Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Overland Flow Time of Concentration

on Very Flat Terrains


Ming-Han Li and Paramjit Chibber
The general formula can be expressed as follows:

Two types of laboratory experiments were conducted to measure overland flow times on surfaces with very low slopes. One was a rainfall test
using a mobile artificial rainfall simulator; the other was an impulse
runoff test. Test plots were 6 ft (1.83 m) wide by 30 ft (9.14 m) long with
slopes ranging from 0.24% to 0.48%. Surface types tested include bare
clay, lawn (short grass), pasture (tall grass), asphalt, and concrete. A
regression analysis was conducted to construct models for predicting flow
times. Results predicted with regressed models were compared with
those from empirical models in the literature. It was found that the slope
variable in the regressed model from rainfall test data is less influential
than that in existing models. Furthermore, the exponent for the slope
variable in the regressed model for the impulse runoff condition is only
1/10th of those in existing models. Overall, most empirical models underestimate overland flow time for laboratory plots with very low slopes.
The slope variable becomes insignificant in governing overland flow
time when the slope is small. Antecedent soil moisture, not included in
most empirical models, significantly affects time of concentration, which
is included in the regressed models.

tc = kLa n b S yi z

(1)

where
tc = time of concentration,
k = constant, and
a, b, y, and z = exponents.
This equation exhibits a linear relationship for the logarithms of the
variables involved.
As indicated in Table 1, all models include the surface slope
variable, S, raised to a negative power ranging from 0.2 to 0.5
that is, the slope variable is in the denominator and, therefore, the
greater the surface slope, the less the time of concentration and
vice versa. The problem occurs when such empirical models are to
be used for estimating time of concentration on very flat terrains
such as coastal plains. If existing empirical models are used, as the
surface slope approaches zero the time of concentration approaches
infinity. Engineering judgment suggests that such long times of
concentration do not represent reality.

Time of concentration is a primary basin parameter that represents the


travel time from the hydraulically furthest point in a watershed to the
outlet. The accuracy of estimation of peak discharge is sensitive to
the accuracy of the estimated time of concentration (1). Therefore,
the importance of an accurate estimate of the time of concentration
cannot be overlooked. For empirical models that estimate time of
concentration, one of the major variables computed is a nominal
value of slope for the watershed. Previous literature (25) collected
and listed most of the popular time-of-concentration formulas. All
these formulas (622) are empirical models and are presented in
Table 1. These models share the same general format formed by
four variables:

METHODOLOGY
Rainfall Test
A rainfall test was conducted using a mobile artificial rainfall simulator (Figure 1) where test surfaces of low slope were prepared.
The simulator included a hydroseeder with a 500-gal (1,890-liter)
water tank and an 18-in.-high (0.46-m) rain rack with spray nozzles
mounted in a 5-ft (1.52-m) spacing. This rainfall simulator was
designed to cover the entire test plot. The hydroseeder pump generated a flow ranging from 15 to 41 gal/min (0.0009 to 0.0026 m3/s).
With this range of flow rates, the precipitation rates generated by the
spray nozzles ranged from 1.5 to 3.3 in./h (38.1 to 83.8 mm/h).
Each test plot measured 6 ft (1.83 m) wide by 30 ft (9.14 m) long.
This size allowed each rainfall test to run for more than 45 min using
the maximum-capacity 500-gal water tank. A 45-min rainfall duration
was used to ensure that each test would reach the time of concentration.
Each rainfall test use the following procedure:

Length of the watershed, L;


Surface roughness (usually Mannings roughness coefficient
for overland flow), n;
Slope of the watershed, S; and
Rainfall intensity, i.

M.-H. Li, Department of Landscape Architecture and Urban Planning, Texas


A&M University and Texas Transportation Institute, 3137 TAMU, College Station, TX 77843-3137. P. Chibber, PBS&J, 1211 Indian Creek Court, Beltsville,
MD 20705. Corresponding author: M.-H. Li, minghan@tamu.edu.

Measure surface soil moisture (except on concrete and asphalt


surfaces),
Determine precipitation rate,
Start hydroseeder pump,
Observe the first flow of runoff at the outlet of the test plot,

Transportation Research Record: Journal of the Transportation Research Board,


No. 2060, Transportation Research Board of the National Academies, Washington,
D.C., 2008, pp. 133140.
DOI: 10.3141/2060-15

133

134

TABLE 1

Transportation Research Record 2060

Summary of Time of Concentration Models

Publication and Year

Equation for Time of Concentration (min)

Williams (1922) (6)

tc = 60LA D S
L = basin length, mi
A = basin area, mi2
D = diameter (mi) of a circular basin of area
S = basin slope, %
tc = KL0.77Sy
L = length of channel/ditch from headwater to outlet, ft
S = average watershed slope, ft/ft
For Tennessee, K = 0.0078 and y = 0.385
For Pennsylvania, K = 0.0013 and y = 0.5
tc = 0.8275 (LN)0.467S 0.233
L = overland flow length, ft
S = overland flow path slope, ft/ft
N = flow retardance factor
tc = 41.025(0.0007i + c)L0.33S 0.333 i 0.667
i = rainfall intensity, in./h
c = retardance coefficient
L = length of flow path, ft
S = slope of flow path, ft/ft
tc = 300L0.5S 0.5
L = basin length, mi
S = basin slope, ft/mi
tc = 60(11.9L3/H)0.385
L = length of longest watercourse, mi
H = elevation difference between divide and outlet, ft
If expressed as Tc = kLanbSyiz format: tc = KL0.77S 0.385
K = conversion constant
tc = 0.94(Ln)0.6S0.3i 0.4
L = length of overland flow, ft
n = Mannings roughness coefficient
S = overland flow plane slope, ft/ft
i = rainfall intensity, in./h
tc = 0.94L0.6 n0.6S 0.3i 0.4
L = length of overland flow, ft
n = Manning roughness coefficient
S = average overland slope, ft/ft
i = rainfall intensity, in./h
tc = 1.8(1.1 C)L0.5S 0.333
C = rational method runoff coefficient
L = length of overland flow, ft
S = surface slope, ft/ft
tc = (1/60)(L/V)
L = length of flow path, ft
V = average velocity in ft/s for various surfaces
(The exponent of S, if converted from Mannings equation, will be 0.5)
tc = 0.66L0.5n0.52S 0.31i 0.38
L = length of flow path, ft
n = roughness coefficient
S = average slope of flow path, ft/ft
i = rainfall intensity, in./h
tc = 0.595(3.15)0.33kC0.33L0.33(2k)S 0.33i 0.33(1+k)
For water at 26C
C, k = constants (for smooth paved surfaces, C = 3, k = 0.5. For grass, C = 1, k = 0)
L = length of overland plane, m
S = slope of overland plane, m/m
i = net rainfall intensity, mm/h
tc = 0.702(1.1 C)L0.5S 0.333
C = rational method runoff coefficient
L = length of overland flow, m
S = surface slope, m/m
tc = 0.0526[(1000/CN) 9]L0.8S 0.5
CN = curve number
L = flow length, ft
S = average watershed slope, %

Kirpich (1940) (7)

Hathaway (1945) (8),


Kerby (1959) (9)

Izzard (1946) (10)

Johnstone and Cross


(1949) (11)
California Culvert
Practice (1955) (12)

Henderson and Wooding


(1964) (13)

Morgali and Linsley


(1965) (14), Aron and
Erborge (1973) (15)

FAA (1970) (16)

U.S. Soil Conservation


Service (1975, 1986)
(17, 18)
Papadakis and Kazan
(1986) (2)

Chen and Wong (1993)


(19), Wong (2005) (20)

TxDOT (1994) (21)

Natural Resources
Conservation Service
(1997) (22)

0.4

1 0.2

NOTE: 1 mi = 1.61 km; 1 ft = 0.3048 m; 1 in. = 25.4 mm.

Remarks
The basin area should be smaller
than 50 mi2 (129.5 km2).

Developed for small drainage basins


in Tennessee and Pennsylvania,
with basin areas from 1 to 112
acres (0.40 to 45.3 ha).
Drainage basins with areas of less
than 10 acres (4.05 ha) and slopes
of less than 0.01.
Hydraulically derived formula;
values of c range from 0.007 for
very smooth pavement to 0.012
for concrete pavement to 0.06 for
dense turf.
Developed for basins with areas
between 25 and 1624 mi2
(64.7 and 4206.1 km2).
Essentially the Kirpich (7) formula;
developed for small mountainous
basins in California.

Based on kinematic wave theory for


flow on an overland area.

Overland flow equation from


kinematic wave analysis of runoff
from developed areas.

Developed from airfield drainage


data assembled by U.S. Corps of
Engineers.
Developed as a sum of individual
travel times. V can be calculated
using Mannings equation.
Developed from USDA Agricultural
Research Service data of 84 small
rural watersheds from 22 states.

Overland flow on test plots of


1 m wide by 25 m long. Slopes of
2% and 5%.

Modified from FAA (16).

For small rural watersheds.

Li and Chibber

135

Weir for overflow

Runoff direction

Reservoir

FIGURE 2

Reservoir used for impulse runoff test.

Tested Surfaces and Number of Tests


FIGURE 1 Rainfall test using artificial
rainfall simulator.

Monitor runoff rate continuously,


Shut down hydroseeder pump after the runoff peaked for about
10 min, and
Continue measuring runoff rate until runoff ceased.

Impulse Runoff Test


The impulse runoff test measured runoff travel time without artificial
rainfall. Both pervious and impervious surfaces were tested. For
pervious surface testing, test plots were kept saturated before testing.
The notion was to minimize the effect of antecedent soil moisture
and infiltration and focus on the effects of surface type, flow rate, and
slope on travel time. Meanwhile, time of concentration was measured as the time for water to travel from the farthest point of the plot
to the outlet.
As with the rainfall test, researchers used a hydroseeder as the
mobile water source. A large reservoir stored water filled by the
hydroseeder. This reservoir, during an impulse runoff test, was
placed on the upstream side of the test plot and as the reservoir
capacity was exceeded, water overflowed the weir onto the plot
(Figure 2). The hydroseeder pump generated a flow ranging from
about 15 to 41 gal/min (0.0009 to 0.0026 m3/s). The size of each
test plot again measured 6 ft (1.83 m) wide by 30 ft (9.14 m) long,
as in the rainfall test. Each impulse runoff test used the following
procedure:

Five surface types were tested in this study. The details of each test
type are described below:
Bare clay (two plots). The texture for the soil tested was 21%
sand, 31% silt, and 48% clay. The slopes of these two bare clay plots
were 0.43% and 0.42%, respectively.
Lawn (two plots). The lawn was established using common
Bermudagrass (Cynodon dactylon). A grass height of less than 2 in.
(5.1 mm) was maintained by mowing. The slopes of these two lawn
plots were 0.48% and 0.24%, respectively.
Pasture (two plots). The lawn was also formed using common
Bermudagrass (C. dactylon). Grasses of 6 in. (0.15 m) or taller were
maintained to simulate pasture conditions. The slopes of these two
pasture plots were 0.48% and 0.24%, respectively.
Concrete (one plot). The plot was built on the taxiway of an old
airport. The slope was 0.35%.
Asphalt (one plot). An asphaltic cold mix was applied on top
of an old airport taxiway to simulate the asphalt surface. The slope
was 0.35%.
The number of tests for different surfaces is presented in Table 2.
The data and the range of that specific data category collected in this
study are summarized in the following list:
Slope: 0.24% to 0.48%;
Rainfall intensity (for rainfall test): 1.5 to 3.3 in./h (38.1 to
83.8 mm/h);

TABLE 2

Number of Tests

Tested Surface

Water pervious test plots, including bare clay and pasture, to


create a saturated condition;
Determine flow rate;
Start hydroseeder pump to generate flow;
Record time when water overtopped the weir; and
Record travel time when water front reached the outlet.

Bare clay
Lawn
Pasture
Concrete
Asphalt

Rainfall Test

Impulse Runoff Test

11
13
12
10

15
Not tested
13
5

16

136

Transportation Research Record 2060

Runoff flow rate (for impulse runoff test): 15 to 41 gal/min


(0.0009 to 0.0026 m3/s);
Antecedent soil moisture (bare clay, lawn, and pasture): 8%
to 54%;
Infiltration rate: 0.0358 to 0.0598 in./h (0.91 to 1.52 mm/h);
Soil texture: 21% sand, 31% silt, and 48% clay (bare clay, lawn,
and pasture plots); and
Raindrop size: 0 to 0.07 in. (0 to 1.8 mm) in diameter.

RESULTS AND DISCUSSION


Rainfall Test
A typical hydrograph of the rainfall test observed in this study is
shown in Figure 3. A rainfall test began at time zero when the artificial rainfall simulator was turned on. Initially, there was a no-flow
period at the outlet. Time of beginning was recorded when the
first flow was observed at the outlet. As the flow plateaued and
fluctuated within 5% of the flow rate, the rate was considered as
the peak, which determined the time to peak. Because time of concentration should involve only hydraulic travel time, the initial
loss process (initial abstraction) was not considered as part of the
time of concentration. Therefore, the time of concentration in the
rainfall test was determined as the time to peak minus the time of
beginning of runoff.
The results of rainfall tests are presented in Table 3. A linear regression analysis on the rainfall test data was conducted to construct a
model to predict the time of concentration. From a preliminary
analysis, it was found that the antecedent soil moisture was negatively
associated with the time of concentration (Figure 4). Therefore, the
antecedent soil moisture variable, , was added to Equation 1 to
create the new model as follows:
tc = kLa n b x S yi z

(2)

where is antecedent soil moisture (percent) and x is an exponent


of .
In the new model, the watershed length L is 30 ft (9.14 m), which
is the length of test plots. For the exponent of length L, the mean value,
0.5, of several models compiled by Papadakis and Kazan (2) was used.
Similarly, regression analysis was conducted to construct models
for the time of beginning, tb, and time to peak, tp, using the same
initial model shown in Equation 1.
The best-fit models for tc, tb, and tp are summarized in Table 4.
Signs and values of exponents for each variable in all three models

Time to Peak
Q
Time of Concentration

Time
Time of Beginning

FIGURE 3

Typical hydrograph in a rainfall test.

are generally similar. When compared with existing models, exponents of variables n, , and i are close to those of existing models
and they are all statistically significant (P < 0.001). However, exponents of the slope variable, S, were found to be insignificant and
their absolute values are less than those (0.33) from existing models.
It appears that S has the least influence on the time of concentration because the absolute values of the exponents for S are the
least among all variables. This could result from the relatively
few slopes tested in the study (0.24%, 0.35%, 0.42%, 0.43%, and
0.48%). Or, if the slope data number is not the cause, the insignificant effect of S might indicate that when the slope is very small,
the effect from variables n, , and i become dominant on the time
of concentration.
The final regressed model for time of concentration can be expressed
as follows:
tc = 0.553 L0.5n 0.32 0.277 S 0.172i 0.646

(3)

This model was compared with some existing models developed


from both overland flow data [Kerby (9), Izzard (10), Henderson
and Wooding (13), and Chen and Wong (19)] and watershed data
[Papadakis and Kazan (2) and Natural Resources Conservation
Service (22)]. Values used for variables in each model are presented
in Table 5. The comparison result is plotted in Figure 5. It is apparent
that all compared models underestimated the time of concentration
for tested conditions. Note that Izzards model predicted the time
reasonably well with the majority underestimated and the minority
overestimated.

Impulse Runoff Test


The results of impulse runoff tests are presented in Table 6. A linear
regression analysis of the impulse runoff test data was also conducted
to construct a model to predict the time of concentration. In this case,
the time of concentration is the travel time from the farthest side of
the test plot to the outlet. The initial regression model was modified
from Equation 1 as follows:
tc = kLa n bQ w S y

(4)

where
tc = time of concentration (min),
L = watershed length (ft) [30 ft (9.14 m) in this
study],
n = Mannings roughness coefficient for overland
flow,
Q = flow rate (gal/min), and
a, b, w, and y = exponents.
Again, the exponent (a) of variable L was set as 0.5, which is the
mean value compiled by Papadakis and Kazan (2) from several
existing models.
The final model is presented in Table 7. Exponents of n and Q are
statistically significant, whereas the exponent of S is insignificant,
similar to the rainfall test result. Furthermore, the exponent for the
slope variable in the regressed model for the impulse runoff condition
is only one-tenth of those in existing models. Again, the slope variable becomes insignificant in governing overland flow time when the
slope is small.

Li and Chibber

137

TABLE 3

Rainfall Test Results

Surface
Bare Clay 1
Bare Clay 2
Bare Clay 3
Bare Clay 4
Bare Clay 5
Bare Clay 6
Bare Clay 7
Bare Clay 8
Bare Clay 9
Bare Clay 10
Bare Clay 11
Lawn 1
Lawn 2
Lawn 3
Lawn 4
Lawn 5
Lawn 6
Lawn 7
Lawn 8
Lawn 9
Lawn 10
Lawn 11
Lawn 12
Lawn 13
Pasture 1
Pasture 2
Pasture 3
Pasture 4
Pasture 5
Pasture 6
Pasture 7
Pasture 8
Pasture 9
Pasture 10
Pasture 11
Pasture 12
Concrete 1
Concrete 2
Concrete 3
Concrete 4
Concrete 5
Concrete 6
Concrete 7
Concrete 8
Concrete 9
Concrete 10
Asphalt 1
Asphalt 2
Asphalt 3
Asphalt 4
Asphalt 5
Asphalt 6
Asphalt 7

Time of
Beginning (min)

Time to
Peak (min)

Antecedent Soil
Moisture (%)

Rainfall Intensity
(mm/h)

Slope
(%)

6
15
6
2
5
22
14
15
10
38
27
9
14
9
9
9
13
7
11
14
10
18
8
10
18
23
11
16
44
18
49
8
15
16
26
17
4
2
2
1
3
2
2
3
3
1
1
1
4
1
1
3
2

14
27
20
12
15
45
37
37
36
71
61
34
54
34
38
36
32
24
55
47
40
57
34
39
50
63
39
49
91
60
95
46
44
51
61
47
12
18
15
16
10
10
14
12
9
11
10
10
14
10
12
10
12

41
20
19
53
41
12
17
17
23
9
10
41
42
44
55
31
44
56
33
19
28
34
42
37
30
26
28
25
18
24
23
38
34
30
30
34
N.A.
N.A.
N.A.
N.A.
N.A.
N.A.
N.A.
N.A.
N.A.
N.A.
N.A.
N.A.
N.A.
N.A.
N.A.
N.A.
N.A.

47
50
54
85
58
43
38
41
41
40
41
38
38
53
46
78
87
83
54
49
56
47
47
52
49
33
75
51
40
42
54
65
53
41
44
39
38
35
32
36
53
56
43
41
44
32
38
41
37
45
43
47
49

0.43
0.43
0.43
0.43
0.43
0.42
0.42
0.42
0.42
0.42
0.42
0.48
0.48
0.48
0.48
0.48
0.48
0.48
0.24
0.24
0.24
0.24
0.24
0.24
0.48
0.48
0.48
0.48
0.48
0.48
0.24
0.24
0.24
0.24
0.24
0.24
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35

Transportation Research Record 2060

60
2

R = 0.6858

40
20
0
0

10

20

30

60

Ant. moisture content


(%)

Ant. moisture content


(%)

138

40
20

R2 = 0.2972

0
0

40

Time of conc. (min)

20

40

60

Time of conc. (min)

(b)
Ant. moisture content
(%)

(a)
40

20

R2 = 0.631

0
0

20

40

60

Time of conc. (min)

(c)
FIGURE 4 Relationship between antecedent moisture content and time of concentration
(rainfall tests): (a) bare clay, (b) lawn, and (c) pasture.

TABLE 4

Regressed Models (Rainfall Test)

R2

Constant

b (for n)

x (for )

y (for S )

z (for i)

tc

53

0.86

tb

53

0.85

tp

53

0.92

0.553
(0.115)a
0.196
(0.753)
0.692
(0.026)

0.320
(<0.001)
0.326
(<0.001)
0.305
(<0.001)

0.277
(<0.001)
0.942
(<0.001)
0.474
(<0.001)

0.172
(0.192)
0.016
(0.947)
0.139
(0.228)

0.646
(<0.001)
0.415
(0.119)
0.631
(<0.001)

p-value in parentheses.

TABLE 5

Surface
Bare clay
Lawn
Pasture
Concrete
Asphalt
a

Variable Values Used for Model Comparison (Rainfall Test)

Kerby Model

Izzard Model

Chen and Wong


Model

N (flow
retardance factor)

c (retardance
coefficient)

C and k
(constant)

0.1
0.3
0.4
0.02
0.02

0.02
0.046
0.06
0.012
0.007

Natural Resources Conservation Service.

Not provided
C = 1, k = 0
Not provided
C = 3, k = 0.5
C = 3, k = 0.5

Papadakis and Kazan/


Henderson and
Wooding Models

NRCSa
Model

n (Manning
roughness n)

Curve
Number

0.0425
0.24
0.41
0.012
0.014

89
78
80
98
98

Li and Chibber

139

TABLE 6

50

Impulse Runoff Test Results

Surface

Observed Tc (min)

40

Bare clay 1
Bare clay 2
Bare clay 3
Bare clay 4
Bare clay 5
Bare clay 6
Bare clay 7
Bare clay 8
Bare clay 9
Bare clay 10
Bare clay 11
Bare clay 12
Bare clay 13
Bare clay 14
Bare clay 15
Pasture 1
Pasture 2
Pasture 3
Pasture 4
Pasture 5
Pasture 6
Pasture 7
Pasture 8
Pasture 9
Pasture 10
Pasture 11
Pasture 12
Pasture 13
Concrete 1
Concrete 2
Concrete 3
Concrete 4
Concrete 5
Asphalt 1
Asphalt 2
Asphalt 3
Asphalt 4
Asphalt 5
Asphalt 6
Asphalt 7
Asphalt 8
Asphalt 9
Asphalt 10
Asphalt 11
Asphalt 12
Asphalt 13
Asphalt 14
Asphalt 15
Asphalt 16

30

This study

20

Izzard
Kerby
Henderson-Wooding
10

Chen-Wong
Papadakis-Kazan
NRCS

0
0

10

20
30
Predicted Tc (min)

40

50

FIGURE 5 Observed versus predicted time of concentration in the


rainfall test by seven models: this study, Papadakis and Karan (2),
Kerby (9), Izzard (10), Henderson and Wooding (13), Chen and
Wong (19), and Natural Resources Conservation Service
(NRCS) (22).

The observed times of concentration were also plotted against those


predicted by the new and existing models (Figure 6). Values used
for variables in each model are presented in Table 8. The Kirpich (7)
and Natural Resources Conservation Service (22) models underestimated the observed times of concentration, the Texas Department
of Transportation (21) overestimated them. The lower left cluster of
data in Figure 6 indicates the test results on smooth surfaces (bare
clay, asphalt, and concrete), and the higher clusters indicate the results
on pasture. On the basis of comparison results, existing models
cannot correctly predict the time of concentration on flat surfaces.

Limitations
The limitations associated with the laboratory experiment are
summarized as follows:
Data were collected only from laboratory experiments. Field
studies are needed to verify the laboratory results.
Only a few flat slopes were tested.
The antecedent soil moisture had a significant effect on the time
of concentration; however, changes in soil moisture during tests
were not monitored.

Travel Time
(min)

Flow Rate
(m3/s)

Slope
(%)

1.28
1.17
0.70
1.02
1.31
1.20
0.73
1.80
1.72
2.27
1.72
2.25
2.23
2.42
2.67
5.60
4.88
5.00
4.07
5.50
6.92
5.53
7.00
6.28
5.53
4.33
4.07
5.83
1.32
1.28
1.33
1.35
1.23
1.02
1.10
1.12
1.33
1.02
1.30
1.02
1.42
1.22
1.08
1.07
1.07
1.15
1.18
1.23
1.63

0.0011
0.0018
0.0023
0.0013
0.0017
0.0016
0.0026
0.0019
0.0015
0.0012
0.0018
0.0020
0.0017
0.0015
0.0011
0.0013
0.0013
0.0013
0.0013
0.0013
0.0016
0.0014
0.0015
0.0009
0.0015
0.0013
0.0019
0.0019
0.0015
0.0015
0.0013
0.0010
0.0009
0.0011
0.0012
0.0007
0.0012
0.0017
0.0011
0.0026
0.0012
0.0008
0.0009
0.0015
0.0015
0.0016
0.0020
0.0023
0.0012

0.43
0.43
0.43
0.43
0.43
0.43
0.43
0.42
0.42
0.42
0.42
0.42
0.42
0.42
0.42
0.24
0.24
0.24
0.24
0.24
0.24
0.24
0.48
0.48
0.48
0.48
0.48
0.48
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35
0.35

conducted. The results indicate that the slope variable was insignificant in governing overland flow time on low-slope surfaces. In
addition, antecedent soil moisture was negatively associated with
the time of concentration and was as influential as surface roughness,
rainfall intensity, and flow rate to the time of concentration of
low-slope surfaces. This finding suggests that soil moisture should

CONCLUSIONS
Existing empirical models of estimating the time of concentration
were mostly developed from data of common watersheds. The ability of these models to predict the time of concentration on very flat
terrains has not been fully analyzed. This study used a controlled
laboratory environment to measure times of concentration on test
plots 6 ft (1.83 m) wide by 30 ft (9.14 m) long with slopes of less
than 0.5%. Artificial rainfall tests and impulse runoff tests were

TABLE 7

tc

Regressed Model (Impulse Runoff Test)

R2

Constant

b (for n)

w (for Q)

y (for S)

49

0.86

0.779
(0.185)a

0.453
(<0.001)

0.537
(<0.001)

0.037
(0.848)

p-value in parentheses.

140

Transportation Research Record 2060

Observed Tc (min)

4
This study
Kirpich

TxDOT
NRCS

0
0

10

12

14

16

Predicted Tc (min)
FIGURE 6 Observed versus predicted time of concentration in the impulse runoff
test by four models: this study, Kirpich (7), Texas Department of Transportation
(TxDOT) (21), and Natural Resources Conservation Service (NRCS) (22).

TABLE 8 Variable Values Used for Model Comparison


(Impulse Runoff Test)
Surface
Bare clay
Pasture
Concrete
Asphalt

4.

Kirpich Model
K

TxDOT Model
C (runoff coefficient)

NRCS Mode
Curve Number

0.0078
0.0078
0.0078
0.0078

0.2
0.15
0.925
0.875

89
80
98
98

5.
6.
7.
8.
9.
10.

be considered in design and a moderate value should be chosen


for risk balancing.
The findings of this study indicate that a direction for future work
would be through field-scale instrumentation and observation of a
number of watersheds with low slopes. The rationale for performing
the laboratory study was the lack of appropriate data as well as the
desire to develop a relationship between the time of concentration
and the explanatory variables without having to collect several years
of data. In light of the realization that antecedent soil moisture can
affect the time of concentration, some representative soil moisture
content data would also be useful.

11.
12.
13.

14.
15.

16.

ACKNOWLEDGMENT

17.

This study was sponsored by the Texas Department of Transportation.

18.
19.

REFERENCES
20.
1. McCuen, R. H., S. L. Wang, and W. J. Rawls. Estimating Urban Time of
Concentration. Journal of Hydraulic Engineering, Vol. 110, No. 7, 1984,
pp. 887904.
2. Papadakis, C. N., and M. N. Kazan. Time of Concentration in Small
Rural Watersheds. Technical Report 101/08/86/CEE. Civil Engineering
Department, University of Cincinnati, Cincinnati, Ohio, 1986.
3. Goitom, T. G., and M. Baker, Jr. Evaluation of Tc Methods in a Small Rural
Watershed. In Channel Flow and Catchment Runoff: Proc. International
Conference for Centennial of Mannings Formula and Kuichlings

21.
22.

Rational Formula (B. C. Yen, ed.), University of Virginia, Charlottesville,


1989, pp. 8896.
Mays, L. W. Water Resources Engineering. John Wiley and Sons,
New York, 2001.
Computer Applications in Hydraulic Engineering, 5th ed. Haestad
Methods, Inc., Watertown, Conn., 2002.
Williams, G. B. Flood Discharges and the Dimensions of Spillways in
India. Engineering (London), Vol. 134, 1922, p. 321.
Kirpich, Z. P. Time of Concentration of Small Agricultural Watersheds.
Civil Engineering, Vol. 10, No. 6, 1940, p. 362.
Hathaway, G. A. Design of Drainage Facilities. Transactions of the
American Society of Civil Engineers, Vol. 110, 1945, p. 697.
Kerby, W. S. Time of Concentration Studies. Civil Engineering, March,
1959.
Izzard, C. F. Hydraulics of Runoff from Developed Surfaces. Proceedings
of the Highway Research Board, Vol. 26, 1946, pp. 129150.
Johnstone, D., and W. P. Cross. Elements of Applied Hydrology. Ronald
Press, New York, 1949.
California Culvert Practice, 2nd ed. Department of Public Works,
Division of Highways, Sacramento, 1955.
Henderson, F. M., and R. A. Wooding. Overland Flow and Groundwater
Flow from a Steady Rain of Finite Duration. Journal of Geophysical
Research, Vol. 69, No. 8, 1964, pp. 15311540.
Morgali, J. R., and R. K. Linsley. Computer Analysis of Overland Flow.
Journal of the Hydraulics Division, Vol. 91, No. HY3, 1965, pp. 81100.
Aron, G., and C. E. Erborge. A Practical Feasibility Study of Flood Peak
Abatement in Urban Areas. Report. U.S. Army Corps of Engineers,
Sacramento, Calif., 1973.
Circular on Airport Drainage. Report A/C 150-5320-5B. Federal
Aviation Administration, U.S. Department of Transportation, Washington,
D.C., 1970.
Urban Hydrology for Small Watersheds. Technical Release 55. U.S.
Soil Conservation Service, Washington, D.C., 1975.
Urban Hydrology for Small Watersheds. Technical Release 55, 2nd ed.
U.S. Soil Conservation Service, Washington, D.C., 1986.
Chen, C. N., and T. S. W. Wong. Critical Rainfall Duration for Maximum
Discharge from Overland Plane. Journal of Hydraulic Engineering,
Vol. 119, No. 9, 1993, pp. 10401045.
Wong, T. S. W. Assessment of Time of Concentration Formulas for
Overland Flow. Journal of Irrigation and Drainage Engineering, Vol. 131,
No. 4, 2005, pp. 383387.
Hydraulic Design Manual. Texas Department of Transportation, Austin,
1994.
PondsPlanning, Design, Construction. Agriculture Handbook No. 590.
U.S. Natural Resources Conservation Service, Washington, D.C., 1997.

The Hydrology, Hydraulics, and Water Quality Committee sponsored publication


of this paper.

You might also like