Sei sulla pagina 1di 17

Environ Fluid Mech (2008) 8:117

DOI 10.1007/s10652-007-9051-6

ORIGINAL ARTICLE

Turbulent velocity profile in fully-developed open channel


flows

Hossein Bonakdari Frdrique Larrarte


Laurent Lassabatere Claude Joannis

Received: 6 April 2007 / Accepted: 26 December 2007 / Published online: 16 January 2008
Springer Science+Business Media B.V. 2008

Abstract The determination of velocity profile in turbulent narrow open channels is a


difficult task due to the significant effects of the anisotropic turbulence that involve the
Prandtls second type of secondary flow occurring in the cross section. With these currents
the maximum velocity appears below the free surface that is called dip phenomenon. The
well-known logarithmic law describes the velocity distribution in the inner region of the
turbulent boundary layer but it is not adapted to define the velocity profile in the outer region
of narrow channels. This paper relies on an analysis of the NavierStokes equations and
yields a new formulation of the vertical velocity profile in the center region of steady, fully
developed turbulent flows in open channels. This formulation is able to predict time aver-
aged primary velocity in the outer region of the turbulent boundary layer for both narrow
and wide open channels. The proposed law is based on the knowledge of the aspect ratio and
involves a parameter CAr depending on the position of the maximum velocity (dip ). dip may
be derived, either from measurements or from an empirical equation given in this paper. A
wide range of longitudinal velocity profile data for narrow open channels has been used for
validating the model. The agreement between the measured and the computed velocities is
rather good, despite the simplification used.

Keywords Dip phenomenon Free surface flow Narrow channel Secondary current
Velocity profile

Nomenclature
Ar Aspect ratio Ar = b/h
b Free surface width (m)

H. Bonakdari
Department of Civil Engineering, University of Razi, Kermanshah, Iran
H. Bonakdari F. Larrarte (B) L. Lassabatere C. Joannis
Water and Environment Division, LCPC, Route de Bouaye, BP 4129, 44341 Bouguenais Cedex, France
e-mail: frederique.larrarte@lcpc.fr

123
2 Environ Fluid Mech (2008) 8:117

F = F +f Statistical approach
F Mean component of F
f Turbulent fluctuations
g Gravitational acceleration (m s2 )
h Water depth (m)
I Turbulence intensity
k Turbulent kinetic energy
ks Roughness height (m)
Von-Karman constant
P Pressure (Pa)
Rh Hydraulic radius (m)
S0 Energy slope (m m1 )
U Velocity component in the x direction (m s1 )
Ufs Velocity component in the x direction at the free surface (m s1 )
V Velocity component in the spanwise y direction (m s1 )
W Velocity component in the vertical z direction (m s1 )
Us Secondary current velocity (m s1 )
u Shear velocity (m s1 )
z0 Roughness length of the surface (m)
ij Kroneckers symbol; ij = 1 if i = j; ij = 0 if i  = j;
dip Relative position of the maximum velocity on a vertical profile
density (kg m3 )
Kinematic viscosity (kg m3 s1 )
Dynamic viscosity (m2 s1 )
t Eddy viscosity
p Wall shear (Pa)
= hz Relative distance from the bottom
Re = uks Roughness Reynolds number
Re = UR
h
Reynolds number
ui uj Reynolds stress tensor
z+ = u z Relative distance

1 Introduction

The velocity vertical distribution in a cross section of the flow in a channel has raised up
investigations of many researchers for many years, due to its practical applications. Nikuradse
[1] has proposed a logarithmic law (log law) to describe the vertical distribution of the primary
velocity. Coles [2] has improved this law by introducing the wake function.
The development of laser technics in the 1980s has provided access to much more
accurate experimental data. So many researchers such as Sarma et al. [3], Nezu and Rodi
[4], Steffler et al. [5] have investigated the so-called dip phenomenon. The fact that a weak
secondary motion in the cross section transports low momentum fluid from the lateral solid
walls to the central section of the channel has already been highlighted by Stearns [6] at
the end of the 19th century. Due to this motion in the cross section plane of the channel,
the velocity maximum takes place below the free surface. This phenomenon occurs in nar-
row open channels with Ar = hb smaller than 5, where Ar is the ratio of the free surface
width b on the water level h. In such situations, we have a complete three dimensional
flow.

123
Environ Fluid Mech (2008) 8:117 3

Sarma et al. [3] have proposed a parabolic law to describe the vertical distribution of
longitudinal velocities in the narrow channels. More recently Yang et al. [7] and Maghrebi
and Rahimpour [8] gave insightful information to the understanding of velocity profile in
the open channel with dip phenomenon. However, the suggested models presented by these
authors require different parameters that limit their application.
In this study, a new law is proposed for the velocity profile in the outer region of either wide
or narrow channels. Firstly, the analysis of the Reynolds averaged NavierStokes (RANS)
equations is presented. Then the suggested law is validated on both experimental results
obtained from a full-scale sites and with different data sets from the literature. Considering
the hypothesis carried out for simplifying the equation, the law is valid in the outer region of
turbulent boundary layers and central region of channels.

2 Theoretical background

This paper deals with fully turbulent flows where the turbulent boundary layer has a composite
layer consisting of inner and outer regions.

2.1 Inner region

The inner region of a turbulent boundary layer is much smaller than the outer region. Its
thickness represents 1020% of the entire boundary layer thickness [9]. In this region, the
generation of turbulent kinetic energy is higher than the rate of dissipation. The behaviour is
different according to the wall rugosity i.e. smooth or rough.

2.1.1 Smooth wall

The inner region is divided in a viscous sub layer, adjacent to the wall, an intermediate area
of transition, and a fully turbulent region. In the viscous sub layer the viscosity is dominant
and it is usually assumed that the mean velocity distribution (U):
U u z
= = z+ (1)
u
where z is the distance from the bottom, u the shear velocity and the kinematic viscosity.
This linear distribution is found to be valid for z+ 5 [10]. In the fully turbulent part of the
inner region, the vertical distribution of the velocity is determined by the logarithmic law:
U 1 u z
= ln +B (2)
u
where is the Von-Karman constant and B is an integration constant. The various researchers
studied these parameters. The value of = 0.41 is a universal constant irrespective of flow
configuration [10]. In contrast, the constant B may depend on flow properties and Table 1
summarizes the values for the constant B obtained from various studies.

2.1.2 Rough wall

In the region close to the wall, velocity distribution is given by the following relationship
[15]:
U 1 z
= ln (3)
u z0

123
4 Environ Fluid Mech (2008) 8:117

Table 1 Various values


Reference B
suggested for the integration
constant (B) in the logarithmic
law Nikuradse [1] 5.50
Bradshaw et al. [12] 5.20
Stefler et al. [5] 5.50
Nezu and Rodi [4] 5.29
Cardoso et al. [13] 5.10
Kirkgz and Ardiclioglou [14] 5.50

Fig. 1 Turbulent boundary layer over the rough surface. (a) Cross section, (b) longitudinal scheme

where z0 is the roughness length of the surface. Einstein and El Samni [16] found that this
reference level, or hypothetical bed, is defined by a slip velocity and located between the top
of the roughness elements and the channel bed (Fig. 1).
This reference level is usually obtained by the adjustment of the experimental data [17].
Using the experimental results of Nikurades [1], Jan et al. [18] proposed various relations
depending on the Re parameter given as follow:

z0 0.11
= , Re 4.0 (4)
ks Re

  
z0 Re 4
= 0.0275 0.007 sin , 4.0 < Re 11.0 (5)
ks 14

  
z0 0.0125 Re 40.5
= 0.0205 + 1 + sin 11.0 < Re < 70.0 (6)
ks 2 59

123
Environ Fluid Mech (2008) 8:117 5

z0
= 0.033 Re 70.0 (7)
ks

where Re = uks is the roughness Reynolds number, ks the roughness height and the
kinetic viscosity.  
By considering Bs = 2.5 ln zko , the logarithmic law can be written:
s
 
U 1 z
= ln + Bs (8)
u ks
2.2 Outer region

In the outer region, the velocity profile can be estimated from the following equation [19]:
 
U 1 z 2 2  
= ln + Bs + w ( ) with w ( ) = sin (9)
u ks 2
where w ( ) is the wake function proposed by Coles [2] who introduced an additive correction
to the log-law.
The Eq. 9 is valid for both smooth and rough surfaces. Indeed, in the outer region, the
velocity gradient is independent of the characteristics of the walls [9]. Many researchers,
Nezu and Rodi [4], Kirkgz [20], Cardoso et al. [13] have checked the validity of this law
for the inner and outer regions of the boundary layer and they have proposed different values
for the parameter . In order to fit their experimental results with this law, Nezu and Rodi
[4] had to decrease the value proposed by Coles [2] ( = 0.55) and they proposed a value
of  = 0.20 for Reynolds numbers higher than 105 . Kirkgz [20] needed to even lower that
parameter until  = 0.10.
The Coles law is a well known function suitable for presenting the vertical velocity profile
rivers and large channels. Unfortunately, it is a monotonic increasing function of z and the
maximum velocity takes place at the water surface. In the following parts of that paper, we
focus on the case of narrow channels with an aspect ratio Ar smaller than 5. In such cases, the
velocity distribution is three dimensional even if the flow is uniform and the primary maxi-
mum velocity U is below the free surface. This phenomenon has been studied by [35,7,21].
Maghrebi and Rahimpour [8] have also investigated on this phenomenon and have proposed
vertical distribution of the velocity for such channels. The determination of the constant
parameters is difficult. To improve the analysis and to take into account these difficulties, a
new analysis was made and a new law is developed further.

3 Simplified Reynolds Averaged NavierStokes equations

Before any mathematical development, Fig. 2 presents the cross section and the coordinates
systems.
As this paper deals with turbulent flows, the RANS equations are used as the starting point
of the mathematical analysis. Thus, in agreement with that method, each flow parameter F
can be defined as:

F =F+f (10)

where F is the time-averaged quantity component and f the turbulent fluctuations of F .

123
6 Environ Fluid Mech (2008) 8:117

Fig. 2 Coordinate system and velocity distribution in a narrow open channel

The RANS momentum equation can be written as:


Ui Ui 1 P 2 Ui
+ Uj = + ui uj + gi (11)
t xj xi xk xk xj
where Ui is the mean velocity in the xi direction, P is the pressure, is the fluid viscosity
 
and ui uj are the components of the Reynolds stress tensor.
In the following part of this paper, Ui are respectively called U, V, W in the stream-
wise, lateral and vertical directions. Using the continuity equation in a steady, uniform, fully
developed turbulent flow, Eq. 11 in the streamwise (x-) direction can be written:
(UV) (UW) 2U 2U
+ = g sin + (uv) + (uw) + 2 + 2 (12)
y z 
y z y z

B 




A C D E F

where g is the acceleration due to gravity, sin the energy slope. In the Eq. 12, (A) is the
secondary flow term in the cross section, (B) the weight component term, (C) and (D) are
respectively the horizontal and vertical Reynolds stresses, (E) and (F) are the horizontal and
vertical diffusion. Yang and McCorquodale [22] have inferred that in the central region,
the vertical gradient ( z ) should be dominating and the horizontal gradient ( y ) could be
neglected. Therefore the Eq. 12 in the central region can be written as follows:
(UW) 2U
= g sin + (uw) + 2 (13)

z
 z

B


z
A D F

123
Environ Fluid Mech (2008) 8:117 7

Equation (13) can be integrated from z (z 0.2h) to h (flow depth) in the z direction
(index z indicates the value of quantity in level of z):
   
dU dU  
UW(h) UW(z) = g (h z) sin + + (uw)h (uw)z (14)
dz h dz z
At the free surface, the vertical velocity component (W), the vertical velocity fluctuation
(w) and the shear velocity dUdz are equal to zero then Eq. 14 becomes:
 
dU
UW(z) = gh ( 1) sin + (uw) z (15)
dz z
where = hz .
In wide channel, the stress (z) = ( dUdz )(z) can be fitted by a linear relation decreasing
from the bottom to the free surface [13]:
(z) = b (1 ) (16)
As most of the shear stress gradient is located close to the walls, and as we focus on the
central region where lateral walls influence is negligible, we assume to use this linear relation
in the central region of any channel [13].
By introducing u2 = 1 b that is the shear velocity, Eq. 15 becomes:
uw UW gh sin
= 2 + (1 ) (1 ) (17)
u2 u u2
gh sin
By introducing = u2 1, this can be written:
uw U W
2
+ = (1 ) (18)
u u u
The Boussinesqs hypothesis introduced the turbulent viscosity concept, then for a uniform
flow, x = 0, the Reynolds stress is given by (Rodi [23]):
dU
uw = t (19)
dz
where t is the turbulent kinematic viscosity that is not an intrinsic property of the fluid.
Yang et al. [7] have proposed:
t  z
= h 1 (20)
u h
Together with Eqs. 19 and 20 gives:
 
d U
uw u
2 = (1 ) (21)
u d
which leads to
d uU U ( ) W ( )
+ = (22)
d u u (1 )
or
dY
+ F ( ) Y ( ) = G ( ) (23)
d

123
8 Environ Fluid Mech (2008) 8:117

with
U ( )

Y ( ) = u
W ( )
F ( ) = u (1 )
(24)

G ( ) =

The general solution for such a differential equation is defined by the following relation:
      t  
Y ( ) = exp F (r) dr exp F (r) dr G (t) dt + Y(i ) (25)
i i i

 
U ( ) W ( )
= exp dr
u i u (1 )
   t  
W ( ) U(i )
exp dr dt + (26)
i i u (1 ) t u
Nezu et al. [25] have highlighted the pattern of secondary currents in open channels, the
transverse flow (V > 0) near the free surface is directed from the side wall towards the channel
center and this current flows down (W < 0) along the channel center from the free surface to
the bottom of the channel. Moreover W equals to 0 at the bottom and at the free surface. This
means that, in the central region of the channel, only the vertical mean velocity contributes
to the secondary current. This pattern has been demonstrated by 3D modeling of flow in the
narrow channel by [26]. Moreover such formulation verifies the constraints on W, i.e. null
value at both free surface and bottom and maximum in the medium of the total height of
water. Among all the formulations that verify that constraints, the following one has been
chosen:
 
W 1 2
= 1 (27)
2 + + CAr
u 2

When such an expression is considered for W(x), the expression of Y(x) can be simplified
to:

   2 
2
i2 + + CAr Ln( ) 4i + i + CAr Ln(i )
U( ) + i + CAr 4
= 2
u 2 i2
2 + + CAr + i + C Ar
2

U(i )

+ (28)
u

where is the Von Karmans constant and has its own physical meaning and can be cal-
culated from the channel characteristics and hydraulic conditions, and where i corresponds
to a certain relative position on the vertical.
To ensure both the continuity with the ln law and the fact that the secondary flow is mainly
vertical (constraint needed to ensure the validity of the differential Eq. 22, the integration is
performed between i = 0.2 and . Thus,
 
U(i ) 1 0.2h
= ln + Bs (29)
u ks

123
Environ Fluid Mech (2008) 8:117 9


   2 
2 i
i2 + + CAr Ln( ) 4 + i + CAr Ln(i )
U( ) + i + CAr 4
= 2
u 2 i2
+ + CAr
2 2 + i + CAr

 
1 0.2h
+ ln + Bs
(30)
ks

The main advantage of this expression is that it is general and can be calculated in all
cases from the physical features of the channel (rugosity, slope, etc.) and on the coefficient
CAr . When, since the physical features are known, Eq. 30 defines a set of functions depend-
ing upon parameter CAr . The maximum of Eq. 30, which corresponds to the position of the
dip phenomenon (dip ), can be completely defined from the value of CAr . Thus, a relation
between CAr and dip can be calculated from the numerical determination of dip for a large
set of values of CAr . Such a relation CAr (dip ) is proposed on the result section. Moreover,
dip can be calculated from the aspect ratio of the channel Ar, from relations proposed in the
literature. Thus, anyone can derive dip and the parameter CAr .

4 Results and discussion

4.1 Meaning of the C(Ar) parameter

Considering the Eq. 23, a new parameter CAr must be discussed. The curvature of the trans-
versal distribution of the vertical mean velocity W is governed by CAr as is smaller than 1.
Figure 3a shows that the greater values of CAr produces the flatter curves. Let us call dip the
position of the maximum velocity on a vertical velocity profile. Figure 3b gives the evolution
of the vertical velocity profile and the value of dip for a practical case with a water level of
0.77 m, a shear stress = 0.032 m/s, a relative roughness 0 = 1.6 103 and = 1.95 [27]. In
this case the relative error between the vertical profile with CAr = 9 and the velocity profile
given by the modified logarithmic law is smaller than 2%.
Considering the derivative of the longitudinal velocity at the dip phenomenon position
(dip ) is zero, an numerical relation has been found between (dip ) and CAr

CAr = D dip
E
(31)

with D = 9.3 and E = 1.7.


There is no physical relation between Ar and dip . Wang et al. [28] have fitted experimen-
tal results with an empirical relation and given the location of the dip phenomenon for the
central vertical profile in a narrow cross section by:
 
2
dip = 0.44 + 0.106Ar + 0.05 sin Ar with Ar < 5.2 (32)
2.6
Yang et al. [7] have used full of measurements in a rectangular smooth open channel and
proposed, for the central vertical profile in a narrow cross section, the following equation:
1
dip =   (33)
Ar
1 + 1.2 exp 2

123
10 Environ Fluid Mech (2008) 8:117

Fig. 3 Effect of C(Ar) on: (a) the proposed equation for W and (b) the velocity distribution and dip phenom-
enon position

Figure 4 sums up the evolution of such relations with the Ar aspect ratio. For both rela-
tions, the dip phenomenon is located at = 0.68 when Ar is equal to 2 whereas the results of
Nezu and Rodi [29] give a maximum velocity at = 0.5. Moreover, these relations deduced
by the Eqs. 23 and 24 over estimates the experimental results obtained in an ovoid shaped
sewer by [27]. In order to fit the experimental results obtained in the literature, we propose
a sigmoid model as follows, see Fig. 5:

C1 + C2 ArC3
dip = (34)
C4 + ArC3
where the constant parameters are: C1 = 42.4, C2 = 1.0, C3 = 4.2, C4 = 94.7.
Moreover, it is important to notice, that the Eq. 25 gives an asymptotical trend (for Ar > 6)
which is in agreement with the physical principles of flow in open channels. Effectively, when
the channel becomes wide, the dip phenomenon position takes place in the free surface and
thus dip tends towards 1.

4.2 Validation of the law

First of all, as the determination of the longitudinal velocity U is based on assumptions on the
vertical velocity W (Eq. 22), a comparison of the vertical velocity with experimental results

123
Environ Fluid Mech (2008) 8:117 11

Fig. 4 Position of the dip phenomenon as a function of the aspect ratio Ar. (a) Ar = 2.44, Umean = 0.67 m/s;
(b) Ar = 2.11, Umean = 0.74 m/s; (c) Ar = 2.05, Umean = 0.88 m/s; (d) Ar = 1.76, Umean = 1.0 m/s

Fig. 5 Validation of the proposed law by experimental results [27]

should have been done. Unfortunately, due the size of the results presented into the literature
it has not been possible to handle properly the experimental values. Moreover, it has not been
possible to obtain such data from contacted authors.

123
12 Environ Fluid Mech (2008) 8:117

Figure 5 compares the experimental profiles obtained by Larrarte [27] with the profile
described by Eq. 26, where the CAr parameter has been calculated with the law given by
Eq. 31. In order to keep this figure easy to read, the comparison has been made with the
Coles law as this is the most well known velocity profile law. Moreover the laws proposed
by Yang et al. [7] or Maghrebi and Rahimpour [8] require various parameters that are difficult
to calibrate for egg shaped geometries.
The relative difference with the experimental results are within 23% of the velocity
calculated with the new law at the same relative height . The law proposed here is more
accurate than the modified logarithmic law. This is particularly clear near the free surface
with a difference of 13% as a consequence to the fact that the modified logarithmic law does
not take into account of the dip phenomenon.

4.3 Application from other tests

Figure 6 compares the experimental profiles obtained in the circular sewer by Knight and
Sterling [30] with the profile described by Eq. 22. The aspect ratio is 2 and dip = 0.65. The
experimental results are within 3% of the velocity calculated with the new law at the same
relative height . The law proposed here is more accurate than the modified logarithmic law.
Near the free surface, a difference of 10% is observed.
In Fig. 7, the values obtained from the isovalues presented by Nezu et al. [25] for a rect-
angular channel with an aspect ratio of 2 and a mean velocity of 0.29 m/s are compared with
the new law. As shown above, the suggested law is able to represent experimental results
with a maximum error of 1% when the modified logarithmic law is within 22%.
All the previous comparisons are for narrow channels. As already shown in Fig. 3b, the
new law allows to reproduce the modified logarithmic law for CAr equal to 9. Both the new
law and the modified logarithmic law are able to fit the experimental results presented by
Tominaga et al. [24] for a rectangular channel with an aspect ratio equal to 7.9 and a mean
velocity equal to 0.37 m/s, as shown in Fig. 8.

Fig. 6 Validation of the proposed law by experimental data in a circular channel, Knight and Sterling

123
Environ Fluid Mech (2008) 8:117 13

Fig. 7 Validation of the proposed law by experimental data in a rectangular canal: narrow channel [25] with
Ar = 2, Umean = 0.29 m/s

Fig. 8 Validation of the proposed law by experimental data in a rectangular canal: wide channel [24] with
Ar = 7.9, Umean = 0.37 m/s

5 Conclusions

An analysis of the NavierStokes equations is presented which gives a new formulation for
the mean primary velocity in the central region of a steady, fully developed turbulent flow
in an open channel. This new law only requires the water level, free surface width, channel
slope and shear velocity.
The comparison with published data shows that the law proposed here fits very well with
the vertical distribution of the mean velocity for both narrow and wide rectangular channels.
Further development should be made to improve the continuity of the derivative in = 0.2
and also to extend the validity of the new law in the inner region. Sensibility analysis are also
requested to quantify the validity of the present law out of the central region.

123
14 Environ Fluid Mech (2008) 8:117

Acknowledgements This work has been supported by GEMCEA, a Public Interest Group for Assessing
Measurements for Water and Sewage. The authors wish to acknowledge the technical staffs of both the Water
and Environment Division and the Direction de lAssainissement de la Communaut Urbaine de Nantes
(Nantes metropolis Wastewater Authority) for their valuable contributions to these experiments.

Appendix A1

This appendix presents the way Eq. 23 is resolved:


dY (x)
+ Y (x)F(x) = G(x) (A1)
dx
We define
x

H (x) = exp F(t)dt (A2)
xo

where xo > 0

H (x0 ) = 1

x x
 
dH (x) d
= F(t)dt exp F(t)dt (A3)
dx dx
xo xo

dH (x)
= F(x)H (x) (A4)
dx
Lets multiply equation (A1) with H (x):
dY (x)
H (x) + Y (x)F(x)H (x) = G(x)H (x) (A5)
dx
This can be written:
dY (x) dH (x)
H (x) + Y (x) = G(x)H (x) (A6)
dx dx
Due to derivation laws, this is equal to:
d(H Y (x))
= G(x)H (x) (A7)
dx
After integration between x0 and x this gives:
x
Y (x)H (x) = G(t)H (t)dt + Y (xo )H (xo ) (A8)
xo

thus
!x
xo G(t)H (t)dt + Y (xo )
Y (x) = (A9)
H (x)

123
Environ Fluid Mech (2008) 8:117 15

Replacing H (x) with its expression in function to functions F(x), Y(x) can be expressed
directtly as a function to F(x) and G(x):
x t x
  
Y (x) = G(t) exp F(r )dr dt + Y (xo ) exp F(r ) dr (A10)
xo tO xO

Appendix A2

The expression XX of W (x) can lead to a great simplification of the global expression of
Y (x). Actually,
!x
G(t)H (t) dt + Y (xo )
xo
Y (x) =
H (x)

x
H (x) = exp F(t) dt
xo

x

W (t)
H (x) = exp dt
u t (1 t)
xo

x x  
W (t) t 1 t2 1
dt = u dt
u t (1 t) 1 2
2t + t + C Ar u t (1 t)
xo xo

x x
W (t) (t + 1)
dt = dt
u t (1 t) 1 2
2t + t + C Ar
xo xo

x   x
W (t) 1 2
dt = Ln t + t + C Ar
u t (1 t) 2 x0
xo

x " #
2 x + x + C Ar
1 2
W (t)
dt = Ln
u t (1 t) 2 x 0 + x 0 + C Ar
1 2
xo

" " ##
2 x + x + C Ar
1 2
H (x) = exp Ln
2 x 0 + x 0 + C Ar
1 2

2 x + x + C Ar
1 2
H (x) =
2 x 0 + x 0 + C Ar
1 2

123
16 Environ Fluid Mech (2008) 8:117

Thus
x x
2 t + t + C Ar
1 2

G(t)H (t) dt = dt
2 x 0 + x 0 + C Ar
1 2 t
xo xo

x x
1 1 2
G(t)H (t) dt = t + t + C Ar dt
1 2
2 x0 + x0 + C Ar 2 t
xo xo

x x
1 1 C Ar
G(t)H (t) dt = 1 2 t +1+ dt
2 x0 + x0 + C Ar 2 t
xo xo

x x
1 1 2
G(t)H (t) dt = 1 2 t + t + C Ar Ln (t)
2 x0 + x0 + C Ar 4 xo
xo

Thus
   
x2 x02
x02 + x + C Ar Ln (x) 4 + x0 + C Ar Ln (x0 )
2 + x 0 + C Ar
4
Y (x) =



x2 x02
2 + x + C Ar +x +C
2 0 Ar



+ Y (x0 )

References

1. Nikuradse J (1950) Laws of flow in rough pipes, Translation in National Advisory Committee for aero-
nautics, Technical Memorandum 1292. NACA, Washington, p 62
2. Coles D (1956) The low of the wake in the turbulent boundary layer. J Fluid Mech 1:191226
3. Sarma KVN, Lakshminarayana P, Lakshmana Rao NS (1983) Velocity distribution in smooth rectangular
open channel. J Hydraul Eng ASCE 109(2):270289
4. Nezu I, Rodi W (1986) Open channel flow measurements with a laser Doppler anemometer. J Hydraul
Eng ASCE 112(5):335355
5. Steffler PM, Rajaratnam N, Peterson AW (1985) LDA measurements in open channel flow. J Hydraul
Eng ASCE 111(1):119130
6. Stearns EP (1883) A reason why the maximum velocity of water flowing in open channels is below the
surface. Trans ASCE 7:331338
7. Yang SQ, Tan SK, Lim SY (2004) Velocity distribution and dip phenomenon in smooth uniform open
channel flow. J Hydraul Eng ASCE 130(12):11791186
8. Maghrebi MF, Rahimpour M (2006) Streamwise velocity distribution in irregular shaped channels having
composite bed roughness. J Flow Meas Instrum 17:237245
9. Cebeci T (2004) Analysis of turbulent flows. 2nd revised and expanded edn. Elsevier Ltd, Oxford
10. Schlichting H (1979) Boundary layer theory, 7th edn. McGraw-Hill
11. Nezu I (2005) Open channel flow turbulence and its research prospect in the 21st century. J Hydraul Eng
ASCE 131(4):229246
12. Bradshaw P, Cebeci T, Whitelaw JH (1981) Engineering calculation methods for turbulent flow. Aca-
demic Press Inc, London

123
Environ Fluid Mech (2008) 8:117 17

13. Cardoso AH, Graf WH, Gust G (1989) Uniform flow in a smooth open channel. J Hydraul Res 27(5):603
616
14. Kirkgz S, Ardiclioglu M (1997) Velocity profiles of developping and developed open channel flow.
J Hydraul Eng ASCE 115(11):10991105
15. Townsend AA (1976) The structure of turbulent shear flow, 2nd edn. Cambridge University Press
16. Einstein HA, El-Samni EA (1949) Hydrodynamic forces on a rough wall. Rev Modern Phys 31(3):520
524
17. Raupach MR, Antonia RA, Rajagopalan S (1991) Rough wall turbulent boundary layers. Appl Mech Rev
44:125
18. Jan CD, Wang JS, Chen TH (2006) Discussion of Simulation of flow and mass dispersion in meandering
channel. J Hydraul Eng ASCE 132(3):339342
19. Jimnez J (2004) Turbulent flow over rough walls. Ann Rev Fluid Mech 173196
20. Kirkgz S (1989) Turbulent velocity profiles for smooth and rough open channel flow. J Hydraul Eng
ASCE 115(11):15431561
21. Parthasarathy RN, Muste M (1994) Velocity measurements in asymmetric turbulent channel flows.
J Hydraul Eng ASCE 120(9):10001020
22. Yang SQ, McCorquodale JA (2004) Determination of boundary shear stress and Reynolds shear stress in
smooth rectangular channel flows. J Hydraul Eng ASCE 130(5):458462
23. Rodi W (1993) Turbulence models and their application in hydraulics, a state of the art review. 3rd edn.
A. A. Balkema, Rotterdam
24. Tominaga A, Nezu I, Ezaki K, Nakagawa H (1989) Three dimensional turbulent structure in straight open
channel flows. J Hydraul Res 27(1):149173
25. Nezu I, Nakagawa H, Rodi W (1989) Significant difference between secondary currents in closed channels
and narrow open channels. In: Proc 23rd IAHR congress, vol A, Delft, pp 125132
26. Bonakdari H (2006) Modlisation des coulements en collecteur dassainissementapplication la con-
ception de points de mesures, Ph. D thesis, University of CaenBasse Normandie, France
27. Larrarte F (2006) Velocity fields in sewers: an experimental study. Flow Meas Instrum 17:282290
28. Wang X, Wang ZY, Yu M, Li D (2001) Velocity profile of sediment suspensions and comparison of log
law and wake law. J Hydraul Res 39(2):211217
29. Nezu I, Rodi W (1985) Experimental study on secondary currents in open channel flow. In: Proc 21st
congress of IAHR, Melbourne, vol 2, pp 115119
30. Knight DW, Sterling M (2000) Boundary shear in circular pipes running partially full. J Hydraul Eng
ASCE 126(4):263275

123

Potrebbero piacerti anche