total volumes of input and output; parameter m controls the effective catchment soil prole; parameter c controls the proportion of rainfall that percolates as recharge to the groundwater; parameter Qo
Copyright 2001 John Wiley & Sons, Ltd. Hydrol. Process. 15, 2281 2295 (2001)

PREDICTING RUNOFF GENERATION

2293

represents the discharge when the recession of streamow commences. Therefore, owing to the shallow wet soil in the catchment, evapotranspiration would not be expected to be strongly limited (indirectly controlled by the n parameter), neither would the residence time for surface ow (controlled by the a parameter). Thus any changes in the c and m parameters would result in variations in recharge, which has a major inuence on streamow. Since dotty plots show that for each parameter there are good simulations across a wide range of parameters, all of these good parameter sets will, however, give different predictions. In this study, therefore, the resulting uncertainty in the predictions was estimated by weighting the predictions of all the acceptable models by their associated degree of belief (as quantied by the likelihood measure values). Figures 7 and 8, therefore, present prediction limits for stream discharge from the selected catchment after conditioning in the 197071 simulation period and updating with the likelihoods on individual years for the entire 197075 period data. Figure 9 depicts the prediction limits after testing the behavioural parameter sets with new rainfall and discharge data for the 19761978 observation period. For most of the simulations the observations are bracketed by the prediction limits, but there are periods when the observations fall outside the prediction limits, especially during the wet period. This indicates some deciencies in the input data or model structure used. As noted from Table I, it seems that a second-order transfer function model (2, 2, 0 and 2, 1, 1) does slightly better in terms of R2 . However the YIC has the most negative value for a rst-order
10.00 Discharge (mm/day) 8.00 6.00 4.00 2.00 0.00 13-Dec-69 01-Jul-70 17-Jan-71 05-Aug-71 21-Feb-72 08-Sep-72 Date
Figure 7. Prediction limits after being conditioned and updated with the 1970 72 data

Observed discharge 95% Prediction limit 5% Prediction limit

12.00 Discharge (mm/day) 10.00 8.00 6.00 4.00 2.00

Observed discharge 95% Prediction limit 5% Prediction limit

0.00 27-Nov-72 15-Jun-73 01-Jan-74 20-Jul-74 05-Feb75 Date

24-Aug75

Figure 8. Prediction limits after conditioning and updating with the 1973 75 data

Copyright 2001 John Wiley & Sons, Ltd.

Hydrol. Process. 15, 2281 2295 (2001)

2294

S. MWAKALILA ET AL.

12.00 Discharge (mm/day) 10.00 8.00 6.00 4.00 2.00

Observed discharge 95% Prediction limit 5% Prediction limit

0.00 02-Dec- 19-Jun-76 05-Jan-77 24-Jul-77 09-Feb-78 28-Aug75 78 Date


Figure 9. Prediction limits after testing with the new 1976 78 discharge data

transfer function (1, 1, 0) with high R2 (good t). Therefore, the second-order transfer functions were not tried owing to the fact that they have larger standard errors of the estimated parameter values, an indication that a model is over-parameterized. It should be noted here that, the aim in this study was to nd a model structure that gives a good t but is parsimonious in having a small number of coefcients. The authors suggest that much of the errors may be due to an inadequate representation of the spatial patterns of rainfall within the catchment. Although the models used in this study are not fully distributed process-based models, these results are perhaps representative of the type of accuracy that might be achieved in predicting the discharge while representing the state of the catchment as a single set of process-related parameters. Further, it proves to be surprisingly difcult in rainfall-runoff modelling to bracket all the discharge observations for at least 90% condence interval.

CONCLUSIONS With a DBM modelling approach, the simplest TFM structure that is consistent with the observations is made up of storage elements for fast ow (surface ow) and slow ow (baseow) components. The advantage of the DBM approach is that the data are allowed to suggest the form of the transfer function used rather than specifying a xed structure beforehand. On the basis of the GLUE approach used in this study and the ndings, it can be concluded that for any rainfall-runoff model there is no single best parameter set that represents the catchment for a range of rainfall-runoff responses, but rather a range of different sets of model parameter values may represent the rainfall-runoff process equally well. Therefore, there may be many representations of a catchment that may be equally valid in terms of their ability to produce acceptable simulations of the available data. The parameter sensitivity analysis with the GLUE approach is essentially a non-parametric method, in that it makes no prior assumptions about the variation or covariation of different parameter values, but only evaluates sets of parameter values in terms of their performance. This study has found that recession ow parameters that control the slow ow component are relatively more sensitive than those of the fast ow component. Within the GLUE framework, multiple acceptable parameter sets give a range of predictions. This may actually be an advantage, since it allows the possibility of assessing the uncertainty in predictions as conditioned on the calibration data and then using that uncertainty as part of the decision-making process
Copyright 2001 John Wiley & Sons, Ltd. Hydrol. Process. 15, 2281 2295 (2001)

PREDICTING RUNOFF GENERATION

2295

arising from the modelling project. However, the present study has found that during the wet period the rainfall-runoff modelling does not bracket all the discharge observations for at least 90% condence interval. This is due to errors in the input data used being dominated by peaks during the wet season and being relatively small in the dry period recession. Most of these errors may be due to an inadequate representation of the spatial patterns of rainfall within the catchment. Although the models used in this study are not fully distributed process-based models, these results are perhaps representative of the type of accuracy that might be achieved in predicting the discharge while representing the state of the catchment as a single set of process-related parameters.

ACKNOWLEDGEMENTS

This work was performed during the doctoral study programme of the rst author at the Institute for Land and Water Management (ILWM), Katholieke Universiteit Leuven (K.U.Leuven), Belgium. The study was funded by the Belgian Administration for Development Co-operation (BADC/ABOS) and the University of Dar es Salaam (UDSM), Tanzania. The rst author is pleased to acknowledge this nancial support. He is also grateful for the accessibility to the streamow data and rainfall data provided by the Water Section of the Department of Civil Engineering at UDSM and the Directorate of Meteorology respectively.

REFERENCE Beven KJ, 1993. Prophecy, reality and uncertainty in distributed hydrological modelling. Advances In Water Resources 16: 41 51. Beven KJ. 2001. Rainfall-Runoff Modelling: The Primer . Wiley: Chichester. Beven KJ, Binley AM. 1992. The future of distributed models: model calibration and uncertainty prediction. Hydrological Processes 6: 279298. Beven KJ, Lamb R, Quinn P, Romanowicz R, Freer J. 1995. TOPMODEL. In Computer Models of Watershed Hydrology, Singh VP (ed.). Water Resources Publications: 627 668. Cameron D, Beven KJ, Tawn J, Blazkova S, Naden P. 1999. Flood frequency estimation by continuous simulation for a gauged upland catchment (with uncertainty). Journal of Hydrology 219: 169 187. DANIDA. 1995. Joint study of integrated water and land management in the Great Ruaha Basin. A Final report. DANIDA/World Bank, 1995. Franks SW, Gineste Ph, Beven KJ Mert Ph. 1998. On constraining the predictions of a distributed model: the incorporation of fuzzy estimates of saturated areas into the calibration process. Water Resources Research 34: 787797. Freer J, McDonnel J, Beven KJ, Brammer D, Burns D, Hooper RP, Kendal C. 1997. Topographic controls on subsurface stormow at the hillslope scale for two hydrologically district small catchments. Hydrological Processes 11(9): 1347 1352. Jakeman AJ, Littlewood IG, Whithead PG. 1990. Computation of the instantaneous unit hydrograph and identiable component ows with application to two small upland catchments. Journal of Hydrology 117: 275300. Jakeman AJ, Chen TH, Post DA, Hornberger GM, Littlewood IG, Whithead PG. 1993. Assessing uncertainties in hydrological response to climate at large scale. In Macroscale Modelling of the Hydrosphere, Vol. 214. Wilkinson WB (ed.). IAHS Publication: 3747. Jakeman AJ, Hornberger GW. 1993. How much complexity is warranted in a rainfall-runoff model? Water Resources Research 29: 2637 2649. Melching CS. 1995. Reliability estimation. In Computer Models for Watershed Hydrology, Singh VP (ed.). Water Resources Publications: 69117. Nash JE, Sutcliffe J. 1970. River ow forecasting through conceptual models, part I A discussion of principles. Journal of Hydrology 10: 282290. Post D, Jakeman AJ. 1996. Relationships between catchment attributes and hydrological response characteristics in small Australian mountain ash catchments. Hydrological Processes 10: 877892. Romanowicz R, Beven KJ, Tawn J. 1994. Evaluation of predictive uncertainty in non-linear hydrological models using a Bayesian approach. In Statistics for the Environment II. Water Related Issues, Barnett V, Turkman KF (eds). Wiley: 297317. Schreider SYU, Jakeman AJ. 1996. Modelling rainfall-runoff from large catchment to basin scale: the Goulburn valley, Victoria. Hydrological Processes 10: 863 876. Sefton CEM, Howarth SM. 1998. Relationships between dynamic response characteristics and physical descriptors of catchments in England and Wales. Journal of Hydrology 211: 116. Young PC. 1984. Recursive Estimation and Time Series Analysis. Springer-Verlag: Berlin. Young PC. 1998. Data-based mechanistic modelling of environmental, ecological, economic and engineering systems. Environmental Modelling Software 13: 105122. Young PC, Beven KJ. 1994. Data-based mechanistics modelling and the rainfall-ow non-linearity. Environmetrics 5: 335 363.

Copyright 2001 John Wiley & Sons, Ltd.

Hydrol. Process. 15, 2281 2295 (2001)

S-ar putea să vă placă și