You are on page 1of 8

Journal of Hydrology 489 (2013) 238–245

Contents lists available at SciVerse ScienceDirect

Journal of Hydrology
journal homepage: www.elsevier.com/locate/jhydrol

An efficient method for DEM-based overland flow routing


Pin-Chun Huang, Kwan Tun Lee ⇑
Department of River and Harbor Engineering, National Taiwan Ocean University, Keelung 202, Taiwan, ROC

a r t i c l e i n f o s u m m a r y

Article history: The digital elevation model (DEM) is frequently used to represent watershed topographic features based
Received 21 November 2012 on a raster or a vector data format. It has been widely linked with flow routing equations for watershed
Received in revised form 2 March 2013 runoff simulation. In this study, a recursive formulation was encoded into the conventional kinematic-
Accepted 12 March 2013
and diffusion-wave routing algorithms to permit a larger time increment, despite the Courant–Fried-
Available online 21 March 2013
This manuscript was handled by
rich–Lewy condition having been violated. To meet the requirement of recursive formulation, a novel
Konstantine P. Georgakakos, Editor-in-Chief, routing sequence was developed to determine the cell-to-cell computational procedure for the DEM data-
with the assistance of Attilio Castellarin, base. The routing sequence can be set either according to the grid elevation in descending order for the
Associate Editor kinematic-wave routing or according to the water stage of the grid in descending order for the diffusion-
wave routing. The recursive formulation for 1D runoff routing was first applied to a conceptual overland
Keywords: plane to demonstrate the precision of the formulation using an analytical solution for verification. The
Overland flow proposed novel routing sequence with the recursive formulation was then applied to two mountain
Computational efficiency watersheds for 2D runoff simulations. The results showed that the efficiency of the proposed method
Digital elevation model was significantly superior to that of the conventional algorithm, especially when applied to a steep
Diffusion wave watershed.
Kinematic wave
Ó 2013 Elsevier B.V. All rights reserved.
Courant–Friedrich–Lewy condition

1. Introduction strains the maximum time increment depending on the wave


celerity; the restriction is especially stringent for steep mountain
Due to considerable progress on the computational technique watersheds having rapid flow. Although nonlinear implicit meth-
over the past few decades, complex numerical algorithms have ods have been proposed and are claimed to be more flexible in
been adopted in hydrological modeling in connection with a assigning the time increment (Li et al., 1975; Fread, 1976), the
high-resolution DEM database. The simplest and most straightfor- use of iterative calculations or matrix operations may disadvan-
ward method for flow direction determination in a DEM dataset is tage their computational efficiency. In addition, such complex cal-
the deterministic eight-node (D8) approach (O’Callaghan and culations require extremely large virtual memory space on a
Mark, 1984), in which a single drainage path for each cell is as- computer.
signed to one of its eight neighbors toward the steepest down- To achieve efficient computing, MacCormack (1982) proposed
slope. Although multiple-flow direction methods have been an unconditional stable scheme to allow for an expanding time
suggested to account for the dispersion of a cell to all neighboring increment in performing unsteady-flow routing. A recursive
cells having lower elevations (Freeman, 1991; Quinn et al., 1991), formulation was added to each time step of the explicit scheme
error may be introduced in the inconsistently physical definition to enable variable modification without complying to the CFL sta-
of upstream drainage areas (Tarboton, 1997; Orlandini et al., bility condition. Nevertheless, in performing the MacCormack
2003). Consequently, the D8 approach is most widely adopted for recursive formulation, the routing sequence from upstream to
flow-direction determination by hydrologists. downstream outlet is strictly enforced. It results that this formula-
To depict the concentration process of surface runoff in a wa- tion cannot be properly performed for the D8 runoff routing in a
tershed, various approximations of St. Venant equations are DEM database.
linked with numerical algorithms for runoff routing. According This study aims to improve the computing efficiency for runoff
to the conventional explicit scheme, the computational efficiency routing in connection with a high-resolution DEM database. To al-
of the scheme is restricted by the Courant–Friedrich–Lewy (CFL) low for application of the D8 method, the MacCormack recursive
stability condition (Courant et al., 1967). The CFL condition con- formulation (MacCormack, 1982) was reformulated and combined
with a novel cell-to-cell routing sequence in the DEM database.
Numerical algorithms, based on kinematic- and diffusion-wave
⇑ Corresponding author. Fax: +886 2 24634122. approximations, were developed to link with the recursive formu-
E-mail address: ktlee@ntou.edu.tw (K.T. Lee). lation for watershed runoff routing. The precision of the numerical

0022-1694/$ - see front matter Ó 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jhydrol.2013.03.014
P.-C. Huang, K.T. Lee / Journal of Hydrology 489 (2013) 238–245 239

algorithms in the runoff simulation was verified using the on the variances between wave celerity and numerical marching
analytical solution for a conceptual overland plane, and then the speed (Dx/Dt). MacCormack (1982) estimated this value as follows:
proposed method was applied to two mountain watersheds to  
Dx
demonstrate the computational efficiency of the method. k ¼ a  Max 0; c  ð9Þ
Dt
2. Governing equations for watershed runoff routing in which a is a coefficient (a negative a should be avoided for the
computation), and c is the wave celerity. This function implies that
Since the D8 method is applied for flow-direction determination when the CFL condition is preserved, k is equal to 0 because the
in this study, the expression of the governing equation is essen- numerical marching speed (Dx/Dt) is larger than the wave celerity.
tially in a 1D (or quasi-2D) form. The continuity equation is written Under other conditions, k is larger than 0, and a modified Dyi can be
as obtained from Eq. (8) and used to estimate the water depth at t + Dt,
@y @q as shown in Eq. (7). The coefficient a shown in Eq. (9) mainly de-
þ ¼ ie ð1Þ pends on the wave celerity and geomorphologic features of the
@t @x
watershed.
in which y is the water depth; ie is the rainfall excess rate, and q is
the flow discharge per unit width, which can be calculated as
  3.1. Verification of the MacCormack stable formulation using an
1 2=3 1=2 analytical solution
q ¼ Vy ¼ R Sf y ð2Þ
n
in which V is the average flow velocity; n is the Manning’s rough- Rainfall-runoff processes on an impervious overland plane were
ness coefficient; R is the hydraulic radius, which is considered equal used as an example to verify the capability of the MacCormack
to y for shallow water flow; and Sf is the friction slope. For a kine- unconditional stable formulation when the CFL condition was vio-
matic-wave (KW) approximation, the friction slope of the flow is as- lated. The overland plane was 850 m long with a slope of 0.3. The
sumed to be equal to the bed slope, as shown by the term runoff simulation on the overland plane began with an initially dry
surface, and the rain lasted 60 min with an intensity of 70 mm/h.
Sf ¼ So ð3Þ The 1D flow models based on the KW and DFW approximations
in which So is the bed slope. Hence, the CFL stability condition of the were developed using a finite-difference scheme, in which the
KW approximation is Manning roughness coefficient was set as 0.1 in the simulations.
To perform the conventional backward-finite-difference scheme
3 nDx (HEC, 1990; Bedient et al., 2008), the time increment was set at
Dt 6 pffiffiffiffiffi ð4Þ
5 So y2=3 2 s to comply with the CFL condition. However, to have the fi-
nite-difference scheme include the MacCormack stable formula-
If the influence of the pressure term on flow routing is further
tion, the time increment was enlarged to 30 s to investigate the
included, the friction slope is expressed as
accuracy of the algorithm when the CFL condition was violated.
@y As shown in Fig. 1a , although the time to equilibrium discharge
Sf ¼ So  ð5Þ
@x calculated using the numerical models was slightly later than that
This equation shows that the water-stage gradient determines of the analytical solution, the hydrographs generated from the con-
the motion of runoff, which is known as the diffusion-wave ventional and the MacCormack algorithms based on KW approxi-
(DFW) model. Corresponding to the CFL stability condition of the mation were close to the analytical solution. As shown in Fig. 1b
diffusion wave, the maximum time increment constraint can be for the DFW model, the simulated hydrograph generated from
denoted as follows (Weinmann and Laurenson, 1979; Huang and the MacCormack algorithm, without complying with the CFL con-
Song, 1985): dition, also showed a fairly similar result to that obtained using
the conventional algorithm. In performing the algorithms, the coef-
Dx ficient a in Eq. (9) was assigned as 1.0 for the KW model and as 0.6
Dt 6 pffiffiffiffiffi ð6Þ
V þ gy for the DFW model to obtain the best fit results with the analytical
where V is the average flow velocity shown in Eq. (2), and g is the solution. This analysis showed that, although its computing time
gravity acceleration. The use of an explicit scheme for runoff routing increment was 15 times larger than that of the conventional algo-
requires strict adherence to the stability criteria at each time step to rithm, the MacCormack stable formulation preserved the accuracy
ensure correct numerical results. of results.

3. MacCormack unconditional stable formulation 4. MacCormack stable formulation for DEM-based overland
flow routing
To alleviate the CFL restriction in performing an explicit
scheme, MacCormack (1982) developed a simple and straightfor- To enable quasi-2D runoff routing on a DEM database, modifica-
ward formulation to ensure numerical stability in computations tions are required for the discrete formulations. Fig. 2 shows an
involving large time increments. According to MacCormack example of the flow direction using the D8 method (O’Callaghan
(1982), the estimated water depth at t + Dt can be expressed as and Mark, 1984), in which each grid has a single downstream
yi ðt þ DtÞ ¼ yi ðtÞ þ Dy0i ð7Þ drainage route but may possess multiple inflow sources from
upstream neighboring grids. The discrete continuity equation
in which using the backward-finite-difference scheme for flow routing on
Dyi þ k DDxt Dy0i1 a quasi-2D system can be expressed as
Dy0i ¼ ð8Þ
1 þ k DDxt Dt 1
Dyi ðt þ DtÞ ¼ ½q ðtÞ  qi ðtÞ þ ½ie ðtÞ þ ie ðt þ DtÞDt ð10Þ
Dx i;IN 2
where Dyi denotes the time-varying increment of water depth at
location i, and Dy0i represents the modified Dyi. Eq. (8) is used to en- where qi(t) denotes the flow discharge per unit width at location i;
sure the stability of computational results. The parameter k depends qi,IN(t) is the total inflow discharge (namely, the accumulated
240 P.-C. Huang, K.T. Lee / Journal of Hydrology 489 (2013) 238–245

0.02 0.02

0.016 0.016

0.012 0.012
q (m 2/s)

q (m 2/s)
0.008 0.008

So=0.3, no=0.1 So=0.3, no=0.1


0.004 KW analytical solution 0.004 KW analytical solution
conventional (KW, Δt=2sec) conventional (DFW, Δt=2sec)
MacCormack (KW, Δt=30sec) MacCormack (DFW, Δt=30sec)
0 0
0 20 40 60 0 20 40 60
time (min) time (min)
(a) kinematic-wave model (b) diffusion-wave model
Fig. 1. Verification of MacCormack stable algorithm using conceptual overland plane.

1 2 3 drainage route to the downstream outlet. In Fig. 3b, the arrow rep-
resents the flow direction for each grid, according to the maximum
elevation drop between grids; the flow direction of the surround-
4 5 6 ing grids in the database is forced to point to the periphery. Hence,
once the outlet of a watershed is assigned, the watershed boundary
can be delineated, as shown by the dotted line in Fig. 3b. In accor-
dance with the flow direction depicted in Fig. 3b, the drainage net-
7 8 9 work can be constructed as shown in Fig. 3c.
The conventional flow-routing sequence in a DEM database typ-
ically starts in the upper-left corner and progresses stepwise to the
Fig. 2. Flow direction assigned in a DEM-based system. lower-right corner. As shown in Fig. 4a, the numeral allocated to
each cell represents the routing sequence in conventional grid-
based computing, and grids lying beyond the watershed boundary
are assigned ‘‘00’’ and are not routed. The conventional routing se-
discharge from upstream neighboring grids), and ie(t) is the rainfall
quence shown in Fig. 4a violates the rule of using the recursive for-
excess rate at location i. For example, the inflow discharge of Grid 5
mulations of Eqs. (12) and (13). For example, the discharge of the
shown in Fig. 2 can be calculated as
upstream-inflow grid (No. 16) should be calculated prior to calcu-
q5;IN ðtÞ ¼ q2 ðtÞ þ q3 ðtÞ þ q6 ðtÞ ð11Þ lating the discharge of the downstream grid (No. 15). Conse-
quently, the grids marked with a rectangle, shown in Fig. 4b,
Consequently, the MacCormack recursive formulation shown in
would hinder runoff routing using the MacCormack recursive for-
Eq. (8) is modified to fit the D8 method on a DEM-based dataset as
mulation in the DEM-based system.
follows:
To relax the restriction of using Eqs. (12) and (13) in KW rout-
Dyi þ k DDxt Dy0i;IN ing, the routing sequence was designed to follow the watershed
Dy0i ¼ ð12Þ topography from higher-elevation grids to lower-elevation grids.
1 þ k DDxt
This enabled that the discharges from the upstream-inflow grids
If the Manning equation is used to estimate the flow velocity, can be calculated prior to calculating the downstream grid dis-
Dy0i;IN in Eq. (12) can be further derived as charge. As shown in Fig. 5a, the numerals allocated to the grids
show the proposed new flow-routing sequence in descending or-
Dy0i;IN ðt þ DtÞ ¼ yi;IN ðt þ DtÞ  yi;IN ðtÞ
!3=5 der of the grid elevations shown in Fig. 3a. Hence, numerals shown
n on the upstream grids in Fig. 5b are always smaller than those
¼ pffiffiffiffiffi qi;IN ðt þ DtÞ  yi;IN ðtÞ ð13Þ shown on the downstream grids. For example, in Fig. 5b, Grids 3
Sf
and 11 (upper-left corner) are connected to the downstream Grid
in which the friction slope Sf can be estimated from Eq. (3) for KW 14. Hence, the discharges of Grids 3 and 11 should be calculated
routing and from Eq. (5) for DFW routing. It should be noted that before the discharge of Grid 14 if Eqs. (12) and (13) are to be imple-
since the total inflow discharge qi,IN(t + 1) is included in the recur- mented. Consequently, the numeral shown on the watershed out-
sive formulations in Eqs. (12) and (13), the discharges of up- let equals the total number of grids in the watershed; that is, 43
stream-inflow grids should be obtained prior to calculating the grids. According to the new routing sequence shown in Fig. 5, the
downstream-grid discharge. MacCormack recursive formulation can then be performed suc-
cessfully. Fig. 6 shows a schematic connection between flow path
4.1. Flow routing sequence for DEM-based watersheds and routing sequence proposed in this study. The dotted line rep-
resents the connection of flow direction between grids, and the so-
Fig. 3a shows an example of the DEM database, in which the nu- lid line represents the routing sequence for performing the
meral shown on each grid represents the elevation of the grid. If proposed algorithm.
the D8 method was adopted to determine the flow direction based The flow routing sequence shown in Figs. 5 and 6 is suitable
on the elevation drop between grids, each grid has a single only for the KW method because the friction slope is assumed
P.-C. Huang, K.T. Lee / Journal of Hydrology 489 (2013) 238–245 241

90 92 91 95 93 92 92 95 93
72 91 87 90 89 87 88 90 78
78 89 93 83 83 85 87 92 74
75 90 72 69 69 82 83 84 77
88 67 70 68 72 70 83 85 89
78 66 69 67 71 72 82 95 83
73 64 63 64 65 82 85 94 81
61 57 59 66 71 71 83 88 87
51 62 67 74 78 82 91 90 95

(a) grid elevation drainage outlet


(b) flow direction

(02)
(01) (03) (04) (05)
(08) (11)
(06) (09) (10)
(07)
(14) (16) (17)
(13) (15)
(12)
(19) (20) (21) (22) (23)
(18)
(29)
(27) (30)
(26) (28)
(24) (25)
(34)
(33) (35)
(31) (32) (36)
(41)
(43)
(38) (39) (40) (42)
(37)

(c) drainage network


Fig. 3. Grid elevation, flow direction, and drainage network in a DEM dataset.

(02)
00 00 00 00 00 00 00 00 00 (01) (03) (04) (05)
00 00 01 02 03 04 05 00 00 (08) (11)
(06) (09) (10)
(07)
00 06 07 08 09 10 11 00 00 (14) (17)
(15) (16)
(13)
00 12 13 14 15 16 17 00 00 (12)
(22)
(19) (20) (21) (23)
00 18 19 20 21 22 23 00 00 (18)
(29)
(27) (30)
(26) (28)
00 24 25 26 27 28 29 30 00 (24) (25)
(34)
00 31 32 33 34 35 36 00 00 (33) (35)
(31) (32) (36)
(41)
00 37 38 39 40 41 42 43 00 (38) (39) (40)
(43)
(42)
00 00 00 00 00 00 00 00 00 (37)

(a) conventional flow-routing (b) drainage network


sequence
Fig. 4. Conventional flow-routing sequence is set to start from the upper-left corner to the lower-right corner of the DEM dataset.

equal to the ground slope in the KW approximation, as described in 5. Model verification for mountain watersheds
Eq. (3). Consequently, the flow routing sequence can be deter-
mined before starting a KW runoff routing based on the relative Geomorphologic and hydrological data from the Goodwin Creek
elevation drop between grids. Nevertheless, in performing the Experimental Watershed in Northern Mississippi, USA, were col-
DFW model, the spatial variation of water stage dominates the lected for the proposed algorithm test. Goodwin Creek is a tribu-
flow direction between grids, as shown in Eq. (5); hence, the rout- tary of Long Creek which flows into Yocona River. In 1981 the
ing sequence should be determined at each time step according to Goodwin Creek watershed was placed under the administration
the temporal variation of water stage between grids. To clearly de- of the National Sedimentation Laboratory of the United States
pict the routing processes of the proposed method, a flowchart for Department of Agriculture. Flow and sediment discharges in Good-
both the KW and DFW models was provided as shown in Fig. 7. win Creek are monitored by 14 gauging stations, and precipitation
242 P.-C. Huang, K.T. Lee / Journal of Hydrology 489 (2013) 238–245

(03)
00 00 00 00 00 00 00 00 00 (11) (05) (09) (08)
00 00 11 03 05 09 08 00 00 (14) (10)
(06) (16) (13)
(02)
00 06 02 14 16 13 10 00 00 (32) (20) (15)
(24) (31)
00 04 24 32 31 20 15 00 00 (04)
(29) (33) (23) (28) (18)
00 35 29 33 23 28 18 00 00 (35)
(19)
(25) (01)
(34) (22)
00 36 30 34 25 22 19 01 00 (36) (30)
(38)
00 40 41 39 38 21 12 00 00 (39) (21)
(40) (41) (12)
(26)
00 43 42 37 27 26 17 07 00 (42) (37) (27)
(07)
(17)
00 00 00 00 00 00 00 00 00 (43)

(a) proposedflow-routing sequence (b) drainage network


Fig. 5. Proposed flow-routing sequence is set according to the grid elevation in descending order for the DEM dataset.

outlet
01 02 03 ..... 11 12 13 14 ..... 43 Input DEM dataset

flow-routing sequence
flow direction DFW model KW model

Fig. 6. Flow-routing sequence and flow direction between grids.

Sorting the water stage Sorting the elevation


of each cell of each cell
is measured by 31 rain gauges located in or near the watershed.
Discharge data from STA05 and the rainfall records from No. 55
and No. 57 rain gauges within the subwatershed were collected
for model verification. The watershed extends over 3.95 km2 with
a mean channel slope of 0.0072 and an overland slope of 0.0380.
i=1
Fig. 8 shows the channel network extracted from a 30-m resolution
digital elevation data set using a DEM (Lee, 1998), in which the wa-
tershed topography is represented by 3890 grids.
Rainfall and discharge data obtained at 10-min intervals from a
Derive Δyi
rainstorm event on 23 March 1996, were used in this study for
using continuity equation (Eq.10)
model verification. The rainfall excess hyetograph was determined
by deducting the abstractions from the rainfall using the Green–
Ampt equation (Green and Ampt, 1911). The Manning’s roughness
Derive Δy′i
used in the KW and DFW models was set as 0.22 which was cali-
using MacCormack equation (Eq.12)
brated based on the hydrological records from Goodwin Creek wa-
tershed by using the conventional algorithm. The allowable i = i +1
maximum time increment determined by the CFL stability condi- Estimate water depth yi ( t + Δt )
tion was 11 s for the KW routing and 7 s for the DFW routing. Con- using Eq.7
sequently, in applying the conventional algorithms to both runoff
models, the time increment was set at 5 s. When the proposed
method was used, the time increment was increased from 10 to Derive discharge qi ( t + Δt )
60 s to investigate simulation precision and computational using momentum equation (Eq.2)
efficiency.
Table 1 shows the detailed simulated results for the rainstorm
on 23 March 1996 in the Goodwin Creek watershed, using various Accumulate discharge
numerical algorithms in KW and DFW routings. The flow hydro- into downstream cell
graph for this storm lasted 16 h, and the recorded peak discharge
was 10.21 m3/s. As shown in Table 1, the simulated peak dis-
charges calculated using the conventional algorithms and the pro-
DFW model KW model
posed method were essentially congruent with the recorded data; Next time step
the relative errors for peak discharge simulations were less than 4%
in all cases. Table 1 also shows that the relative root mean square
error (RMSE) increased from 0.0065 for Dt = 10 s to 0.0167 for
Dt = 60 s in KW routing. Similarly, it increased from 0.0077 for
Stop
Dt = 10 s to 0.0184 for Dt = 60 s in DFW routing. These results
showed that, although the relative RMSE increased with increasing
Dt, RMSE remained less than 2%, even when Dt was 12-fold that Fig. 7. Flowchart of the proposed method.
P.-C. Huang, K.T. Lee / Journal of Hydrology 489 (2013) 238–245 243

reducing computing time for both KW and DFW routings. Fig. 9


shows the simulated results using the conventional algorithm
(Dt = 5 s) and the proposed method for KW and DFW routings
(Dt = 60 s). These results demonstrated that the hydrographs gen-
erated using the proposed method fitted fairly well with the re-
corded data, despite the CFL condition having been violated in
performing the runoff routing.

5.1. Model performance for steep mountain watersheds

As shown in Eqs. (4) and (6), to meet the CFL stability restric-
tion, a small time increment should be selected for steep water-
sheds when performing conventional numerical algorithms. To
Fig. 8. Map of the Goodwin Creek watershed. further investigate the computational efficiency of the proposed
algorithm, it was applied to a steep mountain watershed located
at Su–Ao Township in Northern Taiwan. As shown in Fig. 10, the
used in the conventional algorithm. Thus, the overall performance watershed covers 0.36 km2, and the average slope of the watershed
of the proposed method was extremely satisfactory. The MacCor- is 0.63; these measurements are based on analysis of a 5-m resolu-
mack recursive formulation provided an efficient method for tion DEM dataset (total 14,352 grids). Discharge records are

Table 1
Comparison of observed and simulated hydrographs in the Goodwin Creek watershed using different algorithms for the storm event occurred on 23 March 1996.

Runoff model Numerical algorithm Time increment Dt (s) Relative Qp error Relative RMSE CUP time (min) Relative CPU time
Kinematic wave Conventional 5 0.0303 – 1.05 1.000
This study 10 0.0315 0.0065 0.53 0.505
20 0.0344 0.0081 0.26 0.248
30 0.0392 0.0089 0.18 0.171
40 0.0388 0.0117 0.13 0.124
50 0.0371 0.0132 0.11 0.105
60 0.0354 0.0167 0.09 0.086
Diffusion wave Conventional 5 0.0240 – 1.31 1.000
This study 10 0.0257 0.0077 0.65 0.496
20 0.0284 0.0091 0.33 0.252
30 0.0265 0.0102 0.22 0.168
40 0.0253 0.0143 0.16 0.122
50 0.0232 0.0167 0.13 0.099
60 0.0215 0.0184 0.11 0.084

Note: 1. Relative Qp error = |(Qp)simulated  (Qp)recorded|/(Qp)recorded.


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
PT ffi
1 2
2 Relative RMSE ¼ T t¼1 ½ððQ t Þthis study  ðQ t Þconv entional Þ=ðQ t Þconv entional  .
3. Relative CPU time = (CPU time)this_study/(CPU time)conventional.

40 40
i e (mm/hr)

i e (mm/hr)

30 30
20 20
10 10
0 0

10 1996/03/23 10 1996/03/23
Recorded Recorded
conventional (KW, Δt = 5 sec) conventional (DFW, Δt = 5 sec)
8 this study (KW, Δt = 60 sec) 8 this study (DFW, Δt = 60 sec)
Q (m3/s)

Q (m3/s)

6 6

4 4

2 2

0 0
0 4 8 12 16 0 4 8 12 16
time (hr) time (hr)
(a) (b)
Fig. 9. Comparison of observed and simulated hydrographs generated from different algorithms using (a) kinematic-wave model and (b) diffusion-wave model for the 23
March 1996 storm.
244 P.-C. Huang, K.T. Lee / Journal of Hydrology 489 (2013) 238–245

Rainfall records obtained from the Dong-Ao rain-gauging station


for Typhoon Megi, which occurred on 20 October 2010, were used
as an example for runoff routing to test the efficiency of the pro-
posed method. The Dong-Ao station is situated 2.31 km from the
watershed outlet, and the rainfall records were obtained at 10-
min intervals. The Green–Ampt equation (Green and Ampt, 1911)
was used to deduct the rainfall abstraction to obtain the rainfall ex-
cess. The parameters used in the Green–Ampt equation were
determined according to the soil characteristics of the watershed.
In accordance with the CFL stability restriction for the conven-
tional KW algorithm applied to the Su–Ao watershed, the time
increment calculated by Eq. (4) was equal to 2 s, and it was 0.5 s
(using Eq. (6)) for the conventional DFW algorithm. These values
were smaller than those used in the computation for the Goodwin
Creek watershed (11 s and 7 s, respectively). Consequently, the
runoff routing was expected to be time-consuming if conventional
Fig. 10. Map of the Su–Ao watershed. algorithms were applied for the routing on a steep watershed using
a high resolution DEM database. As mentioned for quasi-2D rout-
ing, the flow routing sequence should be determined before per-
unavailable for this watershed; thus, to calibrate the model param- forming a runoff simulation. For the Su–Ao watershed with its
eters, the Manning coefficient was assigned as 0.6 according to the 14,352 grids, computing time was only 0.02 s by using Shellsort
land cover conditions of the watershed and the roughness values method (Pratt, 1972) to arrange the flow-routing sequence for
suggested by the Hydrologic Engineering Center (HEC, 2000). the whole watershed.

160
i e (mm/hr)

i e (mm/hr)

160
120 120
80 80
40 40
0 0
16 16
2010/10/20 2010/10/20
14
conventional (KW, Δt = 2 sec) conventional (DFW, Δt = 0.5 sec)
this study (KW, Δt = 60 sec) this study (DFW, Δt = 60 sec)
12 12

10
Q (m3/s)

Q (m3/s)

8 8

4 4

0 0
0 12 24 36 48 60 72 84 96 0 12 24 36 48 60 72 84 96
time (hr) time (hr)
(a) (b)
Fig. 11. Comparison of simulated hydrographs for different algorithms using (a) kinematic-wave model and (b) diffusion-wave model for the 20 October 2010 storm.

Table 2
Comparison of observed and simulated hydrographs in the Su–Ao watershed using different algorithms for the storm event occurred on 20 October 2010.

Runoff model Numerical algorithm Time increment Dt (s) Relative RMSE CPU time (min) Relative CPU time
Kinematic wave Conventional 2 – 121.3 1.000
This study 10 0.0024 25.7 0.212
20 0.0067 13.5 0.111
30 0.0081 8.2 0.068
40 0.0098 6.1 0.050
50 0.0134 4.9 0.040
60 0.0182 4.1 0.034
Diffusion wave Conventional 0.5 – 491.5 1.000
This study 10 0.0054 25.9 0.053
20 0.0086 13.6 0.028
30 0.0097 8.3 0.017
40 0.0127 6.4 0.013
50 0.0187 5.1 0.010
60 0.0209 4.2 0.008
P.-C. Huang, K.T. Lee / Journal of Hydrology 489 (2013) 238–245 245

Fig. 11 shows the simulated hydrographs for the example storm Acknowledgments
generated from the KW and DFW models using various algorithms.
For KW routing, as shown in Table 2, although the relative RMSE This research was supported by the National Science Council,
increased with increasing Dt, relative RMSE values remained less Taiwan, ROC. Financial support provided by the National Science
than 2% when the proposed algorithm was used, even when Dt Council is acknowledged.
was increased to more than 30-fold the values used in the conven-
tional algorithm. Moreover, in performing DFW routing, the com- References
puting time for the proposed method (CPU time = 4.2 min for
Dt = 60 s) was a mere 1/117 that required for the conventional Bedient, P.B., Huber, W.C., Vieux, B.E. 2008. Chapter 4: Flood Routing in Hydrology
and Floodplain Analysis. Prentice-Hall, Inc.
algorithm (CPU time = 491.5 min for Dt = 0.5 s), and the relative Courant, R., Friedrichs, K., Lewy, H., 1967. On the partial difference equations of
RMSE value of the hydrograph was only 2.09%. A comparison of mathematical physics. IBM J. 11, 215–234.
CPU times associated with different algorithms is shown in Tables Fread, D.L., 1976. Theoretical Development of an Implicit Dynamic Routing Model.
National Oceanic and Atmoshperic Administration, National Weather Service.
1 and 2. The tables show that the CPU time of the proposed method Freeman, T.G., 1991. Calculating catchment area with divergent flow based on a
was significantly less than that of the conventional algorithm espe- regular grid. Comput. Geosci. 17, 413–422.
cially applied to a steep watershed using a high-resolution DEM Green, W.H., Ampt, G.A., 1911. Studies on soil physics. I: The flow of air and water in
soils. J. Agric. Sci. 4, 1–24.
dataset.
Huang, I., Song, C.C.S., 1985. Stability of dynamic flood routing schemes. J. Hydr.
Eng., ASCE 111 (12), 1497–1506.
6. Conclusion Hydrologic Engineering Center, 2000. Hydrologic Modeling System HEC-HMS,
Technical Reference Manual.
Lee, K.T., 1998. Generating design hydrographs by DEM assisted geomorphic runoff
To reduce the lengthy computing time required to perform an simulation: a case study. J. Am. Water Resour. Assoc. 34 (2), 375–384.
explicit numerical scheme, a recursive formulation combined with Li, R.-M., Simons, D.B., Stevens, M.A., 1975. Nonlinear kinematic wave
approximation for water routing. Water Resour. Res. 11 (2), 245–252.
a novel routing sequence were developed for DEM-based runoff O’Callaghan, J., Mark, D.M., 1984. The extraction of drainage networks from
routing in natural watersheds. The routing sequence was set to fol- digital elevation data. Comput. Vision Graph. Image Process. 28 (3), 323–
low either the grid elevation in descending order for KW routing or 344.
Orlandini, S., Moretti, G., Franchini, M., Aldighieri, B., Testa, B., 2003. Path-based
the water stage in descending order for DFW routing. The MacCor-
methods for the determination of nondispersive drainage directions in grid-
mack recursive formulation can avoid numerical unstable despite based digital elevation models. Water Resour. Res. 39 (6), 1144. http://
the CFL condition being violated. The proposed method was ap- dx.doi.org/10.1029/2002WR001639.
Pratt, V. 1972. Shellsort and Sorting Networks, Ph.D. Dissertation, Computer Science
plied to two watersheds to test its capability for improving compu-
Department, Stanford Univ., US.
tational efficiency with a minimal simulation error. Computational Quinn, P., Beven, K., Chevallier, P., Planchon, O., 1991. The prediction of hillslope
efficiency was shown to be significantly enhanced, especially for flow paths for distributed hydrological modeling using digital terrain models.
steep watersheds. In a test using the example of a storm over a Hydrol. Process. 5, 59–80.
MacCormack, R.W., 1982. A numerical method for solving the equations of
steep watershed, the computing time for performing the proposed compressible viscous flow. AIAA J. 20 (9), 1275–1281.
KW routing method was 1/30 that of performing the conventional Tarboton, D.G., 1997. A new method for the determination of flow directions and
algorithm. For DFW routing, the computing time for the proposed upslope areas in grid digital elevation models. Water Resour. Res. 33 (2), 309–
319.
method was 1/117 that of the conventional algorithm. The pro- Weinmann, P.E., Laurenson, E.M., 1979. Approximate flood routing methods: a
posed numerical method is considered computationally efficient review. J. Hydr. Eng., ASCE 105 (12), 1521–1537.
and offers promise in its application to watershed runoff routing.

You might also like