coupling the xinanjiang model to a kinematic flow model based on digital drainage networks for flood...

12
HYDROLOGICAL PROCESSES Hydrol. Process. 23, 1337–1348 (2009) Published online 4 February 2009 in Wiley InterScience (www.interscience.wiley.com) DOI: 10.1002/hyp.7255 Coupling the Xinanjiang model to a kinematic flow model based on digital drainage networks for flood forecasting Jintao Liu, 1,2 Xi Chen, 1 Jiabao Zhang 2 * and Markus Flury 3 1 State Key Laboratory of Hydrology-Water Resources and Hydraulic Engineering, Hohai University, Nanjing 210098, People’s Republic of China 2 State Experimental Station of Agro-Ecosystem in Fengqiu, State Key Laboratory of Soil and Sustainable Agriculture, Institute of Soil Science, Chinese Academy of Sciences, Nanjing 210008, People’s Republic of China 3 Department of Crop and Soil Science, Center for Multiphase Environmental Research, Washington State University, Pullman, WA 99164-6420, USA Abstract: The Xinanjiang model, which is a conceptual rainfall-runoff model and has been successfully and widely applied in humid and semi-humid regions in China, is coupled by the physically based kinematic wave method based on a digital drainage network. The kinematic wave Xinanjiang model (KWXAJ) uses topography and land use data to simulate runoff and overland flow routing. For the modelling, the catchment is subdivided into numerous hillslopes and consists of a raster grid of flow vectors that define the water flow directions. The Xinanjiang model simulates the runoff yield in each grid cell, and the kinematic wave approach is then applied to a ranked raster network. The grid-based rainfall-runoff model was applied to simulate basin-scale water discharge from an 805-km 2 catchment of the Huaihe River, China. Rainfall and discharge records were available for the years 1984, 1985, 1987, 1998 and 1999. Eight flood events were used to calibrate the model’s parameters and three other flood events were used to validate the grid-based rainfall-runoff model. A Manning’s roughness via a linear flood depth relationship was suggested in this paper for improving flood forecasting. The calibration and validation results show that this model works well. A sensitivity analysis was further performed to evaluate the variation of topography (hillslopes) and land use parameters on catchment discharge. Copyright 2009 John Wiley & Sons, Ltd. KEY WORDS kinematic wave method; runoff infiltration; drainage network; digital elevation model; land use; Xinanjiang model Received 5 March 2008; Accepted 8 December 2008 INTRODUCTION The Xinanjiang model (Zhao, 1992) is a conceptual rainfall-runoff model and has been successfully and widely used in humid, semi-humid and even in dry area of China and elsewhere in the world for flood forecasting since its initial development in the 1970s. The main merit of the Xinanjiang model is that it can account for the spatial distribution of soil moisture storage, which has made it outperform other models in a comparison by Gan et al. (1997). The concept of spatial distribution of soil moisture storage has been implemented in other models, e.g. the VIC model (Liang et al., 1996) and the ARNO model (Todini, 1996). Like most conceptual hydrological models with lumped or semi-distributed structure, the spatial variation of hydrological variables is generally difficult to be considered (Chen et al., 2007), which happens to be the advantage of the distributed hydrological models. In recent decades, owing to their capability of explicit spatial representation of hydrological components and variables, the distributed hydrological models have been increasingly applied to account for spatial variability of * Correspondence to: Jiabao Zhang, State Experimental Station of Agro- Ecosystem in Fengqiu, State Key Laboratory of Soil and Sustainable Agriculture, Institute of Soil Science, Chinese Academy of Sciences, Nanjing 210008, People’s Republic of China. E-mail: [email protected] hydrological processes and to support impact assessment studies (e.g. land use change and climate change stud- ies) (Christensen et al., 2007; Das et al., 2008). At the same time, many distributed hydrological models have been developed for storm-runoff simulations, such as the TOPKAPI model (Todini et al., 2001; Liu et al., 2005), LISFLOOD model (De Roo et al., 2000), or the physi- cally based storm-runoff model (Du et al., 2007). Such distributed storm-runoff models do in principle take into account the main mechanisms of flow generation such as saturated-excess overland flow and lateral subsurface flow (Du et al., 2007). Since the data availability was obviously not up to the mark to leverage distributed models, they did not perform better than lumped mod- els (Reed et al., 2004) in the distributed model inter- comparison project (DMIP). In fact, the model performance can vary depending on many factors such as model structure (distributed or lumped), physiographic characteristics of the basin, data available (resolution/accuracy/quantity), so that no single model is perfect and best for all problems (Du et al., 2007; Das et al., 2008). Though many distributed storm-runoff models, e.g. TOPKAPI model (Liu et al., 2005), Du’s model (Du et al., 2007) and LL-II model (Li et al., 2004), have been operationally running for real time flood forecasting purposes, the Xinanjiang model is still one of the most widely used models in humid basins in China due to the lack of spatial data especially soil Copyright 2009 John Wiley & Sons, Ltd.

Upload: jintao-liu

Post on 11-Jun-2016

213 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Coupling the Xinanjiang model to a kinematic flow model based on digital drainage networks for flood forecasting

HYDROLOGICAL PROCESSESHydrol. Process. 23, 1337–1348 (2009)Published online 4 February 2009 in Wiley InterScience(www.interscience.wiley.com) DOI: 10.1002/hyp.7255

Coupling the Xinanjiang model to a kinematic flow modelbased on digital drainage networks for flood forecasting

Jintao Liu,1,2 Xi Chen,1 Jiabao Zhang2* and Markus Flury3

1 State Key Laboratory of Hydrology-Water Resources and Hydraulic Engineering, Hohai University, Nanjing 210098, People’s Republic of China2 State Experimental Station of Agro-Ecosystem in Fengqiu, State Key Laboratory of Soil and Sustainable Agriculture, Institute of Soil Science,

Chinese Academy of Sciences, Nanjing 210008, People’s Republic of China3 Department of Crop and Soil Science, Center for Multiphase Environmental Research, Washington State University, Pullman, WA 99164-6420, USA

Abstract:

The Xinanjiang model, which is a conceptual rainfall-runoff model and has been successfully and widely applied in humid andsemi-humid regions in China, is coupled by the physically based kinematic wave method based on a digital drainage network.The kinematic wave Xinanjiang model (KWXAJ) uses topography and land use data to simulate runoff and overland flowrouting. For the modelling, the catchment is subdivided into numerous hillslopes and consists of a raster grid of flow vectorsthat define the water flow directions. The Xinanjiang model simulates the runoff yield in each grid cell, and the kinematic waveapproach is then applied to a ranked raster network. The grid-based rainfall-runoff model was applied to simulate basin-scalewater discharge from an 805-km2 catchment of the Huaihe River, China. Rainfall and discharge records were available for theyears 1984, 1985, 1987, 1998 and 1999. Eight flood events were used to calibrate the model’s parameters and three other floodevents were used to validate the grid-based rainfall-runoff model. A Manning’s roughness via a linear flood depth relationshipwas suggested in this paper for improving flood forecasting. The calibration and validation results show that this model workswell. A sensitivity analysis was further performed to evaluate the variation of topography (hillslopes) and land use parameterson catchment discharge. Copyright 2009 John Wiley & Sons, Ltd.

KEY WORDS kinematic wave method; runoff infiltration; drainage network; digital elevation model; land use; Xinanjiang model

Received 5 March 2008; Accepted 8 December 2008

INTRODUCTION

The Xinanjiang model (Zhao, 1992) is a conceptualrainfall-runoff model and has been successfully andwidely used in humid, semi-humid and even in dry areaof China and elsewhere in the world for flood forecastingsince its initial development in the 1970s. The main meritof the Xinanjiang model is that it can account for thespatial distribution of soil moisture storage, which hasmade it outperform other models in a comparison byGan et al. (1997). The concept of spatial distributionof soil moisture storage has been implemented in othermodels, e.g. the VIC model (Liang et al., 1996) andthe ARNO model (Todini, 1996). Like most conceptualhydrological models with lumped or semi-distributedstructure, the spatial variation of hydrological variablesis generally difficult to be considered (Chen et al., 2007),which happens to be the advantage of the distributedhydrological models.

In recent decades, owing to their capability of explicitspatial representation of hydrological components andvariables, the distributed hydrological models have beenincreasingly applied to account for spatial variability of

* Correspondence to: Jiabao Zhang, State Experimental Station of Agro-Ecosystem in Fengqiu, State Key Laboratory of Soil and SustainableAgriculture, Institute of Soil Science, Chinese Academy of Sciences,Nanjing 210008, People’s Republic of China.E-mail: [email protected]

hydrological processes and to support impact assessmentstudies (e.g. land use change and climate change stud-ies) (Christensen et al., 2007; Das et al., 2008). At thesame time, many distributed hydrological models havebeen developed for storm-runoff simulations, such as theTOPKAPI model (Todini et al., 2001; Liu et al., 2005),LISFLOOD model (De Roo et al., 2000), or the physi-cally based storm-runoff model (Du et al., 2007). Suchdistributed storm-runoff models do in principle take intoaccount the main mechanisms of flow generation suchas saturated-excess overland flow and lateral subsurfaceflow (Du et al., 2007). Since the data availability wasobviously not up to the mark to leverage distributedmodels, they did not perform better than lumped mod-els (Reed et al., 2004) in the distributed model inter-comparison project (DMIP).

In fact, the model performance can vary dependingon many factors such as model structure (distributedor lumped), physiographic characteristics of the basin,data available (resolution/accuracy/quantity), so that nosingle model is perfect and best for all problems (Duet al., 2007; Das et al., 2008). Though many distributedstorm-runoff models, e.g. TOPKAPI model (Liu et al.,2005), Du’s model (Du et al., 2007) and LL-II model (Liet al., 2004), have been operationally running for realtime flood forecasting purposes, the Xinanjiang model isstill one of the most widely used models in humid basinsin China due to the lack of spatial data especially soil

Copyright 2009 John Wiley & Sons, Ltd.

Page 2: Coupling the Xinanjiang model to a kinematic flow model based on digital drainage networks for flood forecasting

1338 J. LIU ET AL.

data. Therefore, it is necessary and useful to develop anew type of flood forecasting model based on the existingXinanjiang model to improve the model capability inusing more detailed information, such as topography andland cover, in real time flood forecasting. To coupleweather radar rainfall data with the Xinanjiang model,the rainfall-runoff model was set to the same resolution(1 km ð 1 km) as the weather radar rainfall (Li et al.,2004). A similar model structure and model resolution(1 km ð 1 km) were used by Lu et al. (2008) to couple amesoscale atmosphere model with the Xinanjiang model.Both of the studies (Li et al., 2004; Lu et al., 2008)aimed to couple rainfall data with higher resolutionthat traditional rainfall measured by raingage. Thus theXinanjiang model has been justified for incorporating gridscale resolution as 1 km ð 1 km.

Our main objective here is to develop a grid-basedXinanjiang model that can make use of readily availablerainfall, land use and topographic data for basin-scaleflood forecasting. In this study, a grid-based kinematicwave method is used to couple spatial information intothe Xinanjiang model. The coupling is not a virtue initself, but is a technical way to produce an operationalmethodology, while reflecting the necessary physicalprocesses. In the model, the basic approximation is thatthe tension water capacity is spatially heterogeneity anddistributed within each grid (such as 1 km ð 1 km) likein sub-basin scale. And we also assumed that spatialdistribution of tension water capacity is equally the samein every grid, i.e. the curve can be regarded as anaccumulative function or statistical description of thespatial heterogeneity for all pixels. Runoff generated ona partial area in each grid is averaged on the wholegrid and routed to the downstream grid. The model wastested with a series of rainfall-discharge events from theHuangnizhuang sub-basin of the Huaihe River, China.

ISSUES IN THE FLOW ROUTING MODELSAND A STRATEGY TO MODEL BUILDING

As to the flow routing model used in the concep-tual hydrological models, e.g. Xinanjiang model (Zhao,1992), excess rainfall is usually routed by lumpedapproaches such as the unit hydrograph, flow isochronesor linear reservoir modelling in computation of overlandflow and channel flow. These approaches are concep-tually simple and easy to use in flood forecasting, butit is difficult to represent land cover and topography asspatially distributed entities on a basin scale. For manyyears hydrologists have attempted to relate the hydrologicresponse of watersheds to its topographic structures. Thegeomorphologic instantaneous unit hydrograph (GIUH)method (Rodriguez-Iturbe et al., 1979) is perhaps themost promising development in this direction. Maidmentand Olivera (1996) proposed a spatially distributed unithydrograph model in which the watershed is composedof individual cells and the flow paths from cell to cellis determined from the DEM. Ajward and Muzik (2000)

and Zhang et al. (2003) also developed such a type ofunit hydrograph model based on a GIS. Lohmann et al.(1998) derived a linear and time-invariant scheme andrunoff was routed from each cell to the river system withwithout any feedback to the VIC-2L model and withouttransporting water between neighbouring grids. In thesemodels based on linear systems theory, flow from onecell is not affected by flow from neighbour cells.

The kinematic wave overland flow approach is aphysically based method and is well suited for consid-ering both land cover and topography condition. Theapproach has been frequently used to model the rainfall-runoff process for simple as well as complex water-shed geometries (Kibler and Woolhiser, 1970; Singh andWoolhiser, 1976; Michaud and Sorooshian, 1994). It hasbeen used in many distributed hydrological models, e.g.IHDM (Calver et al., 1995), KINEROS (Smith et al.,1995), WEP (Jia et al., 2001) and WEHY (Kavvas et al.,2004). As the grid-square arrangement is adopted by mostexisted, distributed hydrology model, e.g. SHE (Abbottet al., 1986), TOPKAPI model (Todini et al., 2001; Liuet al., 2005), LISFLOOD model (De Roo et al., 2000), aphysically based storm-runoff model (Du et al., 2007),it was used in our model. However, the area of thebasic unit termed a sub-basin in the Xinanjiang model isquite larger area than that of the grid-square unit of dis-tributed model. So when applying the physically basedkinematic method with the grid-square arrangement intothe Xinanjiang model to allow for the easy importationof data from remotely-sensed source, the basin need to besubdivided to modify the Xinanjiang model’s structures.The distributed flow routing model is not only capableof accounting for spatial variability of hydrological pro-cesses, but it also enables computation of internal fluxes.

THE COMBINED RAINFALL-RUNOFF MODELAND ITS ADAPTION FOR DIGITAL DRAINAGE

NETWORKS

Runoff generation in the kinematic wave Xinanjiangmodel (KWXAJ)

In the Xinanjiang model (Zhao, 1983, 1992), runoffproduction at a point occurs on the repletion of storagecapacity, i.e. infiltration occurs until the soil moisturecapacity is reached. To represent the spatial distributionof the soil moisture storage capacity over the basin, aparabolic curve, i.e. soil moisture storage capacity curve(SMSCC) is used in the Xinanjiang model.

In the modified Xinanjiang model, spatial distributionof rainfall input, vegetation, land use and topographiesof the watershed is achieved by a grid network. Here,the grid-based rainfall-runoff model named the kinematicwave Xinanjiang model (KWXAJ) consists of two parts:a runoff yield model and a kinematic overland flowrouting model for a grid network. The Xinanjiang modelwas used to calculate runoff yield in each grid element.The parameters used in the runoff yield component arelisted in Table I.

Copyright 2009 John Wiley & Sons, Ltd. Hydrol. Process. 23, 1337–1348 (2009)DOI: 10.1002/hyp

Page 3: Coupling the Xinanjiang model to a kinematic flow model based on digital drainage networks for flood forecasting

COUPLING THE XINANJIANG MODEL TO A KINEMATIC FLOW MODEL 1339

Table I. Parameters used the for runoff yield algorithm in the Xinanjiang model

Parameters Description Value

K Ratio of potential evapotranspiration to pan evaporation 1Ð25B Exponent of the spatial distribution curve of tension water storage capacity 0Ð3C Evapotranspiration coefficient of the deep layer 0Ð2IMP (%) Percentage of the impervious area of catchment 0Ð01WUM (mm) Tension water capacity of the upper layer 20WLM (mm) Tension water capacity of the lower layer 60WDM (mm) Tension water capacity of the deep layer 40SM (mm) Free water storage capacity 15Ex Exponent of the spatial distribution curve of free water storage capacity 1Ð2Ki Outflow coefficient of free water storage to the interflow 0Ð25Kg Outflow coefficient of free water storage to the groundwater 0Ð45

h

h0

P+h0+∑ri-E

Overlandinflow qsup

Overlandoutflow qs

Deep layer outflow rg

Subsurfaceinflow ∑ri

A S0

xL

IMPW0

Rainfall P

Evapotranspiration E

WM

SMSCC

Subsurfaceoutflow ri

Infiltration

Figure 1. Schematic of runoff generation and the soil moisture storage capacity curve on hillslope for the modified Xinanjiang model (after Zhao, 1992)

Rainfall-runoff processes within each grid of the com-bined model can be described as follows. When rainfallP exceeds evapotranspiration E, rainfall is infiltrated intosoil reservoir and overland flow qs will not occur untilthe soil reservoir is saturated. Excess water is trans-ferred to downstream grid and finally routed to the outlet.To a given grid, overland flow qsup (the correspondingrunoff depth is h0), subsurface flow (

∑ri ) from upland

grids as well as rainfall are regarded as inflow. So therunoff infiltration, i.e. runon is considered in the mod-ified Xinanjiang model, which makes it differ from thetraditional version of Xinanjiang model. Runoff depth iscalculated by the Xinanjiang model and it is separatedinto overland flow rs, subsurface flow ri, and deep layerflow rg. In Figure 1, the runoff generation on a hillslopewith a length L and slope S0 is shown. The percentageof the imperious area is IMP, W0 and WM are initialareal mean tension, water storage and areal mean tensionwater capacity, respectively.

Runoff routing for digital drainage networks

In the kinematic wave model for raster systems, dis-charge leaving the downstream boundary of a rasterenters the upstream boundary of a higher-level raster andserves to establish the boundary conditions of depth anddischarge required by the kinematic wave method. Over-land flow from any direction to the raster is regarded aslateral inflow. According to the D-8 algorithm (Fairfield

c

a

b

d

i,j

y

x

x’

y’

1 2

3

456

8

7

Figure 2. Raster sketch for overland flow routing

and Leymarie, 1991), any raster in the grid system haseight possible flow directions (Figure 2). When the out-flow direction is parallel to x or y, the length Li,j of thecell is equal to a and the width Wi,j equal to b, when theoutflow direction is parallel to the diagonal such as x0, y0,a presumed raster is defined with its length c and widthd. The length Li,j and the width Wi,j are expressed as:

Li,j D√

a2 C b2 �1�

Wi,j D a Ð b/√

a2 C b2 �2�

Copyright 2009 John Wiley & Sons, Ltd. Hydrol. Process. 23, 1337–1348 (2009)DOI: 10.1002/hyp

Page 4: Coupling the Xinanjiang model to a kinematic flow model based on digital drainage networks for flood forecasting

1340 J. LIU ET AL.

Outlet grid of the 6×6 grid network

Figure 3. (a) Raster flow direction definition for a 6 ð 6 grid network; (b) ranked raster for each raster to the outlet

The drainage network extraction algorithm describedby Martz and Garbrecht (1992) was used to implementthe structure of the digital drainage network, including theflow vectors and the raster slopes. The flow concentrationsystem of the basin is routed according to a simplealgorithm. Each cell has a routing order and the cellsare ranked by the following principles: (1) cells withoutinflow are defined rank 1, and (2) the ranking number ofeach cell is equal to the maximum ranking number amongneighbour upland cells plus 1. There is no flow betweencells of equal rank. An example of the flow network andthe ranking numbers is shown in Figure 3.

The one-dimensional kinematic wave equation forshallow water flow over is given by:

{ ∂h∂t C ∂uh

∂x D qlat�x, t�

S0 D Sf

�3�

Here h is overland flow depth (FD), u is depth averageflow velocity, S0 is the hillslope, qlat�x, t� is the netinflow, and Sf is the friction slope. Under the kinematicwave approximation, the discharge is a function ofoverland FD (Kibler and Woolhiser, 1970):

q D uh D ˛hm �4�

Here q�x, t� is the flow discharge per unit width, ˛ andm are coefficients. Using Manning’s equation, m D 5/3and ˛ D S1/2

0 /n in which n is the Manning coefficient(MC ). Substitution of Equation 4 into Equation 3 yieldsthe kinematic wave equation with one dependent variable:

∂h

∂tC ˛mhm�1 ∂h

∂xD qlat�x, t� �5�

An explicit difference method based on the MacCor-mack scheme (MacCormack, 1971) is employed here tosolve the governing differential equations. The resultingdifference equations are:

Predictor step:

hŁi D hn

i � t

x�qn

iC1/2 � qni�1/2� C qlati

nC1/2t �6�

qniš1/2 D ˛

(hn

iš1 C hni

2

)m

�7�

Corrector step:

hnC1i D 1

2

[hŁ

i C hni � t

x �qnC1/2iC1/2 � qnC1/2

i�1/2 �

CqlatinC1/2t

]�8�

qnC1/2iš1/2 D 1

2˛[

(hn

iš1 C hni

2

)m

C(

hŁiš1 C hŁ

i

2

)m

�9�

The initial condition for overland flow is a dry surface

h�x, 0� D 0 �10�

The boundary condition for cells with their rankingnumber equal to 1 is taken as zero inflow. Thus,

h�0, t� D 0 �11�

For other cells, the boundary condition is given below:

h�0, t� D

N∑kD1

qsup,k

˛

1m

�12�

where N is the number of upstream inflow cells, qsup,k isthe discharge from upstream inflow cell.

Numerical experiments were carried out to comparethe performance of the MacCormack schemes with theanalytical solution of kinematic wave equations (Singh,1996) for a hypothetical hillslope. The simulations wereapplied to a test hillslope of 900 m long and 1 m wide,

Copyright 2009 John Wiley & Sons, Ltd. Hydrol. Process. 23, 1337–1348 (2009)DOI: 10.1002/hyp

Page 5: Coupling the Xinanjiang model to a kinematic flow model based on digital drainage networks for flood forecasting

COUPLING THE XINANJIANG MODEL TO A KINEMATIC FLOW MODEL 1341

012345678

0 30 60 90 120 150 180

Time (min)

Dis

char

ge (

×10−4

m2 /s

)

Analytical solution Numerical solution

Figure 4. Comparison of numerical experiment results between analytical and numerical solutions

which was subjected to a rainfall of 0Ð05 mm/min for120 min. The MC was n D 0Ð02 and the slope of thehillslope was 0Ð75%. The simulation was run for 300 minusing a time step of 90 s and a space discretization of30 m.

The model performance in terms of discharge wasevaluated by the mean absolute error (MAE ):

MAE D 1

n

n∑i

jqin � qiaj �13�

where qa, qn represent flow of analytical and numericalsolution, respectively. The numerical simulation matchedthe analytical solution very well, as shown in Figure 4and confirmed by the low value of MAE D 4Ð58 ð10�6 m2/s.

Overland flows are routed from the most upstream cellsto the outlet cell. Before entering into downstream cells,the subsurface flow of each cell is routed by the ‘lag androute’ technique used in the original Xinanjiang model.Deep layer outflow (qg) is gathered to the groundwaterpool directly (see in Figure 1) and discharge from thegroundwater pool is also routed by the ‘lag and route’technique.

CASE STUDY AREA AND DATA

We selected the Huangnizhuang basin (Figure 5) formodel calibration and validation. This basin has an areaof 805 km2, and is part of the Shiguanhe basin (31°120 �32°180N, 115°170 � 115°550E) of Huaihe River which hasan area of 158 160 km2. The Huangnizhuang basin ischaracterized by mountains and steep slopes. Rainfallis often intense. The annual rainfall is about 1250 mmand the runoff coefficient is about 0Ð6. Temporal-spatialdistribution of the rainfall is uneven and storms arecentralized during June to September.

Event hourly discharge and rainfall and daily pan evap-oration were monitored at the Huangnizhuang station(31°280N, 115°370E). Event hourly rainfall was also mon-itored at seven other locations in the basin (Figure 5).Hydrological data from 1984, 1985, 1987, 1998 and 1999were used in this study. The spatial distribution of the pre-cipitation was obtained using Thiessen polygon interpola-tion (Figure 6a). A DEM for the catchment was available

from the U.S. Geological Survey’s EROS Data Cen-ter (GTOPO30). Elevations in GTOPO30 are regularlyspaced at 30 arcsecond (about 1 km) (Figure 6b). Thecatchment was discretized in cells of size 30 ð 30 inch,and 42 different ranks were assigned (Figure 6c). Thefrequency distribution of the individual ranks is shownin Figure 7. The higher-order cells correspond to peren-nial rivers or ephemeral waterways and cells with higherrank than 27 are defined as channel cells by comparedwith the real channels. We assumed that the MC (n) forchannel cells is equal 0Ð04. Surface slope was derivedfrom DEM data. The slope frequency distribution showsthat slopes of about 0Ð09% were most frequent and theaverage slope was about 0Ð122% (Figure 8).

Vegetation distribution of the basin at a spatialresolution of 30 arcsecond was obtained from theUMD 1 km Global Land Cover database for 1999(http://www.geog.umd.edu/landcover/1km-map.html).Six types of vegetation were identified and the distri-bution is shown in Figure 6d. Nearly 78% of the basinis covered by woodland (Table II). Grids with differentvegetation cover (land use) were assumed to have dif-ferent MCs. The MCs were also different for each stormevent as explained below. The soil type is relatively sin-gular in our mountainous Huangnizhuang basin and thesoil category is Lithosols according to the FAO Soil Mapof the World (FAO, 1998).

RESULTS AND DISCUSSION

Calibration of the model

The Xinanjiang model parameters for Huangnizhuangsub-basin were calibrated by eight flood events. Sincemost parameters such as K, C, IMP, WUM, WLM, WDM,are related to the average climate and surface conditionsof the studied region and as it is one of the tributariesof Shiguanhe basin, they were preset by referring to theexisting results by Li et al. (2004). Among the parametersof Table.I, only B and SM were optimally calibrated asthe time and space scale change (Zhao and Liu, 1995).The best selection for ex is between 1Ð0 and 1Ð5 and itmay be taken as a constant (Zhao and Liu, 1995). Ki andKg are the outflow coefficients of the free water storageto interflow and groundwater. As the recession during ofthe upper interflow storage ordinarily lies between 2 and

Copyright 2009 John Wiley & Sons, Ltd. Hydrol. Process. 23, 1337–1348 (2009)DOI: 10.1002/hyp

Page 6: Coupling the Xinanjiang model to a kinematic flow model based on digital drainage networks for flood forecasting

1342 J. LIU ET AL.

Shiguanhe Basin

Huangnizhuang Sub-Basin

Figure 5. Location of the studied area Huangnizhuang sub-basin in Shiguanhe basin

3 days, it is suggested that the sum Ki C Kg may be takenas 0Ð7–0Ð8 and the ratio of the three runoff componentswill be changed by altering the ratio of Kg /Ki (Zhao andLiu, 1995). Calibration of Manning’s roughness is givenbelow in the following text.

The trial and error method was used and we alteredthe values to make the simulated discharge matchingthe observed one. In the kinematic wave model, theManning’s roughness coefficients for all of the overlandflow cells within the basin were defined according tothe land use information and were calibrated for eachflood event. The set of MCs used for the flood events aresummarized in Table III. The average value of MCs ofdifferent flood events varied between 0Ð082 to 0Ð194 andfor all the flood events selected for calibration the averageMCs is 0Ð116. Those values are comparable to the MCsvalues obtained from experiments by Woolhiser (1975)and Engman (1986). We generally observed that floodevent with large peak flow corresponded to large MC.The average MC was linearly related to the average FD(Figure 9), and the same result was also reported by Wangand Hjelmfelt (1998). This is consistent with the relationfor MCs against FD reported by Barling and Moore(1994). In fact, overland flow at very shallow depthencounters maximum resistance because the vegetationis upright in the flow. As pointed by Wang and Hjelmfelt

(1998), the MC here not only depends on the landcover, but also serves as a correction factor for thehydraulic radius approximation. The linear relationshipgiven between the average MC and the average FDshown in Figure 9 indicates that MC is highly relatedwith the flood grade and has great influence on thesimulated discharge.

Model performance was evaluated by the efficiencycoefficient (EC ):

EC D 1 �∑

�Qobs � Qsim�2

∑�Qobs � Qobs�

2�14�

where Qobs and Qsim are observed and simulated dis-charge (m3/s), respectively, and Qobs is the average valueof observed discharge.

The calibration results are summarized in Table IV.Relative errors of the simulated runoff volume range from�17Ð2 to 0Ð43%. The peak discharge errors range from4Ð2 to 14Ð0%. By and large, the simulated and observedpeak discharges agreed well (Table IV, Figure 10), whichis corroborated by an EC value close to 1.

Validation of the model

Three flood events in 1984 and 1987 were selected tovalidate the performance of the grid-based rainfall-runoff

Copyright 2009 John Wiley & Sons, Ltd. Hydrol. Process. 23, 1337–1348 (2009)DOI: 10.1002/hyp

Page 7: Coupling the Xinanjiang model to a kinematic flow model based on digital drainage networks for flood forecasting

COUPLING THE XINANJIANG MODEL TO A KINEMATIC FLOW MODEL 1343

(a) (b)

(c) (d)

Figure 6. Digital map for Huangnizhaung basin. (a) Thiessen polygons, (b) digital elevation model, (c) sequence number, (d) spatial distribution ofvegetation type

0

100

200

300

400

500

8 >8sequence number

num

ber

of g

rids

1 2 3 4 5 6 7

Figure 7. Number of grids of different sequence for kinematic waverouting

model. Two groups of MCs were used for the validation,(1) the average MCs shown in Table III and (2) the MCsderived from the linear relationship in Figure 9. Thesimulated results for the three flood events are listed inTable V and in Figure 11. Generally, the observed andsimulated hydrographs agree well. The relative errors ofrunoff volume and peak discharge agree are low, and theEC is large (Table V). This indicates that the simulationsare consistent with the observations.

0.00

0.02

0.04

0.06

0.08

0.10

0.01 0.09 0.18 0.26 0.34 0.43

Slope (%)

Freq

uenc

y

Figure 8. Slope histograms for Huangnizhuang basin

Flood event 9, June 1984, was a very small floodcompared with flood event 10 and 11. The average FDwas about 0Ð54 mm. According to the linear relationship(Figure 9), the MC is also relatively small. By using theMC derived from the linear relationship, the relative errorfor the peak discharge was 3Ð7%, the relative error for thetotal runoff was �5Ð3% and EC was 0Ð92 which indicates

Copyright 2009 John Wiley & Sons, Ltd. Hydrol. Process. 23, 1337–1348 (2009)DOI: 10.1002/hyp

Page 8: Coupling the Xinanjiang model to a kinematic flow model based on digital drainage networks for flood forecasting

1344 J. LIU ET AL.

Table II. Vegetation types in the Huangnizhuang basin

Evergreen needle leaf forest Deciduous broad leaf forest Mixed forest Woodland Wooded grassland Cropland

Area �km2� 19Ð6 36Ð9 24Ð8 624Ð4 88Ð0 11Ð3Percentage (%) 2Ð4 4Ð6 3Ð1 77Ð6 10Ð9 1Ð4

Table III. Manning coefficients for the different flood events

Beginning of flooda Vegetation type Average value

Evergreenneedleleaf forest

Deciduousbroadleaf forest

Mixedforest

Woodland Woodedgrassland

Cropland

09/09/1984 0Ð121 0Ð121 0Ð165 0Ð096 0Ð086 0Ð077 0Ð09905/07/1985 0Ð135 0Ð135 0Ð185 0Ð107 0Ð096 0Ð086 0Ð11005/07/1987 0Ð177 0Ð177 0Ð242 0Ð140 0Ð126 0Ð113 0Ð14328/08/1987 0Ð127 0Ð127 0Ð173 0Ð100 0Ð090 0Ð081 0Ð10302/07/1998 0Ð110 0Ð110 0Ð150 0Ð090 0Ð080 0Ð070 0Ð09222/06/1999 0Ð132 0Ð132 0Ð180 0Ð105 0Ð094 0Ð084 0Ð10826/06/1999 0Ð225 0Ð225 0Ð310 0Ð190 0Ð175 0Ð150 0Ð19429/06/1999 0Ð102 0Ð102 0Ð138 0Ð080 0Ð072 0Ð065 0Ð082Average value 0Ð141 0Ð141 0Ð193 0Ð114 0Ð102 0Ð091 0Ð116

a Given as dd/mm/yyyy.

y = 20.337x − 1.0589R2 = 0.823

0

0.5

1

1.5

2

2.5

3

3.5

0 0.05 0.1 0.15 0.2 0.25

Average MC

Ave

rage

FD

(m

m)

Figure 9. Relationship between the average flow depth and the averageMC

a good agreement between simulated and observed data.With the average MC (Table III), the simulated resultswere less accurate. There was a 2-h lag between the twosimulated peaks as the MCs are differed by about 47%.

The average FD of Flood event 10, August 1984, wasabout 0Ð91 mm and the MCs derived from the linearrelationship were somewhat smaller than the averagevalue. Using the MC derived from the linear relationshipand the average MC, the relative error for the peakdischarge is �6Ð6, �19Ð0% and the relative error forthe total runoff is 3Ð3, �4Ð4% and EC is 0Ð93, 0Ð88,respectively. The lag between the two simulated peakdischarge was 1 h.

Flood event 11, August 1987, was one of the biggestflood events in this basin. The average FD was about2Ð54 mm. The MCs derived by the linear relationshipare larger than the average MCs. Again, peak dischargeoccurred 1 h earlier when using the averaged MC ascompared to the one derived from the linear relationship.

By and large, the simulations show that the MCderived from the linear relationship described the exper-imental data below better than the average MC. Overall,good agreement between model simulations and observeddata was observed. All of the simulated results usingthe average MCs and the MCs derived from the linear

Table IV. Calibration results by the grid-based rainfall-runoff model for different flood events

No. Beginning offlooda

Observedpeak

discharge (m3/s)

Simulatedpeak

discharge (m3/s)

Relativeerror(%)

Observedrunoff

volume (ð104 m3)

Simulatedrunoff

volume (ð104 m3)

Relativeerror(%)

EfficiencyCoefficient

1 09/09/1984 418Ð0 458Ð3 9Ð6 2559Ð6 2528Ð4 �1Ð2 0Ð912 05/07/1985 1100Ð0 1000Ð7 �9Ð0 3506Ð2 3860Ð8 10Ð1 0Ð903 05/07/1987 1127Ð6 1133Ð5 0Ð52 4361Ð4 4568Ð9 4Ð7 0Ð934 28/08/1987 924Ð6 1054Ð2 14Ð0 4398Ð6 4302Ð1 �2Ð2 0Ð875 02/07/1998 213Ð0 227Ð0 6Ð6 1034Ð8 1016Ð1 �1Ð8 0Ð886 22/06/1999 765Ð0 710Ð2 �7Ð2 3172Ð2 3765Ð6 18Ð7 0Ð897 26/06/1999 2150Ð0 2242Ð7 4Ð3 9359Ð8 9400Ð3 0Ð43 0Ð968 29/06/1999 319Ð0 332Ð3 4Ð2 1402Ð3 1161Ð2 �17Ð2 0Ð71

a Given as dd/mm/yyyy.

Copyright 2009 John Wiley & Sons, Ltd. Hydrol. Process. 23, 1337–1348 (2009)DOI: 10.1002/hyp

Page 9: Coupling the Xinanjiang model to a kinematic flow model based on digital drainage networks for flood forecasting

COUPLING THE XINANJIANG MODEL TO A KINEMATIC FLOW MODEL 1345

R2 = 0.988

0

500

1000

1500

2000

2500

0 500 1000 1500 2000 2500

Observed peak discharge (m3/s)

Sim

ulat

ed p

eak

disc

harg

e (m

3 /s)

Figure 10. Comparison of observed and simulated peak discharges

relationship reach the accuracy standard according to TheAccuracy Standard of Hydrological Forecasting in China(Hydrological Bureau of China, 2000).

Sensitivity analysis

It has been shown in the flood validation simulationsthat the discharge is greatly influenced by the MC. WhenMCs were increased, as in the simulations of Flood 9and 10, the simulated peak discharge was reduced andpeak time deferred. And when MCs were decreased as incase of the Flood 11, the simulated peak discharge wasenlarged and peak time advanced.

We carried out an uncertainty analysis to evaluate theeffects of topography and land use on hydrological waterdischarge. In the kinematic wave approach, topographyand land are represented by the parameter ˛, i.e., thesurface slope (So) and the MC (n). For the uncertaintyanalysis, we individually varied surface slope and rough-ness by š5 and š10% of their original values.

One of the biggest flood events, Flood 7, was used inthe uncertainty analysis. The relevant results are shown inTable VI and Figure 12. As the slopes became steeper,the total volume of discharge as well as the peak dis-charge increased. Correspondingly, as MCs increased,both total discharge and peak discharge decreased. More-over, when roughness was increased by 10% the peakdischarge occurred 1 h later and when it was decreased

by 10% the peak discharge occurred 1 h earlier than theobserved one as shown in Table VI and Figure 12.

We calculate the relative variation of total runoff andpeak discharge as a function of slope and Manningroughness. The total runoff varied from �2Ð24 to 1Ð92%and peak discharge varied from �3Ð72 to 2Ð09% whensurface slope changed by š10% (Figure 13). WhenManning roughness was changed by š10%, the totalrunoff varied from 4Ð11 to �4Ð23% and peak dischargefrom 5Ð88 to �6Ð55% (Figure 13). These results showthat Manning roughness is the more sensitive parameterthan the slope, and that the peak discharge is moresensitive than the total runoff.

From this uncertainty analysis, it was clear that surfaceslope and Manning roughness are two most importantcharacteristics that determine the overland flow routing.One deduction may be that if a large area of the catchmentwas reclaimed as crop land or as urban development, forboth of which the surface roughness is low, the runoffvolume will increase as will the peak discharge. Evenworse, the peak time of the discharge will be advanced.This can ultimately lead to enhanced soil erosion by moreintensive overland flow and flooding.

SUMMARY AND CONCLUSIONS

A modified Xinanjiang model was developed for basin-scale flood modelling. The KWXAJ works on a rastersystem for representing the spatial heterogeneity of thebasin and can fully utilize distributed information ofmodel’s inputs, such as rainfall from weather radar,topography and land uses. The Xinanjiang model wasused for grid runoff yield in the rainfall-runoff model.The kinematic wave method was used as overland flowrouting model for the grid-based rainfall-runoff model inthis paper. And before determining runoff depth for a cellusing Xinanjiang model, inflow from upstream cells, i.e.the runoff infiltration, was considered.

The main objective of this paper has been to demon-strate the practical implementation of a basin-scalerainfall-runoff model coupled by a distributed flow rout-ing model for flood forecasting and its ability to simu-late a variety of basin responses. The simulations match

Table V. Validation results by the grid-based rainfall-runoff model for different flood events

No. Beginning offlooda

Observedpeak

discharge (m3/s)

Simulatedpeak

discharge (m3/s)

Relativeerror(%)

Observedrunoff

volume (ð104 m3)

Simulatedrunoff

volume (ð104 m3)

Relativeerror(%)

Efficiencycoefficient

Manning coefficients are derived by the linear relationship between the average FD and the average MC9 08/08/1984 718Ð0 670Ð4 �6Ð6 1821Ð2 1882Ð2 3Ð3 0Ð9310 12/06/1984 164Ð3 170Ð3 3Ð7 985Ð1 933Ð3 �5Ð3 0Ð9211 22/08/1987 1928Ð2 1617Ð5 �16Ð1 5660Ð1 6571Ð5 16Ð1 0Ð92

Manning coefficients are derived by the average value in Table III9 08/08/1984 718Ð0 581Ð5 �19Ð0 1821Ð2 1740Ð7 �4Ð4 0Ð8810 12/06/1984 164Ð3 132Ð1 �19Ð7 985Ð1 837Ð7 �15Ð0 0Ð7911 22/08/1987 1928Ð2 1912Ð3 �0Ð82 5660Ð1 6699Ð9 18Ð4 0Ð86

a Given as dd/mm/yyyy.

Copyright 2009 John Wiley & Sons, Ltd. Hydrol. Process. 23, 1337–1348 (2009)DOI: 10.1002/hyp

Page 10: Coupling the Xinanjiang model to a kinematic flow model based on digital drainage networks for flood forecasting

1346 J. LIU ET AL.

0

20

40

60

80

100

120

140

160

180

1984−6−129:00

1984−6−1213:00

1984−6−1217:00

1984−6−1221:00

1984−6−131:00

1984−6−135:00

1984−6−139:00

Time

Dis

char

ge (

m3 /s

)

0

5

10

15

20

25

30

Rai

nfal

l (m

m/h

)

rainfall

observation

using MC by the linearrelationshipusing average MC

(a)

0

100

200

300

400

500

600

700

800

1984/8/88:00

1984/8/812:00

1984/8/816:00

1984/8/820:00

1984/8/90:00

1984/8/94:00

1984/8/98:00

Time

Dis

char

ge (

m3 /s

)

0

10

20

30

40

50

Rai

nfal

l (m

m/h

)rainfall

observation

using MC by the linearrelationshipusing average MC

(b)

0

500

1000

1500

2000

2500

1987−8−228:00

1987−8−2212:00

1987−8−2216:00

1987−8−2220:00

1987−8−230:00

1987−8−234:00

1987−8−238:00

1987−8−2312:00

Time

Dis

char

ge (

m3 /s

)

0

10

20

30

40

50

60

Rai

nfal

l (m

m/h

)

rainfall

observationusing MC by the linearrelationshipusing average MC

(c)

Figure 11. Observed and simulated hydrograph for three flood events in the Huangnizhuang station. (a) Flood event 9, (b) Flood event 10, (c) Floodevent 11

Table VI. Simulation results under different topography and land use

Different topographyand land usea

Total runoffvolume

(ð104 m3)

Relativeerror (%)

Peakdischarge

(m3/s)

Relativeerror (%)

Peak timeerror (h)b

Efficiencycoefficient

Observation 9359Ð8 2150SLOP90 9189Ð3 1Ð8 2159Ð3 �0Ð4 0 0Ð95SLOP95 9297Ð9 0Ð66 2206Ð1 �2Ð6 0 0Ð95SLOP105 9494Ð1 �1Ð4 2270Ð1 �5Ð6 0 0Ð96SLOP110 9582Ð6 �2Ð39 2290Ð1 �6Ð5 0 0Ð96ROUGH90 9794Ð4 �4Ð7 2377Ð3 �10Ð6 1 0Ð96ROUGH95 9596Ð3 �2Ð5 2292Ð7 �6Ð6 0 0Ð96ROUGH105 9204Ð4 1Ð7 2166Ð5 �0Ð77 0 0Ð96ROUGH110 9011Ð3 3Ð7 2100Ð9 1Ð8 �1 0Ð94

a SLOP90, SLOP95, SLOP105, SLOP110 represent slopes equal to original slope being multiplied by 0Ð90, 0Ð95 1Ð05, 1Ð10, respectively, while thevalue of the roughness are not changed. ROUGH90, ROUGH95, ROUGH105 and ROUGH110 have the same meaning for the Manning coefficientsas explained for the slope.b ‘0’ means the simulated peak time equals to the observed peak time, ‘1’ means the simulated peak time is 1 h ahead of the observed peak time,‘�1’ means the simulated peak time is 1 h behind of the observed peak time.

Copyright 2009 John Wiley & Sons, Ltd. Hydrol. Process. 23, 1337–1348 (2009)DOI: 10.1002/hyp

Page 11: Coupling the Xinanjiang model to a kinematic flow model based on digital drainage networks for flood forecasting

COUPLING THE XINANJIANG MODEL TO A KINEMATIC FLOW MODEL 1347

600

900

1200

1500

1800

2100

2400

1999−6−2710:00

1999−6−2713:00

1999−6−2716:00

1999−6−2719:00

1999−6−2722:00

Time

Dis

char

ge (

m3 /s

)

obs. sim. slope90

slope95 slope105 slope110

rough90 rough95 rough105

rough110

Figure 12. Simulated hydrograph of Huangnizhuang station by changingslope and Manning coefficients

−5

−4

−3

−2

−1

0

1

2

3

0.9 0.95 1 1.05 1.1

Relative variation for slop

Rel

ativ

e va

riat

ion

(%)

runoffpeak discharge

−8

−6

−4

−2

0

2

4

6

8

0.9 0.95 1 1.05 1.1

Relative variation for roughness

Rel

ativ

e va

riat

ion

(%)

runoffpeak discharge

(b)

(a)

Figure 13. Total runoff and peak discharge variation (%) versus variationof surface slope (a) and roughness (b)

the observations closely. The calibration and validationresults show that MC is highly related with the floodgrade and large flood events have larger values of MCswhereas small events have smaller values. A relationshipbetween MC and FD is derived in this paper. A sen-sitivity analysis was used to reveal the performance ofthis model. The effects of topography and land use wereassessed in a series of numerical simulations. The analy-sis showed that both of surface slope and land use con-siderably affect total predicted runoff volume and peakdischarge. The sensitivity analysis indicates that changesin land use, such as changing forest land to crop or urbanland or town, have the potential to increase peak flow andadvance the peak time.

Although the grid-based model has been verified inflood event modelling, we realize that soil and other landsurface conditions are strong heterogonous, even in a1 ð 1 km2 pixel. In this case, pixel parameters from theseveral groups or a singular set can not be fully accumu-lated or integrated into the spatial distribution curve ofthe Xinanjiang model (XM). This is what we believe thelimitation of the present grid-based hydrological models.Even though these models’ structure in description ofspatial distribution is physically reasonable, their appli-cations in watershed hydrological simulation still existuncertainties due to difficulty in parameter determinationfor each grid. Adopting the tension water capacity curveof XM in our model development can be regarded asa compromise between physical reasonability and appli-cation feasibility for the grid-based hydrological mod-elling. So when using the Xinanjiang model with smallgrid scales (1 km ð 1 km), many questions are still to beanswered. For instance (1) What is the spatial distributionof the tension water capacity within each grid? (2) Howcan the tension water capacity be estimated from veg-etation and soil texture data? (3) How should we routerunoff generated in partial area of the grids?

ACKNOWLEDGEMENTS

This work was supported by the Innovation programof Chinese Academy of Sciences (kzcx2-yw-406), theNational Basic Research Program of China (973 project:2005CB121103), the Key Research Grant from ChineseMinistry of Education (Project No. 308012) and theNational Natural Science Foundation of China (GrantNo.: 40801013).

REFERENCES

Abbott MB, Bathurst JC, Cunge JA, O’Connell PE. 1986. An introductionto the European Hydrological System-Systeme HydrologiqueEuropeen, “SHE”, 2, Structure of a physically-based, distributedmodeling system. Journal of Hydrology 87: 61–77.

Ajward MH, Muzik I. 2000. A Spatially Varied Unit Hydrograph Model.Journal of Environmental Hydrology 8: 1–8.

Barling RD, Moore ID. 1994. Role of buffer strips in managementof waterway pollution: a review. Environmental Management 18:543–558.

Calver A, Wood WL. 1995. The institute of hydrology distributed model.In Computer Models of Watershed Hydrology , Singh VP (ed.). WaterResource publications: Colorado; 595–626, Chapter 17.

Chen X, Chen YD, Xu CY. 2007. A distributed monthly hydrologicalmodel for integrating spatial variations of basin topography andrainfall. Hydrological Processes 21: 242–252.

Christensen N, Lettenmaier DP. 2007. A multimodel ensemble approachto assessment of climate change impacts on the hydrology and waterresources of the Colorado River basin. Hydrology and Earth SystemSciences 11: 1417–1434.

Das T, Bardossy A, Zehe E, He E. 2008. Comparison of conceptualmodel performance using different representations of spatial variability.Journal of Hydrology 356: 106–118.

De Roo APJ, Wesseling CG, Van Deursen WPA. 2000. Physically-based river basin modelling within a GIS: The LISFLOOD model.Hydrological Processes 14: 1981–1992.

Du JK, Xie SP, Xu YP, Xu CY, Singh VP. 2007. Development and testingof a simple physically-based distributed rainfall-runoff model for stormrunoff simulation in humid forested basins. Journal of Hydrology 336:334–346.

Copyright 2009 John Wiley & Sons, Ltd. Hydrol. Process. 23, 1337–1348 (2009)DOI: 10.1002/hyp

Page 12: Coupling the Xinanjiang model to a kinematic flow model based on digital drainage networks for flood forecasting

1348 J. LIU ET AL.

Engman ET. 1986. Roughness coefficients for routing surface runoff.Journal of Irrigation and Drainage Engineering-ASCE 112(1): 39–53.

Fairfield J, Leymarie P. 1991. Drainage networks from grid digitalelevation models. Water Resources Research 27: 709–717.

FAO. 1998. Land and water digital media series N-1, December, ISBN92-5-104050-8.

Gan TF, Dlamini EM, Biftu GF. 1997. Effects of model complexity andstructure, data quality, and objective functions on hydrologic modeling.Journal of Hydrology 192: 81–103.

Jia YW, Ni GS, Kawahara Y, Suetsugi T. 2001. Development ofWEP model and its application to an urban watershed. HydrologicalProcesses 15: 2175–2194.

Kavvas ML, Chen ZQ, Dogrul C, Yoon JY, Ohara N, Liang L,Aksoy H, Anderson ML, Yoshitani J, Fukami K, Matsuura T. 2004.Watershed environmental hydrology (WEHY) model based on upscaledconservation equations: hydrologic module. Journal of HydraulicEngineering-ASCE 9: 450–464.

Kibler DF, Woolhiser DA. 1970. Mathematical properties of thekinematic cascade. Journal of Hydrology 15: 131–147.

Li ZJ, Ge WZ, Liu JT, Zhao K. 2004a. Coupling between weatherradar rainfall data and a distributed hydrological model for real-timeflood forecasting. Hydrological Sciences Journal-Journal Des SciencesHydrologiques 49: 945–958.

Li L, Zhu RR, Zhong MJ. 2004b. Application of LL-II distributed rainfall-runoff model based on GIS. Water Resources Power 22(1): 8–11, (inChinese).

Liang X, Lettenmaier DP, Wood EF. 1996. One-dimensional statisticaldynamic representation of subgrid spatial variability of precipitationin the two-layer variable infiltration capacity model. Journal ofGeophysical Research 101(D16): 21: 403–21, 422.

Liu Z, Martina MVL, Todini E. 2005. Flood forecasting using a fullydistributed model: application of the TOPKAPI model to the UpperXixian Catchment. Hydrology and Earth System Sciences 9: 347–364.

Lohmann D, Raschke E, Nijssen B, Lettenmaier DP. 1998. Regionalscale hydrology: I. Formulation of the VIC-2L model coupled to arouting model. Hydrological Sciences Journal-Journal des SciencesHydrologiques 43: 131–141.

Lu GH, Wu ZY, Wen L, Lin CA, Zhang JY, Yang Y. 2008. Real-timeflood forecast and flood alert map over the Huaihe River Basin inChina using a coupled hydro-meteorological modeling system. Sciencein China Series E-Technological Sciences 51: 1049–1063.

MacCormack RW. 1971. Numerical solution of the interaction of a shockwave with a laminar boundary layer. Lecture Notes in Physics 8:151–163, Springer-Verlag.

Maidment DR, Olivera F. 1996. Unit Hydrograph Derived From ASpatially Distributed Velocity Field. Hydrological Processes 10:831–844.

Martz LW, Garbrecht J. 1992. Numerical definition of drainage networkand subcatchment areas from digital elevation models. Computers &Geosciences 18: 747–761.

Michaud JD, Sorooshian S. 1994. Comparison of simple versus complexdistributed runoff models on a midsized, semiarid watershed. WaterResources Research 30: 596–605.

Reed S, Koren V, Smith M, Zhang Z, Moreda F, Seo DJ, DMIPParticipants. 2004. Overall distributed model intercomparison projectresults. Journal of Hydrology 298: 27–60.

Rodriguez-Iturbe I, Valdes JB. 1979. The geomorphologic structure ofhydrologic response. Water Resources Research 15: 1409–1420.

Singh VP. 1996. Kinematic Wave Modeling in Water Resources: SurfaceWater Hydrology . John Wiley and Sons: New York.

Singh VP, Woolhiser DA. 1976. A nonlinear kinematic wave model forwatershed surface runoff. Journal of Hydrology 31: 221–243.

Smith ER, Goodrich DC, Woolhiser DA, Unkrich CL. 1995. KINEROS-A kinematic runoff and erosion model. In Computer Models ofWatershed Hydrology , Chapter 7, Singh VP (Ed.). Water ResourcesPublication: Colorado; 697–732.

Todini E. 1996. The ARNO rainfall-runoff model. Journal of Hydrology175: 339–382.

Todini E, Ciarapica L. 2001. The TOPKAPI model. In Mathematicalmodels of large watershed hydrology , Chapter 12, Singh VP (Ed.).Water Resources Publications: Colorado, USA.

Wang MH, Hjelmfelt AT. 1998. DEM based overland flow routing model.Journal of Hydraulic Engineering-ASCE 3: 1–8.

Woolhiser DA. 1975. Simulation of unsteady overland flow. In UnsteadyFlow in Open Channels , Vol. II Mahmood K, Yevjevich V (eds). WaterResources Publications: Fort Collins, Co; 502.

Zhang Q, Tan LY, Xu YP, Ge XP. 2003. Exploration of distributed unithydrograph mode to rainfall-runoff. Journal of Nanjing University 39:139–143, (in Chinese).

Zhao RJ. 1983. Watershed Hydrological Model—Xinanjiang Model andShanbei Model . Water & Power Press: Beijing, (in Chinese).

Zhao RJ. 1992. The Xinanjiang model applied in China. Journal ofHydrology 135: 371–381.

Zhao RJ, Liu XR. 1995. The Xinanjiang model. In Computer Models ofWatershed Hydrology , Singh VP (ed.). Water Resources Publication:Colorado; 215–232, Chapter 7.

Copyright 2009 John Wiley & Sons, Ltd. Hydrol. Process. 23, 1337–1348 (2009)DOI: 10.1002/hyp