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

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/327601362

Journal of Hydraulic Engineering

Article in Journal of Hydraulic Engineering · September 2018

CITATIONS READS
2 693

4 authors, including:

Eunjin Han
Inje University
9 PUBLICATIONS 15 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Advection & Diffusion View project

ECO-RIVER OF RESULTS View project

All content following this page was uploaded by Eunjin Han on 12 September 2018.

The user has requested enhancement of the downloaded file.


Analysis of Two-Dimensional Mixing in Natural Streams
Based on Transient Tracer Tests
Il Won Seo, A.M.ASCE 1; Hwang Jeong Choi 2; Young Do Kim 3; and Eun Jin Han 4

Abstract: To investigate the two-dimensional dispersion characteristics of the dissolved contaminant in natural rivers, tracer experiments
were performed based on the instantaneous injection of Rhodamine WT solution into small- to medium-sized rivers in Korea. The relations
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

between the dispersion coefficients and hydraulic and geometric variables of the river were analyzed in depth. The experimental results show
that transverse dispersion coefficient tends to increase as the friction term, U=U  , and the channel aspect ratio, W=H, increase. The effect of the
channel curvature on the rate of transverse mixing was found to be very distinct in that DT =HU  has a linear relation with UW=U  Rc , even
though the slope of the fitted line in the plot of DT =HU  versus UW=U  Rc in a log-log scale was much smaller than the slope of the equation
by Yotsukura and Sayre. The longitudinal dispersion coefficient also tends to increase as U=U  and W=H increase. However, DL is rather
insensitive to the channel curvature. The values of the longitudinal dispersion coefficient from field studies are much higher than the theoretical
value suggested by Elder. The observed longitudinal dispersion coefficient is one order of magnitude larger than Elder’s coefficient, and
DL =DT of this study ranges 26.6–332. DOI: 10.1061/(ASCE)HY.1943-7900.0001118. © 2016 American Society of Civil Engineers.
Author keywords: Pollutant mixing; Two-dimensional transport model; Longitudinal and transverse dispersion coefficients; Tracer tests;
Routing procedures; Channel curvature.

Introduction The two-dimensional advection-dispersion equation (2D ADE)


can be derived by integrating the three-dimensional transport
In large rivers of many countries, river waters are used for the equation for turbulent flow over the channel depth (Fischer et al.
water supply of municipal and agricultural uses. Because these 1979). The 2D ADE for natural streams, in which the principle flow
rivers flow through large cities, the rivers are prone to pollutant direction changes along the channel direction, is written in the cur-
spills and accidents. Furthermore, there are many sewage treatment vilinear coordinate system as (Seo et al. 2008; Lee and Seo 2013)
plants from which wastewater effluent is discharged into rivers.    
When a pollutant is introduced into a river, it undergoes different ∂C ∂C ∂C 1 ∂ ∂C 1 ∂ ∂C
þu þv ¼ hDL þ hDT ð1Þ
mixing stages, as shown in Fig. 1. In the first stage (near field), the ∂t ∂x ∂y h ∂x ∂x h ∂y ∂y
vertical mixing is rapidly completed in most natural rivers, of which
the depth is much smaller than the width of the channel. Then, in where C = depth-averaged concentration; u and v = depth-averaged
the next stage, the mixing occurs in transverse and longitudinal velocities in the x and y directions, respectively; h = local depth;
directions. This stage is referred to the intermediate field, which DL = longitudinal dispersion coefficient; and DT = transverse
is located in the middle reach between the near field and the far dispersion coefficient. In this equation, as shown in Fig. 1, x means
field, as shown in Fig. 1. When pollutants are introduced into rivers, the stream-wise direction, and y is the coordinate which is normal
the mixing stage of the intermediate field is very long compared to to the flow direction, approximately regarded as the span-wise
the first stage. Thus, in this case, the two-dimensional transport direction.
model, which contains both the advection and dispersion terms, In using 2D ADE for the analysis of mixing in two directions in
needs to be used for the analysis of both transverse and longitudinal shallow flows of rivers and estuaries, both longitudinal and trans-
mixing of polluted cloud in the intermediate region. verse dispersion coefficients should be provided. Between the two
dispersion coefficients of 2D ADE, the transverse dispersion of
1 river mixing has been actively studied by many researchers (Fischer
Professor, Dept. of Civil and Environmental Engineering, Seoul
National Univ., 1 Gwanak-ro, Gwanak-gu, Seoul 151-744, South Korea. et al. 1979; Rutherford 1994; Baek and Seo 2013). In most of the
E-mail: seoilwon@snu.ac.kr field studies on transverse mixing in natural rivers, researchers,
2
Ph.D. Student, Dept. of Civil and Environmental Engineering, Seoul such as Yotsukura et al. (1970), Yotsukura and Cobb (1972), Holley
National Univ., 1 Gwanak-ro, Gwanak-gu, Seoul 151-744, South Korea and Abraham (1973), Sayre (1979), Beltaos (1980), Lau and
(corresponding author). E-mail: hjeong@snu.ac.kr Krishnappan (1981), Holly and Nerat (1983), and Pernik and
3
Associate Professor, Dept. of Environmental Science and Engineering, Roberts (1985), evaluated the transverse dispersion coefficient
Nakdong River Environmental Research Center, Inje Univ., 197 Inje-ro, based on tracer tests or tracing studies conducted under steady state
Gimhae, Gyengnam 50834, South Korea. conditions. The values of the transverse dispersion coefficient ob-
4
Postdoctoral Research Associate, Nakdong River Environmental served in this way show wide ranges, according to the hydraulic
Research Center, Inje Univ., 197 Inje-ro, Gimhae, Gyengnam 50834,
and geometric characteristics of the river as summarized by Jeon
South Korea.
Note. This manuscript was submitted on October 17, 2014; approved on et al. (2007).
October 14, 2015; published online on April 25, 2016. Discussion period Unlike the transverse dispersion coefficient, the longitudinal
open until September 25, 2016; separate discussions must be submitted for dispersion coefficient of 2D ADE, DL , has not yet been studied
individual papers. This paper is part of the Journal of Hydraulic Engineer- abundantly. This is because most of the field studies for river mix-
ing, © ASCE, ISSN 0733-9429. ing during the 1970s and 1980s were conducted under steady-state

© ASCE 04016020-1 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


Vertical, Transverse and Transverse and where S = cross-sectional mean concentration; V = cross-sectional
Longitudinal Mixing Longitudinal Dispersion
(3-D analysis) (2-D analysis) mean velocity; and K = bulk longitudinal dispersion coefficient,
C Longitudinal Dispersion which should encompass the combined effects of the lateral veloc-
Pollutant (1-D analysis) ity shear and the vertical shear (Fischer et al. 1979). As summarized
C
Injection C by Seo and Cheong (1998) and Wallis and Manson (2004), K of the
1D ADE has long been investigated. Thus, a number of times, the
literature has reported the observed values of K in natural rivers and
formulae for the prediction of K. However, as Kalinowska and
Rowinski (2012) maintained, the longitudinal dispersion coeffi-
cient appearing in the 2D equation, Eq. (1), is not the same param-
s eter as the longitudinal dispersion coefficient in the 1D equation,
n
Eq. (4). Thus, it is suggested that the longitudinal dispersion coef-
ficient obtained from the 2D tracer tests under the transient input
Intermediate field condition should be applied for the models of 2D ADE, even
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

though, so far, there are not enough studies to quantify the values
Fig. 1. Conceptual diagram of pollutant mixing in rivers (adapted from of DL in a reliable way.
Kilpatrick and Wilson 1989) The objective of this study is to investigate the two-dimensional
dispersion characteristics of natural rivers and to analyze the rela-
tion between dispersion coefficients and hydraulic and geometric
conditions, and then, only the transverse dispersion coefficient was variables of the river. In this study, tracer tests were performed
calculated, neglecting the longitudinal dispersion term in the 2D based on the instantaneous injection of Rhodamine WT solution
ADE [Eq. (1)]. It was only in recent years that tracer studies were into small- to medium-sized rivers in Korea. Both hydraulic and
performed under transient conditions based on the slug input of the dispersion data were collected to calculate both the longitudinal
tracer, and in that way, both longitudinal and transverse dispersion and transverse dispersion coefficients. The steady-stream-tube rout-
coefficients could be calculated from the data of 2D concentration- ing procedure (S-STRP) and the transient stream-tube routing pro-
time curves (Seo et al. 2006; Baek and Seo 2010; Sun et al. 2011). cedure (T-STRP) were used to calculate DL and DT . The effects of
For this reason, in the modeling of the 2D ADE for the analysis of the hydraulic and geometric variables on the dispersion coefficients
pollutant mixing in natural streams, for DL, the theoretical value were analyzed.
proposed by Elder (1959) for the shear flow over the infinitely wide
plane, as given in Eq. (2), has been adopted without much valida-
tion (Lee and Seo 2007; Seo et al. 2008, 2010) Theoretical Backgrounds
DL
¼ 5.93 ð2Þ Transverse Mixing Coefficient
HU 
Transverse mixing in natural streams has been studied by a number
where H = average depth of the river; and U  = shear velocity. of researchers based on field experiments in channels with different
However, the dispersion study conducted in the laboratory me- curves and irregularities (Fischer et al. 1979; Rutherford 1994).
andering channel by Baek et al. (2006) showed that DL varied in a From their results, it can be concluded that, in straight channels,
wider range than estimated by Eq. (3). In most cases in which the the transverse mixing depends on several factors, including bottom
aspect ratios (the ratio of the channel width, W to the depth, H) shear velocity, channel aspect ratio, depth variations, and other
were larger than 6, the values of DL =HU  were found in the range channel irregularities, whereas in the meandering channel, secon-
of 10–20, which is two to four times the value suggested by Elder. dary circulation in the bends may generate additional transverse
This difference was found in natural streams in a more profound transport. Based on the field data collected in channels with differ-
way. Seo et al. (2006) and Baek and Seo (2010) conducted the ent curves and irregularities, Fischer et al. (1979) suggested ranges
tracer study under slug tests in natural streams of which the aspect of the transverse dispersion coefficient considering the effect of
ratios were 60–80, a range ten times bigger than the laboratory channel curvature as
flume conditions. They reported that the dimensionless value of
DL ranges from 10 to 90. This means DL in natural streams is much DT
¼ 0.1 ∼ 0.2 for straight uniform channel
bigger than the value of the laboratory meandering channel, as well HU 
as of Elder’s theoretical equation. More recently, Sun et al. (2011), 0.3 ∼ 0.9 for gently meandering stream ð4Þ
based on the 2D tracer tests under the transient condition, reported
that DL =WU  ≈ 1 and DT =HU  ≈ 0.6 in the St. Clair River, Fischer et al. (1979) also maintained that higher values are ex-
Canada, of which the aspect ratio is approximately 80. That means pected if the channel has the sharp curves, even though values or
that DL =HU  was approximately 80, which would fall into the ranges of the transverse dispersion coefficient were not suggested.
range of the values reported by Baek and Seo (2010). Although, they did not define the exact meaning of “sharp curves”
Taking into account the difference between the observed values versus “slowly meandering,” they suggested, based on the analysis
and the theoretical value mentioned previously, one misdoing in the of data collected in Missouri River by Yotsukura et al. (1970) and
selection of DL for the modeling of 2D ADE is using the Elder’s Sayre and Yeh (1973), that a gently meandering river is one that
value without much calibration and validation, whereas another follows the condition given as follows:
misconception is using the longitudinal dispersion coefficient of
the one-dimensional advection-dispersion equation (1D ADE), U W
≤2 ð5Þ
given as follows: U  Rc
 
∂S ∂S ∂ ∂S
þV ¼ K ð3Þ where U = reach-average of the depth-averaged velocity; W =
∂t ∂x ∂x ∂x channel width; and Rc = radius of channel curvature.

© ASCE 04016020-2 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


Rutherford (1994), after analyzing reported transverse proposed a hybrid equation, in which constants of the theoretical
dispersion coefficients in straight and meandering natural channels, equation were decided by regression coefficients. Comparing re-
also proposed ranges of the transverse dispersion coefficient as sults by Boxall and Guymer (2003) and Baek and Seo (2011), they
follows: suggested the following equation:
  2 
2 1 − exp − β
DT DT
¼ 0.15 ∼ 0.30 for straight channel ¼ ðαPÞ ð12Þ
HU  HU  P
¼ 0.30 ∼ 0.90 for gently meandering channel
where P ¼ ðU=U  ÞðH=Rc Þ; and α, β = regression coefficients. By
¼ 1.00 ∼ 3.00 for sharp curved channel ð6Þ fitting to 18 dispersion sets from the field experiments (Rutherford
1994; Seo et al. 2006), both regression coefficients were decided
Rutherford did not give a clear definition of “straight,” “gently as 77.88.
meandering,” and “sharp curved.” Instead, an approximate quanti-
tative criterion for classifying channels as straight or meandering in Methods for the Calculation of Dispersion Coefficients
terms of the effects on the transverse mixing can be found in works Methods for calculation of the dispersion coefficients have been
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

of Almquist and Holley (1985). Using the equation by Yotsukura proposed by many researchers (Seo et al. 2006; Baek et al. 2006;
and Sayre (1976), they suggested the criterion for the gently me- Baek and Seo 2010). When the observed concentration data are
andering stream where the secondary currents are strong enough to available, DT can be calculated using both moment methods and
increase the transverse mixing as the routing procedures, whereas DL can be calculated using only
routing procedures (Baek and Seo 2010). Moment-based methods
U W
≥1 ð7Þ were suggested by many researchers, including Sayre and Chang
U  Rc (1968), Holley et al. (1972), Krishnappan and Lau (1977), and
Beltaos (1980). Even though moment-based methods have been
In the equations given previously, the effect of channel curvature
widely used, these methods were hampered by the restriction that
was not accounted for in an explicit manner. The effect of channel
the skewed concentration profile induced by the recirculation zones
curvature on the transverse mixing was first studied theoretically by
and channel irregularities makes it difficult to compute the mean-
Fischer (1969) in which he utilized velocity profiles of the secon-
ingful values of the second moments or variances. Furthermore,
dary currents given by Rozovskii (1957). Using the results of
when the concentration data was obtained by the transient tracer
laboratory tests, which were performed at a curved channel, he sug-
tests with the instantaneous injection, data should be converted
gested the equation given as follows:
to the steady-state condition for the application of the moment
 2  2 methods. To overcome these disadvantages, the routing procedure
DT U H
¼ 25 ð8Þ was proposed by Fischer (1968), to calculate the longitudinal
HU  U Rc
dispersion coefficient for the 1D ADE. For calculation of both
Yotsukura and Sayre (1976) maintained that a better correlation DL and DT of 2D ADE, Baek et al. (2006) developed a two-
was obtained substituting channel width (W) for water depth (H) in dimensional routing procedure based on the analytical solution
Eq. (8) based on the data sets acquired in the Missouri River around of the 2D ADE for the unsteady condition. Using the stream-tube
a bend. For field data, their result was concept, Seo et al. (2006) developed the stream-tube routing pro-
 2  2 cedure for the steady-state conditions, and Baek and Seo (2010)
DT U W derived the stream-tube routing procedure for the transient condi-
¼ 0.4  ð9Þ
HU  U Rc tions. In this study, among the several routing procedures men-
tioned previously, two routing procedures were used to calculate
After analyzing a further set of experiments in the Missouri the dispersion coefficients from the observed concentration data.
River using the stream-tube model, Sayre (1979) rewrote Eq. (8) as
 2  2 Steady-Stream-Tube Routing Procedure
DT U W In this study, the steady-stream-tube routing procedure (S-STRP)
¼ ð0.3 ∼ 0.9Þ ð10Þ
HU  U Rc was used to calculate the transverse dispersion coefficient based
on the dosage profiles, which were converted from the two-
Recently, following Fischer (1969)’s approach, Boxall and dimensional concentration-time curves obtained in the tracer tests.
Guymer (2003) employed growth and decay terms of transverse As mentioned previously, in this method, the stream-tube model
velocity with longitudinal distance, instead of the fully developed (STM) was also used because in the STM, a fixed value of dis-
velocity, to reflect the stream-wise variations of the transverse ve- charge is attached to a fixed streamline, by moving the coordinate
locity. Baek and Seo (2008, 2011) also derived the theoretical system back and forth across the cross-section along with the flow;
equations, in which they utilized velocity profiles of the secondary thus, it can efficiently deal with the lateral migration of the water
currents given by Kikkawa et al. (1976) and Odgaard (1986), re- flow and the tracer cloud in meandering rivers, as was found in
spectively. By incorporating the stream-wise variations of the trans- Figs. 8 and 9.
verse velocity, their equations could represent the variation of the The steady-state equation can be obtained by integrating Eq. (1)
transverse dispersion coefficient along the curved channel. Baek with respect to time as (Beltaos 1975)
and Seo’s (2008) equation read as    
 2  2 ∂θ ∂θ 1 ∂ ∂θ 1 ∂ ∂θ
u þv ¼ hDL þ hDT ð13Þ
DT U H ∂x ∂y h ∂x ∂x h ∂y ∂y
¼ α fðxÞ ð11Þ
HU  U Rc
where θ = dosage of tracer, which is defined as
where α = dimensionless constant that depends on von Karman Z ∞
constant and the shear velocity; and fðxÞ = sine function of θðx; yÞ ¼ Cðx; y; tÞdt ð14Þ
the stream-wise distance. More recently, Baek and Seo (2013) 0

© ASCE 04016020-3 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


For the steady-state concentration condition, the longitudinal
dispersion can be negligible, and if the transverse velocity is also
negligibly small, then Eq. (13) can be simplified as
∂θ ∂2θ
u ¼ DT 2 ð15Þ
∂x ∂y

Applying the stream-tube concept originally suggested by


Yotsukura and Cobb (1972) into Eq. (15) and normalizing yields
∂S ∂2S
¼ BC 2 ð16Þ
∂x ∂η
where η ¼ q=Q; and S ¼ θ=Θ. The cumulative discharge, q, the
total discharge, Q, and the total dosage, Θ, are defined as
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

Z y
q¼ uhdy ð17aÞ
0 (a)
Z W
Q¼ uhdy ð17bÞ
0

Z 1
Θ¼ θdη ð17cÞ
0

and BC = bulk dispersion coefficient, given by


ψH2 U
BC ¼ DT ð18Þ
Q2
where Ψ = normalized shape factor found to have a range of
1.0–3.6 (Sayre 1979; Beltaos 1980).
The analytical solution to Eq. (16) for a vertical line source
being instantaneously injected at x ¼ 0, η ¼ ω can be obtained as
 
Θ ðη − ωÞ2
Sðx; ηÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi exp − ð19Þ (b)
4πBC x 4BC x
Fig. 2. Conceptual diagrams of (a) S-STRP; (b) T-STRP
Then, following the similar approach of the 1D routing pro-
cedure by Fischer (1968), the routing equation can be derived by
substituting the measured concentration profile for Θ in Eq. (19) as transverse dispersion coefficients using the observed concentration-
follows (Seo et al. 2006): time curves (Baek and Seo 2010). The stream-tube model equation
Z 1   for transient condition can be derived applying the stream-tube
Sðx1 ; ωÞ ðη − ωÞ2
Sðx2 ; ηÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi exp − dω ð20Þ concept to Eq. (1) without any conversion of the concentration into
0 4πBC ðx2 − x1 Þ 4BC ðx2 − x1 Þ
the dosage, and the result is
where Sðx2 ; ηÞ = predicted dosage profile as a function of the
normalized cumulative discharge, η at downstream section, x2 ; ∂C ∂2C ∂2C
¼ D L 2 þ ST 2 ð21Þ
Sðx1 ; ηÞ = measured dosage profile as a function of η at upstream ∂t ∂x ∂η
section, x1 ; and ω = normalized dummy variable of integration.
As the original routing procedure proposed by Fischer (1968), where ST ¼ UBC . Suppose that the input is comprised of a distrib-
in order to obtain DT from Eq. (20), the dosage profile at a down- uted series of separate slugs; then, the initial condition is given as
stream section is calculated using the measured profile at an up- Cðx; η; 0Þ ¼ fðx; ηÞ. Then, the analytical solution of Eq. (21) can
stream section as the input data, after an initial guess value of be derived by superposition integral as
DT is chosen. Then, as shown in Fig. 2, the calculated profile is
matched with the measured profile. This procedure is repeated until Z Z  
1 fðζ; ωÞ
∞ ðx − ζÞ2
the optimal value of DT is found. Then, this optimal value is re- Cðx; η; tÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffi exp −
garded as the observed dispersion coefficient (Fischer et al. 1979). 0 −∞ 4πt DL ST 4DL t
 2

Although the irregularities of the river can be accounted for using ðη − ωÞ
Eq. (20), the longitudinal dispersion coefficient cannot be calcu- × exp − dζdω ð22Þ
4ST t
lated because the concentration was converted into a dosage
(Seo et al. 2006).
where ζ = dummy variable of integration. The routing equation
Transient Stream-Tube Routing Procedure for the spatial concentration profile at time t2 can be derived by
In this study, the transient stream-tube routing procedure (T-STRP) substituting the measured concentration profile at time t1 for
was also adopted to calculate the both the longitudinal and fðζ; ωÞ in Eq. (22)

© ASCE 04016020-4 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


Z Z  
1 Cðζ; ω; t1 Þ
∞ ðx − ζÞ2 by the S-STRP, Eq. (20), except that when applying T-STRP,
Cðx; η; t2 Þ ¼ pffiffiffiffiffiffiffiffiffiffiffiffi exp − multiple regression techniques should be used to fit the predicted
0 −∞ 4πðt2 − t1 Þ DL ST 4DL ðt2 − t1 Þ
 2
 concentration to the measured concentration because two
ðη − ωÞ
× exp − dζdω ð23Þ parameters (longitudinal and transverse dispersion coefficients)
4ST ðt2 − t1 Þ are determined simultaneously through the single equation
[Fig. 2(b)].
where Cðx; η; t2 Þ = spatial profile of predicted concentration at t2 ;
and Cðx; ω; t1 Þ = spatial profile of measured concentration at t1 .
Employing the frozen-cloud approximation (Fischer 1968), Field Tracer Tests
Eq. (23) can be converted into the routing equation for the temporal
concentration profile as
nR h 2 i o Sites of Field Studies
1 ;ω;τ ÞU ffi ðt̄2 −t̄1 −tþτ Þ2
Z 1 −∞ ∞ pCðxffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi exp − U 4D ð t̄ − t̄ Þ dτ Field dispersion studies were conducted at five streams in Korea:
4πDL ðt̄2 −t̄1 Þ L 2 1
Cðx2 ; η; tÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Daegok Creek, Daepo Creek, Gam Creek, Han Creek, and Miho
0 4πBC ðx2 − x1 Þ
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

  Creek. The layouts of nine tests are shown in Figs. 3–5. As shown
ðη − ωÞ 2
in this figure, four streams are tributaries of the Nakdong River,
× exp − dω ð24Þ
4BC ðx2 − x1 Þ while Miho Creek flows into the Kum River. Among four streams
flowing into the Nakdong River, Daepo Creek and Daegok Creek
where Cðx2 ; η; tÞ = temporal profile of the predicted concentration are small-sized streams in terms of the stream discharge and the
at the downstream section, x2 ; Cðx1 ; ω; τ Þ = temporal profile of channel width, and tracer tests in these streams were conducted
measured concentration at the upstream section, x1 ; τ = dummy in the straight reaches. However, tests in the Gam Creek, Han Creek
time variable of integration; and t̄1 and t̄2 = mean times of passage of Nakdong River, and Miho Creek of Kum River were conducted
in sections x1 and x2 , respectively. at relatively larger scale, and the test reaches of these streams were
Eq. (24) is the combination of the 1D routing equation for the mild to sharp curved sections. The main reason for choosing test
longitudinal dispersion coefficient and the stream-tube routing reaches in these streams was that these streams, especially Gam
equation for the transverse dispersion, which is identical to the Creek, Han Creek, and Miho Creek, are very vulnerable to water
S-STRP equation [Eq. (20)]. As well documented by Baek and pollution accidents because they flow through large cities where
Seo (2010), the main merit of using this method is that both DL large heavy industry complexes are located. In recent years, acci-
and DT can be obtained simultaneously in the intermediate field dental spills of chemicals and oils caused by fire at chemical plants
of the natural river, by routing the concentration time curves at suc- or other accidents occurred several times in these streams. On top
cessive transverse points of the cumulative discharge, η, from the of that, at the downstream of the confluence of these tributary
upstream section to the downstream section. Furthermore, by streams in the Nakdong River and Kum River, many water intake
adopting the stream-tube concept, irregularities of the channel facilities are located in order to supply river water to the water treat-
can be reflected, and the pattern of the stream-wise variation of ment plants for municipal, industrial, and agricultural uses in these
the dispersion coefficients can be revealed because they are calcu- areas. Thus, the detailed 2D mixing characteristics, including
lated at each section the along river. dispersion coefficients, need to be investigated through tracer tests
The process of determining the longitudinal and transverse in order to set up the measures against water quality accidents in
dispersion coefficients by Eq. (24) is basically identical to that these rivers.

Nakdong
River

Daegok 0 3 6 km
Daepo
Creek
Nam River Creek
DP-R1

Injection Point Sec. 1 Sec. 3

Flow
Sec. 2
Sorak2 bridge
Flow

Injection
DG-R1 Point CHINA
Sec. 1
Sec. 2
Sec. 3 KOREA

Fig. 3. Map of field sites: Daegok Creek/Daepo Creek

© ASCE 04016020-5 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


GA-R4
GA-R1 Sec. 3
Sec. 4
Sec. 2
Sec. 3
0 1.5 3 km
Sec. 1 Sec. 2
Gampo Sec. 1
bridge

Injection Point
Injection Point
Flow Flow

GA-R3

Sec. 3 Nakdong
River
Sec. 2
Daedong Sec. 1
bridge Han Creek
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

Flow Injection Point

CHINA
Sec. 3
GA-R2
Sec. 2
Gam Creek
Sec. 1
HA-R1 KOREA
Gampo Injection Point
bridge Sec. 2
Flow
Injection Point Sec. 3 Sec. 1
Flow

Fig. 4. Map of field sites: Gam Creek/Han Creek

In this study, tracer tests were performed at nine sites with differ- Tracer Test Using the Rhodamine WT
ent hydraulic and geometric characteristics in order to study the
effects of those parameters on pollutant mixing in the intermediate As mentioned previously, not many tracer tests have been con-
field. The outline of test sites and conditions is summarized in ducted based on the instantaneous introduction of tracer solution
Table 1. The authors conducted the tests during the spring and fall for the analysis of two-dimensional dispersion in natural streams.
seasons of years 2011–2013, when the stream discharges were rel- Thus, in this study, in order to obtain both longitudinal and trans-
atively low because the impact of a water quality accident is usually verse dispersion coefficients, tracer tests were conducted using slug
more serious during a low flow period. inputs, following the same method established by Seo et al. (2006).

Flow

Injection Point
0 1.5 3 km
MH-R1
Sec. 1 Flow
Sec. 2 Injection Point
Sec. 4
MH-R2
Sec. 3 Sec. 1
Miho Creek
Sec. 2
Sec. 3
Sec. 4

CHINA

KOREA

Gum River

Fig. 5. Map of field site: Miho Creek

© ASCE 04016020-6 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


Table 1. Outline of Field Tracer Tests
Test case Stream Shape of test reach Discharge (cm) Test period
DG-R1 Daegok Creek Straight 0.70 June 20, 2013
DP-R1 Daepo Creek Straight 2.42 September 19, 2012
GA-R1 Gam Creek S-curved 5.74 September 19–21, 2011
GA-R2 Gam Creek S-curved 2.67 October 5–7, 2011
GA-R3 Gam Creek S-curved 8.43 November 8–10, 2011
GA-R4 Gam Creek S-curved 4.43 October 25–27, 2013
HA-R1 Han Creek Mild curved 1.79 September 12–13, 2013
MH-R1 Miho Creek Sharp curved 11.17 May 23–25, 2013
MH-R2 Miho Creek Mild curved 4.84 November 1–3, 2013

Compared to continuous injection, the instantaneous injection into the center of the stream section at the injection site shown
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

method has the merits that it does not harm the water quality in Fig. 6
and the ecosystem of the river and is also a better option in an  
economic sense because the tracer is introduced in a short time Qm L 0.93
V s ¼ 2.0 × 10−3 Cp ð25Þ
interval. U
In this study, Rhodamine WT was selected as the tracer. This
fluorescence dye has been widely used as a tracer for mixing study where V s = volume of stock Rhodamine WT 20% dye in liters;
in streams and rivers because it is mostly conservative and easily Qm = maximum stream discharge at the downstream site;
detectable with low background concentration (Pernik and Roberts L = distance to the downstream monitoring site; and Cp = peak
1985; Mukherjee et al. 2005; Rowinski et al. 2007; Sun et al. 2011). concentration at the downstream site. In order to realize the
The influence of fluorescent dyes on small aquatic organisms in vertically averaged mixing of the intermediate field, the tracer
rivers was well documented by Rowinski and Chrzanowski was injected as a full-depth vertical line source using a specially
(2011). Besides fluorescent dyes, as summarized by Mukherjee designed apparatus (Seo et al. 2006), as shown in Fig. 6. The tracer
et al. (2005), other tracers, such as salt solutions, radioisotopes, concentration measurements were undertaken at three to four
and volatile tracers of propane and sulfur hexafluoride, were used sections downstream of the injection point.
for field tests to determine flow rates and solute travel times in As shown in Fig. 6, most of the measurements were done simul-
rivers. taneously at each site, using an in situ electronic fluorometers (YSI
In this study, the 20% dye of Rhodamine WT, which was made 600-OMS sonde with 6130 Rhodamine WT sensor), whereas dur-
following the dosage formula by Kilpatrick (1970), was released ing an early stage of the field test campaign in 2011, measurements

Fig. 6. Pictures of field test (images by Hwang Jeong Choi): (a) injection of the Rhodamine WT dye; (b) installation of fluorometer sensor;
(c) measurement of velocity using ADCP; (d) measurement of velocity using ADV

© ASCE 04016020-7 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


in some cases were undertaken by manual grab samples. The dye as shown in Fig. 6. At each section, the velocity and the depth were
concentrations were detected at five points transversely and at the measured at intervals of approximately 1 m in the transverse
midstream vertically at each point. Concentrations were recorded direction.
every 30 s to obtain the concentration-time curves at each measur-
ing point. Fig. 7 illustrates the operation principle of the YSI Rhod-
amine sensor. The fluorescence dye in the flowing water tends Analysis of Field Test Data
absorb the energy of the exited light from the light source of
the sensor. Then, the emitted light of which wave length is different
Hydraulic Data
from that of the excited light is detected to measure amount of
fluorescence dye in the water sample compartment of the YSI For the analysis of the hydraulic data, the first step was to average
sonde (600-OMS). The range of sensor used in this study is 0 the velocities measured at several points vertically over the channel
to 200 μg=L, and accuracy is 5% of the reading. Resolution is depth to calculate the depth-averaged velocity. The distributions of
0.1 μg=L, and water temperature range is −5 to 50°C. Each sensor the velocity and the depth versus the lateral coordinate y were used
was calibrated using standard solutions before use in the field tests. to compute the cumulative discharge, q, and the total discharge, Q
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

The total length of the test reach for the nine tests in five streams [Eq. (17)], for the stream-tube model used in the routing procedure
ranged from 80 to 1,194 m, as shown in Figs. 3–5. As mentioned in the next section.
previously, measurements were undertaken at three to four sections, A typical example of the distributions of the primary velocity
depending upon the conditions of each test reach. The distance and the depth at the injection section and the measuring sections
from the injection point to the first measuring section was decided is shown in Fig. 8. This figure, Case GA-R2, which was conducted
using the equation given as follows (Fischer et al. 1979) in the slowly curved reach downstream of the Gampo Bridge at the
Gam Creek, shown in Fig. 4, illustrates the velocity distribution
UH2 skewed to the left at the injection section and Section 1 then skewed
LCV ¼ 0.3 ð26Þ
εV to the right at the next sections. Thus, the main body of the water
flow migrated from the left bank toward the right bank, as water
where LCV = distance of complete vertical mixing; and εV = vertical flowed downstream of the test reach. This shift of the maximum
turbulent mixing coefficient. Then, the distance from the injection velocity line (the thalweg line) would significantly affect the behav-
point to the last measuring section was decided, considering the ior of the transported cloud, as seen in the concentration-time
distance for complete transverse mixing, which is given as follows curves at each lateral point shown in Fig. 9 and also in the contour
(Fischer et al. 1979) of the tracer cloud shown in Fig. 10. Further, compared to the uni-
form flow over the infinitely wide plane of which Elder derived his
UW 2
LCT ¼ 0.1 ð27Þ theoretical equation, Eq. (2), the strong shear flow with velocity
DT variation in the transverse direction that occurred in the test reach,
as shown in Fig. 8, would induce large longitudinal dispersion.
where LCT = distance of complete transverse mixing. The second step was to average the depth-averaged velocity
At the measuring sections, lateral tag lines were installed using a over the cross section. The cross-sectional-averaged hydraulic data
rope, which was fixed on steel piles installed at both banks. These at each section are listed in Table 2. In this table, U  is the
tag lines were set up perpendicular to the principle flow direction in cross-sectional averaged friction velocity, which was derived from
order to measure the concentration of the dye cloud along the curvi- Manning’s equation for uniform flow in an open channel, given as
linear coordinate system. The interval of measuring sections was follows:
chosen as 20–100 m, as shown in Figs. 3–5, so that noticeable
pffiffiffi
change in the cloud concentration would be detected when moving n g
from one section to the next section. Using these tag lines, both U  ¼ 1=6 U ð28Þ
R
hydraulic and concentration data were obtained, either using a boat
or by walking, as shown in Fig. 6. Velocity and depth data were where R = hydraulic radius; n = Manning’s roughness coefficient;
measured using the ADCP (TELEDYNE RDI: StreamPro) where and g = gravitational acceleration. Assuming that the flow in the
the depth of the stream was deep enough, whereas the 3D ADV streams during the test periods was almost uniform flow under
(SonTek: FlowTracker) was used under shallow depth conditions, the steady-state discharge condition, the cross-sectional-averaged

Light Source
Green LED Photodetector

Optical Filter

excited light emitted light


540 nm 580 nm
YSI-6130
Rhodamine sensor wavelength
Optical Fiber

Rhodamine WT

YSI-600OMS Sonde

Fig. 7. Operation principle of Rhodamine WT sensor (adapted from YSI 2012)

© ASCE 04016020-8 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


value of the measured velocity was used to calculate the shear reach-averaged values were used in the analysis of the dispersion
velocity, U  by Eq. (28), with Manning’s n estimated based on coefficients and in the derivation of the empirical equation of the
the river bed condition and bed materials in the streams. For the longitudinal dispersion coefficient.
test reaches in this study, the reach-averaged value of W=Rc , which
was used by Yotsukura and Sayre (1976) and Sayre (1979) in their Concentration Data
empirical equation of the transverse dispersion coefficient, ranges
from 0.014 to 0.192. The concentration data obtained from the tracer tests were first
The third step of averaging was to obtain the reach-averaged smoothed by the moving average method to remove the high-
data, as listed in Table 2, by averaging the cross-sectional average frequency fluctuations, so called white noise, in the data. The
values over the whole test reach. In this study, the cross-sectional concentration-time curves at the five lateral points across the meas-
average values of the velocity and the depth were used for the uring section indicated in Fig. 8 for Case GA-R2 are plotted
application of the routing procedures to calculate the dispersion three-dimensionally in Fig. 9. In this figure, the background con-
coefficients from the measured concentration curves, whereas the centration was set to zero, assuming that the tracer used in this
study was the conservative substance. The unfiltered concentration
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

curves at each point are also shown in this figure. Comparing con-
1 centration curves before and after the filtering in this figure, the
Velocity high-frequency fluctuations usually found in the slow-moving
u (m/s)

Depth
Tracer Detector stream cases are not seen in this case.

0.5

0
h (m)

-0.5
0 5 10 15 20 25 30
(a) y (m)

1
u (m/s)

0.5

0
h (m)

-0.5
0 5 10 15 20
(b) y (m)

1
u (m/s)

0.5

0
h (m)

-0.5
0 5 10 15 20
(c) y (m)

Fig. 8. Velocity and depth distributions of case GA-R2: (a) Section 1; Fig. 9. Concentration-time curves at each section for GA-R2:
(b) Section 2; (c) Section 3 (a) Section 1; (b) Section 2; (c) Section 3

© ASCE 04016020-9 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


This figure demonstrates that all the concentration-time curves Conc. (ppb)
are not symmetrical and are skewed to the left, which means that 28
the time to centroid is bigger than the time to peak concentration.
299,600
This left-skewed curve is very natural for the C-t data obtained 24
from the measurements at the fixed points, based on the Eulerian
approach. However, long tails in the falling limb of the C-t curves 299,400 20
usually make it difficult to calculate the dispersion coefficient by

y (m)
16
the moment-based methods, by which the variance of the C-t 299,200
curves was input. 12
The most noteworthy thing in Fig. 9 is that at Section 1, C-t 299,000
curves of the higher peak concentration were found at the center 8
and near the left bank, which means that the main cloud of the tracer 298,800
was moving through the lanes close to the left bank. This is because 4
the main body of the stream water was flowing in the left side at the
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

0
first section, as shown in Fig. 8. However, as the cloud was trans- 130,200 130,400 130,600 130,800 131,000 131,200 131,400
porting toward the downstream sections, C-t curves of the higher (a) x (m)
peak concentration were found at the lanes close to the right bank of
the channel. This means that, at the downstream sections, the tracer 24
cloud was transporting leaning toward the right bank, in which
299,600
high-flow discharge was occurring. This migration of the tracer 20
cloud back and forth in the lateral direction was expected because
299,400
velocity distribution shift occurred, as shown in Fig. 8, because of 16
the channel meandering at the test reach of the Gam Creek. This is

y (m)
also true for most of the other test cases at the meandering streams. 299,200
12
For this reason, the stream-tube model, in which a fixed value of
discharge is attached to a fixed streamline by moving the coordinate 299,000 8
system back and forth within the cross-section, along with the flow,
was adopted for the calculation of the dispersion coefficient from 298,800 4
the concentration-time curves in this study. In order to visualize this
migrating behavior of the tracer cloud more vividly, the concentra- 0
tion contours of the tracer at certain time intervals are plotted in 130,200 130,400 130,600 130,800 131,000 131,200 131,400
Fig. 10. In this figure, one can clearly see the center of the cloud (b) x (m)
in which the peak concentration occurred was located in the left side
of the channel, when it was moving at the upstream section,
whereas the core was found near the right bank in the downstream 1.8
299,600
section. The other noteworthy thing in this figure is that the tracer 1.6
cloud dispersed more actively in the longitudinal direction than in 1.4
the transverse direction. As mentioned briefly earlier in this section, 299,400
1.2
this strong longitudinal dispersion was mainly caused by the con-
y (m)

tribution from the transverse variation of the stream-wise velocity, 299,200 1


and this effect on the longitudinal dispersion will be analyzed in 0.8
greater detail in later sections. 299,000
0.6

0.4
Analysis of Transverse Dispersion Coefficient 298,800
0.2
As described in the previous section, for the application of the 0
S-STRP, the dosage-dimensionless discharge curve was first ob- 130,200 130,400 130,600 130,800 131,000 131,200 131,400
tained by integrating the concentration-time distribution over time (c) x (m)
at each section. Then, the routing procedure, Eq. (20), was applied
Fig. 10. Behavior of tracer cloud for GA-R2: (a) 21 min; (b) 26 min;
to different spans of the test reach, for example, Sec. 1–Sec. 2,
(c) 36 min
Sec. 1–Sec. 3, and Sec. 2–Sec. 3. In the application of the S-STRP,
the reach-averaged values of the depth and the velocity were input-
ted into the routing equations. Among different values of the
dispersion coefficient for each test, the value that has the minimum across the channel, which are shown in Fig. 8. Thus, as the tracer
error and the longest span, over which the routing was applied, was was transporting downstream, the tracer cloud tends to skew to the
selected as the representative dispersion coefficient for the test. right side of the channel for Case GA-R2.
These representative values for each test by the S-STRP are listed The application of the T-STRP is more complicated than
in Table 3, and examples of curve fittings are depicted in Fig. 11. In S-STRP because both DL and DT should be simultaneously deter-
this figure, one can find that, in most cases, the fittings are quite mined using a single equation, Eq. (24). In this study, the Gaussian-
good, showing RMSE of less than 0.113. In some tests, dosage Newton method, one of the multiple nonlinear regression methods
profiles are not exactly symmetrical, even though the dosages (Kennedy and Gentle 1980), was used to select optimal values
are plotted against the normalized cumulative discharge, η. This for DL and DT that gave the best fit between the predicted curve
is partly because the tracer was injected at the center of the injection and the observed curve at the downstream section. Another pre-
section, without considering variations of the depth and the velocity process needed for the application of the T-STRP was that the

© ASCE 04016020-10 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


Table 2. Summary of the Cross-Sectional-Averaged Hydraulic Data
Stream Test case Section number U (m/s) U (m/s) H (m) W (m) RC (m) Sn
Daegok Creek DG-R1 1 0.29 0.035 0.26 11.0 — —
2 0.15 0.015 0.45 12.0 — —
3 0.08 0.008 0.64 13.0 — —
Average 0.17 0.019 0.45 12.0 880.3 1.03
Daepo Creek DP-R1 1 0.65 0.061 0.38 10.5 — —
2 0.71 0.068 0.40 9.0 — —
3 0.59 0.055 0.52 8.0 — —
Average 0.65 0.061 0.43 9.2 308.4 1.03
Gam Creek GA-R1 1 0.60 0.066 0.37 32.5 — —
2 0.70 0.072 0.45 20.0 — —
3 0.61 0.060 0.54 22.0 — —
Average 0.64 0.066 0.45 24.8 919.1 1.05
GA-R2 1 0.42 0.046 0.20 30.0 — —
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

2 0.53 0.050 0.28 19.0 — —


3 0.56 0.056 0.26 19.9 — —
Average 0.50 0.051 0.25 23.0 919.1 1.05
GA-R3 1 0.50 0.053 0.36 50.0 — —
2 0.60 0.065 0.45 29.0 — —
3 0.59 0.063 0.27 56.0 — —
Average 0.56 0.060 0.36 45.0 824.9 1.15
GA-R4 1 0.58 0.051 0.38 29.0 — —
2 0.52 0.055 0.27 36.0 — —
3 0.50 0.055 0.28 36.0 — —
4 0.53 0.058 0.28 33.0 — —
Average 0.53 0.055 0.30 33.5 316.7 1.13
Han Creek HA-R1 1 0.22 0.038 0.55 14.5 — —
2 0.22 0.039 0.50 17.2 — —
3 0.25 0.048 0.46 19.0 — —
Average 0.23 0.042 0.50 16.9 1133.9 1.28
Miho Creek MH-R1 1 0.37 0.040 1.09 45.0 — —
2 0.27 0.029 1.25 38.0 — —
3 0.22 0.022 1.47 38.0 — —
4 0.21 0.026 1.28 49.0 — —
Average 0.27 0.030 1.27 42.5 221.3 1.55
MH-R2 1 0.44 0.049 0.63 21.0 — —
2 0.25 0.033 0.44 45.0 — —
3 0.50 0.063 0.40 27.0 — —
Average 0.40 0.048 0.49 31.0 345.6 1.54

Table 3. Comparison of Observed Dispersion Coefficients by Two Routing Procedures


S-STRP T-STRP
 
Case Routing sections DT (m =s)
2
DT =HU RMSE (m =s) 2
DT (m =s)
2
DT =HU DL (m2 =s) DT =HU RMSE (m2 =s)
DG-R1 Section 1–3 0.0028 0.32 0.007 0.0031 0.36 0.234 27.31 2.634
DP-R1 Section 1–3 0.0141 0.53 0.012 0.0087 0.33 0.347 13.04 4.898
GA-R1 Section 1–3 0.0133 0.45 0.003 0.0229 0.77 4.460 149.52 0.459
GA-R2 Section 1–3 0.0094 0.75 0.015 0.0062 0.50 0.559 44.47 2.024
GA-R3 Section 1–2 0.0209 0.96 0.013 0.0193 0.88 1.010 46.20 5.571
GA-R4 Section 2–4 0.0071 0.43 0.013 0.0119 0.72 0.801 48.48 3.721
HA-R1 Section 1–3 0.0087 0.41 0.063 0.0172 0.82 0.497 23.62 2.817
MH-R1 Section 1–4 0.0262 0.69 0.038 0.0361 0.96 0.963 25.52 0.837
MH-R2 Section 1–3 0.0137 0.58 0.007 0.0050 0.21 0.653 27.62 2.039

concentration-time-lateral distance ðC − t − yÞ curves should be mountain by varying the two dispersion coefficients, DL and
transformed into the concentration-time-normalized cumulative DT . The results for each test are given in Table 3.
discharge ðC − t − ηÞ curves. Then, transformed curves at discrete In Table 3, the dimensionless values of the transverse dispersion
lateral points should be interpolated along η to form the three- coefficient calculated by S-STRP for the nine tests performed in
dimensional mountain made of the continuous concentration curves. this campaign range from 0.32 to 0.96, which fall in the category
This procedure was necessary because in this study, concentration of a gently meandering channel by Rutherford (1994), Eq. (6). The
curves were not obtained at the same η at all measuring sections. DT calculated by the T-STRP show similar trend to those by the
Thus, in the T-STRP, the predicted concentration mountain at the S-STRP. By this method, the dispersion coefficients for some cases
downstream section was fitted to the measured concentration of Han Creek and Miho Creek gave a discrepancy with the values

© ASCE 04016020-11 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


(a) (b)
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

(c) (d)

Fig. 11. Application of the steady stream-tube routing procedure: (a) DG-R1; (b) GA-R1; (c) GA-R2; (d) GA-R3

by the S-STRP. This difference is considered to arise from the dif- shown in Fig. 5, played a major role in increasing the rate of the
ferent level of error in the fitting by the two methods. Thus, final transverse dispersion compared with the aspect ratio.
values for DT were selected from two methods considering RMSE The relation between DT =HU  and dimensionless values of the
of each method and listed in Table 4 with dimensionless hydraulic hydraulic and geometric parameters of the river channel are plotted
and geometric parameters. in Fig. 12. Data sets available from previous field tests are also
Among the nine cases, Cases R3 and R4 of Gam Creek show a shown in this figure. As shown in this figure, there are all positive
larger dispersion coefficient than the other cases. It is considered relations between DT =HU  and W=H, U=U  , and W=Rc . Fig. 12(a)
that the higher aspect ratio, W=H, of the Gam Creek had a major shows that DT increases as W=H increases. Blanckaert (2001) and
effect on this larger transverse dispersion coefficient, as maintained Seo et al. (2006) explained that the reason for this increase is the
in the previous research (Seo et al. 2006; Baek et al. 2006). Case secondary current becomes stronger, and thus, the transverse
MH-R1 shows the largest value of DT , even though the aspect ratio dispersion coefficient becomes larger at the channel sections of
of this reach was not large. In this case, the channel curvature, as higher width-to-depth ratio. The transverse dispersion coefficient

Table 4. Summary of Tracer Test Data


Reference Case U=U W=H W=RC ðU=U Þ × ðW=RC Þ DT =HU  DL =HU DL =DT
This study DG-R1 9.07 26.71 0.0137 0.1243 0.32 27.31 85.34
DP-R1 10.7 21.12 0.0294 0.3134 0.33 13.04 39.52
GA-R1 9.67 54.96 0.0269 0.2601 0.45 149.5 332.3
GA-R2 9.85 93.00 0.0253 0.2492 0.50 44.47 88.94
GA-R3 9.33 124.5 0.0543 0.5066 0.88 46.20 52.50
GA-R4 9.73 110.8 0.1049 1.0207 0.72 48.48 67.33
HA-R1 5.53 33.58 0.0148 0.0818 0.41 23.62 57.61
MH-R1 9.05 33.40 0.1917 1.7349 0.96 25.52 26.58
MH-R2 8.21 63.29 0.0897 0.7364 0.58 27.62 47.62
Seo et al. (2006) and Beak C-Expt 1 5.48 92.71 0.1121 0.6143 0.69 31.45 45.58
and Seo (2010) S-Expt 1 6.94 78.26 0.1417 0.9836 0.52 26.87 51.67
S-Expt 2 10.4 63.73 0.1130 1.1708 0.79 22.66 28.68
S-Expt 3 6.74 117.8 0.0149 0.1003 0.36 25.35 70.42
H-Expt 1 7.45 78.13 0.1339 0.9974 0.64 87.70 137.0
H-Expt 2 3.68 63.55 0.1250 0.4607 0.23 28.40 123.5
H-Expt 3 3.77 69.07 0.1887 0.7122 0.32 9.80 30.63

© ASCE 04016020-12 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


DT/HU*

(a)
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

10 (a)

0.1
1 10 100
(b) U/U*

(b)
Fig. 13. DT and DL against the dimensionless parameters of the
rivers: (a) DT =HU versus ðU=U Þ=ðW=Rc Þ; (b) DL =HU versus
DT/HU*

ðU=U Þ=ðW=Rc Þ

DT were quite large, played a dominant role. Thus, the criteria of


the gently meandering river given by Fischer et al. (1979), Eq. (5),
cannot explain the tendency of recent data of which DT =HU  were
generally smaller than 1.
(c)

Fig. 12. Relation between DT and hydraulic and geometric parameters: Analysis of Longitudinal Dispersion Coefficient
(a) DT =HU versus W=H; (b) DT =HU versus U=U ; (c) DT =HU In Fig. 14, DL =HU  is plotted against the dimensionless parame-
versus W=Rc ters of the rivers. Data sets from previous field tests are also plotted
in this figure. As in the case of the transverse dispersion coefficient,
DL are also proportional to three-dimensionless parameters, W=H
tends to increase as the friction term, U=U  , increases, as shown in and U=U  . This increase of the longitudinal dispersion coefficient
Fig. 12(b). Fig. 12(c) indicates that DT also increases as the cur- with a larger aspect ratio and friction term was well explained
vature of the stream increases, i.e., as W=Rc increases. This is by Seo and Cheong (1998), and Baek et al. (2006). However,
considered to be because the secondary current, which becomes Fig. 14(b) reveals that DL is rather insensitive to W=Rc , contrary
stronger as the curvature of the stream increases, enhance the trans- to the behavior of DT . This tendency is also shown in Fig. 13(b) in
verse mixing (Seo et al. 2006). To analyze the effects of the channel which DL =HU  was plotted against UW=U  Rc in a log-log scale.
curvature on the rate of transverse mixing more quantitatively, In Fig. 14, most of the values of the longitudinal dispersion co-
a relation between DT =HU  and UW=U  Rc was plotted in a efficient from this study and the previous study were much higher
log-log scale in Fig. 13. In this figure, as Yotsukura and Sayre than the theoretical value suggested by Elder (1959), as mentioned
(1976) and Fischer et al. (1979) maintained, a linear relation earlier. The ratio of observed longitudinal dispersion coefficient to
may be found between DT =HU  and UW=U  Rc even though Elder’s coefficient ranges from 1.65 to 25.2. Elder’s equation was
the slope of the fitted line was 0.35, which is much smaller than derived under the assumption of uniform flow with a vertical shear
the slope of Eq. (9), of which slope was 2, suggested by Yotsukura in an infinitely wide, open channel, in which he employed the
and Sayre (1976) based on Yotsukura et al. (1970) and Sayre and logarithmic law for the vertical profile of the longitudinal flow
Yeh (1973). The discrepancy is because their result was derived in the derivation of the longitudinal dispersion coefficient. How-
using too few data so that data collected in Missouri River, of which ever, unlike the infinitely wide, open channel, the transverse shear

© ASCE 04016020-13 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


field study at Hongcheon River, was in the range of 30.6–137;
while this ratio by Sun et al. (2011), found from their tracer tests
in St. Clair River, was even larger, ranging from 176 to 200. The
DL =DT of this study, as shown in Table 4, ranges from 26.6 to 332.
Although on the other hand, the ratio of the longitudinal dispersion
coefficient of 1D ADE to the transverse dispersion coefficient,
K=DT , was reported to be much bigger than the DL =DT of this
study. Gharbi and Verrette (1998), using 48 dispersion data sets
from rivers of the United States and Canada, summarized that
K=DT ranged from 200 to 59,500. This means that K=DT is
two orders of magnitude bigger than DL =DT . Thus, the order of
magnitude of these coefficients can be suggested as
(a)
DLElder
<D L
< DKT
DT DT ð29Þ
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

101 10 2 104

where DLElder = longitudinal dispersion coefficient suggested by


Elder (1959), as given in Eq. (2).
From the foregoing discussion, it is very clear that in natural
rivers, DL of 2D ADE would be quite different from Elder’s equa-
tion. However, because there were not many theoretical and exper-
imental researches on DL of 2D ADE application, Eq. (2) was
frequently utilized as DL in the modeling of 2D ADE (Lee and
Seo 2007; Seo et al. 2008). This uncalibrated use of DL in the com-
puter models usually induced results that were substantially differ-
ent from field observations of solute concentration, particularly in
(b) systems of complex geometry and bathymetry (Piasecki and
Katopodes 1999). Thus, a more reliable predictive method for
DL of 2D ADE application is needed for accurate simulation by
the numerical model.

Conclusions
DL/HU*

In this study, the two-dimensional dispersion of dissolved contam-


inants in natural rivers was investigated through slug tracer experi-
ments. Tracer tests were conducted based on the instantaneous
injection of Rhodamine WT solution into small- to medium-sized
rivers in Korea. The effects of the hydraulic and geometric variables
on the longitudinal dispersion coefficient, DL , and transverse
(c) dispersion coefficient, DT , were analyzed. The findings of this
study are as follows.
Fig. 14. Relation between DL and hydraulic and geometric parameters: The transverse dispersion coefficient tends to increase as the
(a) DL =HU versus W=H; (b) DL =HU  versus U=U ; (c) DL =HU friction term, U=U  , increases. This was also true for the relations
versus W=Rc between DT =HU  and W=H and W=Rc . This is because the sec-
ondary current, which becomes stronger as the aspect ratio and the
sinuosity of the stream increases, enhanced the transverse mixing,
as Blanckaert (2001) and Seo et al. (2006) explained. The effect of
also exists in natural rivers as well as the vertical shear, in terms of the channel curvature on the rate of transverse mixing was
velocity variation, as shown in Fig. 8. Thus, in the tracer tests con- explained more distinctly in the plot of DT =HU  versus
ducted in natural streams of finite width, the effect of the transverse UW=U  Rc in a log-log scale, in which DT =HU  has a linear rela-
shear, in addition to the vertical shear, contributes to the longitu- tion with UW=U  Rc , even though the slope of the fitted line was
dinal dispersion. In fact, as Fischer et al. (1979) maintained, the 0.35, which is much smaller than the slope of Eq. (9), of which
contribution of the transverse variation of the primary velocity is slope was 2, suggested by Yotsukura and Sayre (1976). Because
far more significant than that of the vertical variation because both Eqs. (5) and (9) were derived using so few data that data col-
the width is usually much bigger than the depth of the channel. lected in Missouri River, of which DT were quite large, played a
In this study, the ratio DL =DT was calculated in order to inves- dominant role. Thus, it can be concluded that the criteria of the
tigate the relative order of the longitudinal dispersion to the trans- gently meandering river, of which DT =HU  is less than 1 suggested
verse dispersion and is listed in Table 4. Piasecki and Katopodes by Fischer et al. (1979), Eq. (5), cannot explain the tendency of
(1999) presented the ratio of the lumped longitudinal dispersion to recent data.
the transverse dispersion of approximately 70, based on numerical The dimensionless longitudinal dispersion coefficient,
study for field tests at the Potomac River, although the dispersion DL =HU  , is also proportional to the dimensionless parameters,
coefficients showed a wide spatial variation. The ratio, DL =DT , re- W=H and U=U  . This increase of the longitudinal dispersion
ported by Seo et al. (2006) and Baek and Seo (2010) from their coefficient with larger aspect ratio and friction term was well

© ASCE 04016020-14 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


explained by Seo and Cheong (1998) and Baek et al. (2006). How- Fischer, H. B. (1969). “The effect of bends on dispersion in streams.” Water
ever, a plot of DT =HU  versus UW=U  Rc in a log-log scale clearly Resour. Res., 5(2), 496–506.
demonstrates that DL is rather insensitive to the channel curvature. Fischer, H. B., List, E. J., Koh, R. C. Y., Imberger, J., and Brooks, N. H.
The values of the longitudinal dispersion coefficient from this (1979). Mixing in inland and coastal waters, Academic Press,
study and the previous study are much higher than the theoretical New York.
value suggested by Elder (1959). The ratio of observed longitudinal Gharbi, S., and Verrette, J. L. (1998). “Relation between longitudinal and
transversal mixing coefficients in natural streams.” J. Hydraul. Res.,
dispersion coefficient to Elder’s coefficient ranges from 1.65 to
36(1), 43–53.
25.2, which means that the observed longitudinal dispersion coef-
Holley, E. R., and Abraham, G. (1973). “Field tests of transverse mixing in
ficient is one order of magnitude larger than Elder’s coefficient. rivers.” J. Hydraul. Div., 99(23), 2313–2331.
DL =DT of this study ranges from 26.6 to 332, whereas the reported Holley, E. R., Siemons, J., and Abraham, G. (1972). “Some aspects of an-
value of the ratio of the longitudinal dispersion coefficient of 1D alyzing transverse diffusion in rivers.” J. Hydraul. Res., 10(1), 27–57.
ADE to the transverse dispersion coefficient, K=DT , was 200– Holly, F. M., and Nerat, G., Jr. (1983). “Field calibration of stream-tube
59,500 (Gharbi and Verrette 1998). This means that K=DT is dispersion model.” J. Hydraul. Eng., 10.1061/(ASCE)0733-9429(1983)
two orders of magnitude bigger than DL =DT . This clearly demon- 109:11(1455), 1455–1470.
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

strates that the longitudinal dispersion coefficient of 2D ADE is not Jackman, A. P., and Yotsukura, N. (1977). “Thermal loading of natural
the same parameter as the longitudinal dispersion coefficient of 1D streams.” U.S. Government Printing Office, Washington, DC.
ADE, K, nor as Elder’s theoretical coefficient. Jeon, T. M., Baek, K. O., and Seo, I. W. (2007). “Development of an
empirical equation for the transverse dispersion coefficient in natural
streams.” Environ. Fluid Mech., 7(4), 317–329.
Acknowledgments Kalinowska, M. B., and Rowinski, P. M. (2012). “Uncertainty in compu-
tations of the spread of warm water in a river-lessons from environmen-
This research was supported by a grant (11 Technology Innovation tal impact assessment.” Hydrol. Earth Syst. Sci. Discuss., 9(5),
C05) from the Construction Technology Innovation Program 5871–5904.
funded by Ministry of Land, Infrastructure and Transport and Kennedy, W. J., and Gentle, J. E. (1980). Statistical computing, Marcel
Dekker, New York.
by the Water Quality Control Center of National Institute of Envir-
Kikkawa, H., Ikeda, S., and Kitagawa, A. (1976). “Flow and bend
onmental Research (NIER) of the Korean government. This work
topography in curved open channels.” J. Hydraul. Div., 102(HY9),
was conducted at the Engineering Research Institute of Seoul 1327–1342.
National University in Seoul, Korea. The authors would like to ex- Kilpatrick, F. A. (1970). “Dosage requirements for slug injections of
press their gratitude to K. O. Baek for valuable comments on the rhodamine BA and WT dyes.” U.S. Government Printing Office,
manuscript. Washington, DC.
Kilpatrick, F. A., and Wilson, J. F., Jr. (1989). “Measurement of time of
travel in streams by dye tracing.” Chapter A9, USGS techniques of
References water-resources investigations, U.S. Geological Survey, Denver, 2.
Krishnappan, B. G., and Lau, Y. L. (1977). “Transverse mixing in meander-
Almquist, C. W., and Holley, E. R. (1985). “Transverse mixing in meander- ing channels with varying bottom topography.” J. Hydraul. Res., 15(4),
ing laboratory channels with rectangular and naturally varying cross- 351–370.
sections.” Technical Rep. CRWR 205, Univ. of Texas, Austin, TX. Lau, Y. L., and Krishnappan, B. G. (1981). “Modeling transverse mixing in
Baek, K. O., and Seo, I. W. (2008). “Prediction of transverse dispersion natural streams.” J. Hrdraul. Div., 107(2), 209–226.
coefficient using vertical profile of secondary flow in meandering Lee, M. E., and Seo, I. W. (2007). “Analysis of pollutant transport in the
channels.” KSCE J. Civ. Eng., 12(6), 417–426. Han river with tidal current using a 2D finite element model.” J. Hydro-
Baek, K. O., and Seo, I. W. (2010). “Routing procedures for observed Environ. Res., 1(1), 30–42.
dispersion coefficients in two-dimensional river mixing.” Adv. Water Lee, M. E., and Seo, I. W. (2013). “Spatially variable dispersion coeffi-
Res., 33(12), 1551–1559. cients in meandering channels.” J. Hydraul. Eng., 10.1061/(ASCE)
Baek, K. O., and Seo, I. W. (2011). “Transverse dispersion caused by sec- HY.1943-7900.0000669, 141–153.
ondary flow in curved channels.” J. Hydraul. Eng., 10.1061/(ASCE)HY Mukherjee, A., Fryar, A. E., and LaSage, D. M. (2005). “Using tracer tests
.1943-7900.0000428, 1126–1134.
to assess natural attenuation of contaminants along a channelized
Baek, K. O., and Seo, I. W. (2013). “Empirical equation for transverse
coastal plain stream.” Environ. Eng. Geosci., 11(4), 371–382.
dispersion coefficient based on theoretical background in river bends.”
Odgaard, A. J. (1986). “Meander-flow model. I: Development.” J. Hydraul.
Environ. Fluid Mech., 13(5), 465–477.
Eng., 10.1061/(ASCE)0733-9429(1986)112:12(1117), 1117–1135.
Baek, K. O., Seo, I. W., and Jung, S. J. (2006). “Evaluation of dispersion
Pernik, M., and Roberts, P. J. W. (1985). “Mixing in a river-floodplain sys-
coefficients in meandering channels from transient tracer tests.”
tem.” Rep. No. Scegit 85-107, Georgia Institute of Technology, Atlanta.
J. Hydraul. Eng., 10.1061/(ASCE)0733-9429(2006)132:10(1021),
1021–1032. Piasecki, M., and Katopodes, N. (1999). “Identification of stream
Beltaos, S. (1975). “Evaluation of transverse mixing coefficients from slug dispersion coefficients by adjoint sensitivity method.” J. Hydraul.
tests.” J. Hydraul. Res., 13(4), 351–360. Eng., 10.1061/(ASCE)0733-9429(1999)125:7(714), 714–724.
Beltaos, S. (1980). “Transverse mixing tests in natural streams.” J. Hydraul. Rowinski, P. M., and Chrzanowski, M. M. (2011). “Influence of selected
Div., 106(10), 1607–1625. fluorescent dyes on small aquatic organisms.” Acta Geophys., 59(1),
Blanckaert, K. (2001). “A model for flow in strongly curved channel 91–109.
bends.” Proc., 29th IAHR Congress, International Association for Rowinski, P. M., Guymer, I., Bielonko, A., Napiorkowski, J. J., Pearson, J.,
Hydraulic Research, Delft, Netherlands. and Piotrowski, A. (2007). “Large scale tracer study of mixing in a
Boxall, J. B., and Guymer, I. (2003). “Analysis and prediction of transverse natural Lowland river.” Proc., 32nd Congress of IAHR, Vol. 13, IAHR,
mixing coefficients in natural channels.” J. Hydraul. Eng., 10.1061/ 15–33.
(ASCE)0733-9429(2003)129:2(129), 129–139. Rozovskii, I. L. (1957). “Flow of water in bends of open channels.”
Elder, J. W. (1959). “The dispersion of marked fluid in turbulent shear Academy of Sciences of the Ukrainian SSR, Jerusalem.
flow.” J. Fluid Mech., 5(12), 544–560. Rutherford, J. C. (1994). River mixing, Wiley, Chichester, U.K.
Fischer, H. B. (1968). “Dispersion predictions in natural streams.” J. Sanit. Sayre, W. W. (1979). “Shore-attached thermal plumes in rivers.” Modelling
Eng. Div., 94(A5), 927–943. in rivers, H. W. Shen, ed., Wiley-Interscience, London, 15.1–15.44.

© ASCE 04016020-15 J. Hydraul. Eng.

J. Hydraul. Eng., 2016, 142(8): 04016020


Sayre, W. W., and Chang, F. M. (1968). “A laboratory investigation of Somlyódy, A. (1982). “Water-quality modelling: A comparison of
open-channel dispersion processes for dissolved, suspended, and float- transport-oriented and ecology-oriented approaches.” Ecol. Modell.,
ing dispersants.” USGS, Reston, VA. 17(3–4), 183–207.
Sayre, W. W., and Yeh, T. P. (1973). “Transverse mixing characteristics of Sun, Y., Wells, M., Bailey, S., Anderson, E. J., and Schwab, D. J. (2011).
the Missouri River downstream from the Cooper Nuclear Station.” “Study on the physical mixing patterns in the St. Clair River by dye
Technical Rep., IIHR-145, Iowa Institute of Hydraulic Research, Univ. release.” Proc., 15th Workshop on Physical Processes in Natural
of Iowa, Iowa City, IA. Waters, National Water Research Institute of Environment Canada,
Seo, I. W., Baek, K. O., and Jeon, T. M. (2006). “Analysis of transverse Univ. of Guelph, ON, Canada, 180–182.
mixing in natural streams under slug tests.” J. Hydraul. Res., 44(3), Wallis, S. G., and Manson, J. R. (2004). “Methods for predicting dispersion
350–362. coefficients in rivers.” Proc., ICE-Water Manage., 157(3), 131–141.
Seo, I. W., and Cheong, T. S. (1998). “Predicting longitudinal dispersion Yotsukura, N., and Cobb, E. D. (1972). “Transverse diffusion of solutes in
coefficient in natural streams.” J. Hydraul. Eng., 10.1061/(ASCE)0733- natural streams.” USGS, Washington, DC.
9429(1998)124:1(25), 25–32. Yotsukura, N., Fischer, H. B., and Sayre, W. W. (1970). “Measurement of
Seo, I. W., Choi, H. J., and Song, C. G. (2010). “Two-dimensional finite mixing characteristics of the Missouri river between Sioux city.” USGS,
element model for analysis of heat transport in rivers.” Proc., 6th ISEH Washington, DC.
2010, IAHR, 277–282. Yotsukura, N., and Sayre, W. W. (1976). “Transverse mixing in natural
Downloaded from ascelibrary.org by Hwang Jeong Choi on 07/19/16. Copyright ASCE. For personal use only; all rights reserved.

Seo, I. W., Lee, M. E., and Baek, K. O. (2008). “2D modeling of hetero- channels.” Water Resour. Res., 12(4), 695–704.
geneous dispersion in meandering channels.” J. Hydraul. Eng., YSI. (2012). “6-series multiparameter water quality sondes: User manual.”
10.1061/(ASCE)0733-9429(2008)134:2(196), 196–204. Yellow Springs, OH.

© ASCE 04016020-16 J. Hydraul. Eng.

View publication stats J. Hydraul. Eng., 2016, 142(8): 04016020

You might also like