Lecture Notes in Turbulence
Lecture Notes in Turbulence
B. MUTLU SUMER
Technical University of Denmark
DTU Mekanik, Section for Fluid Mechanics, Coastal & Maritime Engineering
Building 403, 2800 Lyngby, Denmark, bms@mek.dtu.dk
Revised 2013
Contents
1 Basic equations
1.1 Method of averaging and Reynolds decomposition
1.2 Continuity equation . . . . . . . . . . . . . . . . .
1.3 Equations of motion . . . . . . . . . . . . . . . .
1.4 Energy equation . . . . . . . . . . . . . . . . . . .
1.4.1 Energy equation for laminar ows . . . . .
1.4.2 Energy equation for the mean ow . . . .
1.4.3 Energy equation for the uctuating ow .
1.5 References . . . . . . . . . . . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
7
7
8
9
11
11
12
13
15
17
. 17
. 17
. 23
.
.
.
.
.
.
.
.
.
31
43
48
49
57
59
61
66
69
3 Statistical analysis
73
3.1 Probability density function . . . . . . . . . . . . . . . . . . . 73
3
CONTENTS
3.2 Correlation analysis . . . . . . . . . . . . . . . . . . . . . . .
3.2.1 Space correlations . . . . . . . . . . . . . . . . . . . .
3.2.2 Time correlations . . . . . . . . . . . . . . . . . . . .
3.3 Spectrum analysis . . . . . . . . . . . . . . . . . . . . . . . .
3.3.1 General considerations . . . . . . . . . . . . . . . . .
3.3.2 Energy balance in wave number space . . . . . . . . .
3.3.3 Kolmogoro s theory. Universal equilibrium range and
inertial subrange . . . . . . . . . . . . . . . . . . . .
3.3.4 One-dimensional spectrum . . . . . . . . . . . . . . .
3.4 References . . . . . . . . . . . . . . . . . . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
. . . . . . .
layer depth
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
.
.
.
.
.
.
76
76
81
84
85
87
. 91
. 94
. 98
.
.
.
.
.
.
.
.
.
.
.
.
.
99
99
103
104
113
113
114
118
119
121
123
124
126
129
131
. 132
. 132
. 135
. 136
. 140
. 140
. 141
. 144
. 144
. 144
. 146
CONTENTS
. .
. .
. .
F1
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
155
. 155
. 156
. 157
. 157
. 160
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
169
169
171
171
F2
. .
. .
. .
. .
. .
. .
. .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
172
172
174
175
181
181
183
189
function
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
CONTENTS
Chapter 1
Basic equations
1.1
1
T
component
t0 +T
ui dt
(1.1)
t0
in which t0 is any arbitrary time, and T is the time over which the mean is
taken. Clearly T should be su ciently large to give a reliable mean value.
This is called time averaging. There are also other kinds of averaging, such
as ensemble averaging, or moving averaging. We shall adopt the ensemble
(1.2)
= f +g
= af
= a
f
=
s
= fg
(1.3)
f = f
f = 0
fg = fg
fh
= f
h =0
1.2
Continuity equation
(1.5)
(1.6)
(1.7)
To sum up, Eq. 1.6 is the continuity equation for the mean velocity and
Eq. 1.7 is the continuity equation for its uctuating part.
1.3
Equations of motion
ui
ui
+ uj
) = gi +
t
xj
ij
(1.8)
xj
in which t is the time, gi is the volume force (such as the gravity force) and
xi is the Cartesian coordinates. The quantity ij is the stress on the uid,
and given as
ui
uj
p ij + (
+
)
(1.9)
ij =
xj
xi
The preceding relation is the constitutive equation for a Newtonian uid.
Here, p is the pressure, and is the viscosity, and ij is the Kronecker delta
dened as
ij
ij
= 1 for i = j,
= 0 for i = j
(1.10)
From the continuity equation (Eq. 1.4), the N.-S. equation becomes
ui
+
( ui uj ) = gi +
t
xj
ij
xj
(1.11)
ij
xj
(1.12)
10
Now, the second term on the left hand side of the preceding equation (using
the Reynolds decomposition)
xj
( ui uj ) = uj
ui
+
( ui uj )
xj
xj
(1.13)
ui
ui
+ uj
) = gi +
(
t
xj
xj
ij
ui uj )
(1.14)
ui uj =
u1 u1
u2 u1
u3 u1
u1 u2
u2 u2
u3 u2
u1 u3
u2 u3
u3 u3
(1.15)
As seen, these stresses form a symmetrical second order tensor (the Reynolds
stress tensor).
To summarize at this point, there are four equations for the mean ow.
These are
1. the continuity equation (Eq. 1.6) and
2. three equations of motion (Eq. 1.14),
and there are ten unknowns (namely, three components of the velocity,
ui , and the pressure, p, and six components of the Reynolds stress,
ui uj ).
Therefore the system is not closed. This problem is known as the closure
problem of turbulence. We shall return to this problem later.
(In the case of laminar ow, there are four unknowns (namely, three components of the velocity, ui , and the pressure, p) and four equations, namely
the continuity equation (Eq. 1.4) and three N.-S. equation (Eqs. 1.8 and
1.9). Therefore, the system is closed.)
1.4
11
Energy equation
1.4.1
N.-S. equation:
ui
+u
t
ui
= gi
x
ui
x x
1 p
+
xi
(1.16)
ui
+ uj u
t
ui
= uj gi
x
uj
2
1 p
ui
+ uj
xi
x x
(1.17)
uj
and solving
gives
uj
uj
t
uj
t
ui
(1.18)
ui
= (ui uj )
t
t
1 p
+
xj
ui (gj
uj
x x
uj
)
x
(1.19)
2
ui
ui
= ui gi + ui
x
x x
(1.20)
ui
u
xi x
on the right hand side of the equation (this term is actually zero due to
continuity), and after some algebra, the following equation is obtained
K
+
[Ku + pu
t
x
ui (
ui
u
+
)] = u g
x
xi
(1.21)
12
1
1
ui ui = (u2 + v 2 + w2 )
2
2
is the energy dissipation per unit volume of uid and per unit time
= (
1.4.2
(1.22)
ui
u
ui
+
)
x
xi x
(1.23)
The equation of motion for the mean ow (the Reynolds equation, Eq.1.14):
ui
+ u
t
ui
= gi
x
p
+
xj
ui
+
(
x x
x
(1.24)
ui u )
Now, in the preceding subsection, the energy equation, Eq. 1.21, is obtained from the equation of motion, Eq. 1.16. In a similar manner, the energy
equation for the mean ow can be obtained from the equation of motion for
the mean ow (i.e., from Eq. 1.24). The result is
K0
+
[K0 u + ui u ui +p u
t
x
ui (
ui
u
+
)] = u g
x
xi
ui u )
(1.25)
in which K0 is the kinetic energy per unit volume of uid for the mean ow
K0 =
1
1
ui ui = (u2 + v2 + w2 )
2
2
(1.26)
and 0 is the energy dissipation per unit volume of uid and per unit time,
again, for the mean ow
0
= (
ui
u
ui
+
)
x
xi x
(1.27)
Now, note the last term on the right hand side of Eq. 1.25. We shall return
to this term in the next subsection.
ui
x
1.4.3
13
ui (
ui
u
+
)] = u g
x
xi
(1.28)
= u g
ui (
ui u )
ui
u
+
)] =
x
xi
ui
x
(1.29)
ui (
ui
u
+
)] = u g
x
xi
t +(
ui u )
(1.30)
in which Kt is the kinetic energy per unit volume of uid for the uctuating
ow (i.e., the kinetic energy of turbulence)
Kt =
1
1
ui ui = (u 2 + v 2 + w 2 )
2
2
(1.31)
and t is the viscous dissipation of the turbulent energy per unit volume of
uid and per unit time
t
= (
ui
u
u
+
) i
x
xi x
(1.32)
Eq. 1.25 is the energy equation for the mean ow, and Eq.1.30 is that
for turbulence. As seen, the term ( ui u ) xui is common in both equations,
but with opposite signs. This implies that if this term is an energy gain for
turbulence, then it will be an energy loss for the mean ow, or vice versa.
Now, we shall show that this term is an energy gain for turbulence.
Integrating Eq.1.30 over a volume V gives
t
Kt dV +
V
F n dS =
S
u g dV
V
t
V
dV +
(
V
ui u )
ui
dV
x
(1.33)
ui
x
14
in which
F = Kt u +
1
u uu +pu
2 i i
ui (
ui
u
+
)
x
xi
(1.34)
and S is the surface encircling the volume V. Here, the volume integral in
the second term on the left hand side of Eq. 1.33 is converted to a surface
integral, using the Green-Gauss theorem. The rst term on the right hand
side of Eq. 1.33 drops when the volume force is taken as the gravity force,
since g = 0. The second term on the left hand side of Eq. 1.33, on the other
hand, is the energy ux at the surface S. Therefore, it is not a source term.
Now, consider a situation where there is no inux of turbulent energy. In
this case, the maintenance of turbulence (i.e., t V Kt dV
0 in Eq. 1.33) is
ui
only possible when V ( ui u ) x dV in Eq. 1.33 is positive, meaning that
the term ( ui u ) xui in the turbulence energy equation (Eq. 1.30) should
be positive. In other words, this term is an energy gain for turbulence, and
therefore it is an energy loss for the mean ow. This is an important result,
because it implies that turbulence extracts its energy from the mean ow, and
the rate at which the energy is extracted from the mean ow is ( ui u ) xui .
An important implication of the above result is that turbulence is generated only when there exists a velocity gradient in the ow xui . When
there is no velocity gradient, no turbulence will be generated. In this case,
any eld of turbulence introduced into the ow will be dissipated (the decay of turbulence). See Monin and Yaglom (1973, pp. 373-388) for further
discussion.
Example 1 Find the turbulence energy generation and the total energy dissipation (per unit volume of uid and per unit time) for a turbulent boundary
ow in an open channel (Fig. 1.2).
The energy extracted from the mean ow is (
ow, this will be
u
( uv)
y
ui u )
ui
x
This is the energy extracted from the mean ow per unit time and per unit
volume of uid (Fig. 1.2 c). As seen from the gure, the energy generation
for turbulence is zero at the free surface while it increases tremendously as
the wall is approached.
1.5. REFERENCES
15
Figure 1.2: (a) Shear stress, (b) velocity and (c) turbulent energy extracted from
the mean ow.
From Eq. 1.25, the total energy dissipation, on the other hand, is
0
in which
be
+(
ui u )
ui
x
u u
u
)
+( uv)
y y
y
From the preceding relation, it is seen that the energy dissipation is very
large near the wall, and it decreases with the distance from the wall, similar
to the variation of the turbulent energy generation (Fig. 1.2 c).
(
1.5
References
16
Chapter 2
Steady boundary layers
In this chapter, we will study steady, turbulent boundary layers over a wall,
in a channel, or in a pipe. (Examples of turbulent boundary layers over a
wall may be: ow over a at plate, ow over the seabed, the atmospheric
boundary-layer ow over the earth surface, etc.).
We will rst concentrate on the ow close to a wall (a general analysis),
and then we will study the ow close to a smooth wall, and subsequently
that close to a rough wall. (The latter will include the ow near the bed
of a channel, and that near the wall of a pipe, etc.). Finally, we will turn
our attention to ows in channels/pipes, considering the entire channel/pipe
cross section.
2.1
2.1.1
18
u
u
u
+ u + v ) = gx
t
x
y
2
2
p
u
u
+ ( 2 + 2)+ (
x
x
y
x
u 2) +
uv)
(2.1)
Since ut = 0, v = 0, g x = 0, xp = 0, u is a function of only y, and the ow is
uniform in the x direction (i.e., ux = 0), then the Reynolds equation will
be
d2 u
d
( uv)=0
+
(2.2)
dy 2 dy
or,
d du
d
(
+ ( u v )) =
=0
(2.3)
dy dy
dy
in which is the total shear stress:
=
du
+(
dy
uv)
(2.4)
which is composed of two parts, namely the viscous part (the rst term), and
the part induced by turbulence (the Reynolds stress, the second term).
Integrating 2.3 gives
= c;
c = constant
(2.5)
(2.6)
19
is the wall shear stress. From Eqs. 2.4 and 2.6, one obtains
(=
du
+(
dy
u v )) =
(2.7)
0,
y,
, )=0
u yUf
,
)=0
Uf
(2.10)
can be
(2.11)
(2.12)
20
Solving
u
Uf
(2.13)
u
= f (y + )
Uf
(2.14)
in which
y+ =
yUf
(2.15)
Eq. 2.14 is known as the law of the wall. The quantities Uf and
the inner ow parameters, or wall parameters.
are called
u v )) =
d
=
dy
gx =
gS =
(2.16)
Sy + c;
c: constant
(2.17)
21
=0
(2.18)
= Sh(1
(2.19)
The wall shear stress (the bed shear stress) will then be
0
(2.20)
= Sh
y
)
h
0 (1
(2.21)
0 (1
y
)
h
(2.22)
i.e., near the wall, the shear stress can be assumed to be constant and equal
to the wall shear stress,
0 . This thin layer of uid is called the constant
stress layer.
As seen, the ow in the constant stress layer is similar to the idealized
ow considered in the preceding section; in both ows, the shear stress is
constant across the depth. The immediate implication of this result is that,
the velocity distribution close to the wall in the present case (i.e., in the
constant stress layer) is given by the law of the wall (Eq. 2.14):
u
= f (y + )
Uf
(2.23)
Example 3 Boundary layer in a pipe. Flow close to the wall of the pipe.
Now, consider a fully developed turbulent boundary layer in a pipe (Fig.
2.3).
Similar to the previous example, the x component of the Reynolds equation in the present case:
1
p
(r ) =
(2.24)
r r
x
22
p r2
+c
x2
(2.25)
in which c is a constant.
The boundary condition at the center line of the pipe
r=0:
(2.26)
=0
=2
r
or
D
(2.27)
0 (1
2y
)
D
(2.28)
(2.29)
2.1.2
23
du
+(
dy
uv)
du
dy
(2.30)
24
or
0
du
dy
(2.31)
y+c
(2.32)
Integrating gives
u=
Uf2
u=0
y=0:
Uf2
(2.34)
(2.35)
The velocity varies linearly with the distance from the wall. This analytical
expression agrees well with the experiments for y + < 5. Beyond y + = 5, the
data begins to deviate from the analytical expression (Fig. 2.5). The layer in
which the analytical expression agrees with the experimental data is called
the viscous sublayer:
v
=5
Uf
, or alternatively
+
v
=5
(2.36)
It must be mentioned that, although this very thin layer of uid is called
viscous sublayer, turbulence from the main body of the ow does penetrate
into this layer. This can be seen in the velocity signal in the viscous sublayer,
and also in the wall shear stress signal in the form of turbulent uctuations.
To get a feel of how thin this layer is, for example, for Uf = O(1 cm/s),
2
O(1) = O(0.05 cm) = O(0.5 mm).
v = 5 Uf = 5 10
Mean velocity for large values of y + . Logarithmic layer.
Fig. 2.4 shows that, for large values of y + , the viscous part of the total
shear stress is small compared with the turbulent part, du
( u v ). (It
dy
25
should be noted that, although y + is large, yet y is still small compared with
h so that
0 , namely y is in the constant stress layer).
Now, the shear stress generates the velocity variation over the depth; i.e.,
the cause-and-e ect relationship between the shear stress and the velocity
variation du
is such that generates du
.
dy
dy
Since the shear stress for large values of y + is practically uninuenced
by the viscosity, the velocity variation du
should also be uninuenced by the
dy
viscosity; therefore, dropping the viscosity , from Eq. 2.9,
du
d
= F ( 0 , y, )
dy
dy
(2.37)
or
du
, 0 , y, ) = 0
(2.38)
dy
From dimensional analysis, this dimensional equation is converted to the
following nondimensional equation:
(
du 1
)=0
dy Uf
(2.39)
(2.40)
(2.41)
(2.42)
26
Figure 2.5: Velocity distribution. Taken from Monin and Yaglom (1973).
1. Experiments done in pipe ows, in channel ows, and in boundary layer
ows over walls, etc. have all conrmed the logarithmic law with
A = 2.5,
B = 5.1
(2.43)
27
0,
du
,
dy
70
u2
vu
wu
uv
v2
wv
uw
vw
w2
(2.44)
28
uv
(2.45)
Now, close to the wall, these quantities should depend on the same independent quantities as in the case of the mean velocity (Eq. 2.9), namely
0,
y,
(2.46)
v 2 = Uf f2 (y + ),
w 2 = Uf f3 (y + ),
u v = Uf2 f4 (y + )
(2.47)
The functions f1 , ..f4 are universal functions, and their explicit forms are to
be determined from experiments.
For large values of y + , we know that the variation of the velocity with y,
i.e., du
, is independent of the viscosity (see the argument in conjunction with
dy
, and
Eq. 2.37). Since the turbulence is generated by the velocity gradient du
dy
du
since dy is independent of the viscosity, then the turbulence should also be
independent of the viscosity. Therefore, for large values of y + , the viscosity
29
Figure 2.7: Turbulence quantities. Taken from Monin and Yaglom (1973).
should drop out in the expressions in Eq. 2.47, meaning that the turbulence
quantities in Eq. 2.47 should tend to constant values:
u2
= f1 (y + )
Uf
A1 ,
v2
= f2 (y + )
Uf
A2 ,
w2
= f3 (y + )
Uf
uv
= f4 (y + )
Uf2
(2.48)
A3 ,
A4
Fig. 2.7 shows the experimental results. As expected, the universal func-
30
tions f1 , ..f4 approach zero, as y + goes to zero, and also, in conformity with
Eqs. 2.48, they tend to constant values as y + takes very large values. From
the gure, the latter constant values are
2.3, A2
A1
0.9, A3
1.7, and A4
(2.49)
1
u uu +pu
2 i i
ui (
ui
u
+
)
x
xi
(2.51)
The second term in Eq. 2.50 represents the turbulence production, and
the third term the viscous dissipation of turbulent energy.
The rst term, on the other hand, represents the ux of turbulent energy;
it consists of four parts:
1. Kt u represents the transport of turbulent energy by convection;
2.
1
2
ui (
ui
x
u
xi
For the present boundary-layer ow, the turbulent energy budget, Eq.
2.50, reduces to
u
F
+ t=0
( uv)
(2.52)
x
y
It can easily be seen that F in the present case consists of only items 2, 3
and 4 above.
Fig. 2.8 shows the contributions to the turbulent energy budget of the
previously mentioned e ects plotted versus the distance from the wall y + . It
is seen that most of the energy production and energy dissipation takes place
near the wall in the bu er layer, namely 5 < y + < 70.
31
2.1.3
Mean velocity.
Consider that the wall is now covered with roughness elements (Fig. 2.9).
Let the height of the roughness elements be k. Assume that k is su ciently
large (larger than the thickness of the viscous sublayer) so that it inuences
the ow.
In this case, there is one additional parameter to describe the ow, namely
the roughness height, k. Therefore, Eq. 2.9 will read
u = F ( 0 , y,
, , k)
(2.53)
Now, the total shear stress is given as in the case of the smooth wall (Eq.
32
du
+(
dy
(2.54)
uv)
du
,
dy
u v , will in the
2. the other is caused by the vortex shedding from the roughness elements.
(Observations show that lee-wake vortices are shed from the roughness
elements into the main body of the ow in a continuous manner, Sumer
et al., 2001).
in
Clearly, for large distances from the wall, (1) the viscous part du
dy
the total shear stress (Eq. 2.54) is negligible, as discussed in the previous
section, and (2) the vortex-shedding-induced part of
u v is also negligible,
because the vortices shed into the ow cannot penetrate to large distances
due to their limited life time. This means that, for large distances from the
wall, the shear stress is practically independent of the viscosity and the wall
roughness.
Now, as discussed in the previous section, in the cause-and-e ect relationship between the shear stress and the velocity variation, the shear stress
is the cause and the velocity variation du
is the e ect. Since the shear stress
dy
for large distances from the wall is independent of the viscosity and the wall
33
(2.55)
This is precisely the same equation as in the case of the smooth wall (Eq.
2.37). Therefore, the result obtained in conjunction with Eq. 2.37 is directly
applicable (Eq. 2.41), i.e.,
u = AUf ln y + B1
(2.56)
Here, A is the same constant as that in the case of the smooth wall, A =
2.5. Our analysis implies that this constant has the same value, irrespective
of the category of the wall (smooth, or rough). Therefore, A must be a
universal constant.
In the present context, the preceding expression may be written as
u = AUf ln(
y
)
y0
(2.57)
in which y0 is called the roughness length (not to be confused with the roughness height k). Eq. 2.53 suggests that this new quantity should be a function
of , k, 0 and ,
y0 = y0 ( , k, 0 , )
(2.58)
and from dimensional analysis,
y0 = k g(
kUf
(2.59)
kU
ks Uf
(2.60)
34
2. Completely rough wall when s f > 70. (In this case, the viscous sublayer is completely destructed, and the roughness elements are completely exposed to the main body of the ow); and
3. Transitional wall when 5 <
ks Uf
< 70.
Now, in the case of the hydraulically smooth wall, the function g takes
the following form
ks Uf
a
g(
) = ks Uf
(2.61)
This is because ks will drop out only when g is given in the form depicted
in the preceding equation. Here a is a constant. Inserting the preceding
expression into Eq. 2.60
a
y0 = ks ks Uf
and from Eq. 2.57
u = AUf ln(
y
y
1 yUf
)
) = AUf ln( a ) = AUf ln(
y0
a
Uf
Comparing the preceding equation with Eqs. 2.42 and 2.43 gives the constant
a as 1/9.
In the case of the completely rough wall, on the other hand, the function g is given as
ks Uf
)=b
(2.62)
g(
This is because the viscosity does not play any role, and therefore should
drop out of formulation; and this is possible only when g is given as in the
preceding form. Here, b is another constant. Nikuradses experiments show
that this constant is
ks Uf
1
g(
)=b=
(2.63)
30
35
ks
30
(2.64)
30y
)
ks
(2.65)
This is the logarithmic law for ow over a completely rough wall. Fig. 2.10
compares the above expression with the experimental data (Grass, 1971, Fig.
4). As seen, the logarithmic law begins to deviate from the measured velocity
prole for y 0.2ks . This is not unexpected, because it has been seen that,
for the velocity prole to satisfy the logarithmic law, y should be su ciently
large, so large that the variation of the velocity du
is independent of the
dy
wall roughness; otherwise, the velocity distribution will clearly not satisfy
the logarithmic law, as revealed by Fig. 2.10.
In the case of the transitional wall, the function g is given in Fig. 2.11.
The data can be represented by the following empirical expression:
g(ks+ ) =
1/9
1
+
exp[ 140(ks+ + 6)
+
ks
30
1.7
36
37
sometimes reach values as much as 5k and even more, depending on the shape
of the roughness elements, the packing pattern as well as the roughnessheight-to-ow-depth ratio (the latter is for large values of this ratio such as
> O(0.3)) (Bayazit, 1983). In particular, ks =2k for walls covered by sand,
pebbles, or stones the size in the range 0.54-46 mm (Kamphuis, 1974); and
ks =(3 6)k for the ground covered with ordinary grass or agricultural crops
(Monin and Yaglom, 1973). Bayazit (1983) reports the following values for
the equivalent sand roughness for various boundaries:
Author
Roughness
ks
Leopold et al. (1964)
Limerinos (1970)
Bayazit (1976)
Charlton et al. (1978)
Hey (1979)
Thompson and Campbell (1979)
Gladki (1979)
Denker (1980)
Bray (1980)
Gri ths (1981)
Gravel
Gravel
Closely packed spheres
Gravel
Gravel
Gravel
Gravel
Closely-packed cylinders
Gravel
Gravel
3.5D84
3.5D84
2.5D
3.5D90
3.5D84
4.5D90
2.5D80
2D
3.5D84 (3.1D90 )
5D50
38
covered with ripples, ks is found to be ks =(2 3)k with 2-D ripples (Fredse et
al., 2000), k being the ripple height. In the case of 3-D ripples, the previous
relation can, to a rst approximation, be used. For very high velocities,
ripples are washed away and the bed becomes plane again; This sedimenttransport regime is called the sheet-ow regime (Sumer et al., 1996). In this
case, the ks value can be calculated from the following empirical relations
(Sumer et al., 1996):
1. No-suspension sediment transport (w/Uf > 0.8
ks
= 2 + 0.6
k
2.5
0<
1):
2
(2.66)
1):
2
2.5
(2.67)
Uf2
g(s
1)k
(2.68)
in which s is the specic gravity of sediment grains and g the acceleration due to gravity.
Theoretical wall.
In the case of the rough wall, the exact location of the wall is not well
dened. Does it lie at the top of the roughness elements, or does it lie at
the base bottom, or does it lie somewhere between? This issue brings in the
concept of the theoretical wall.
Formally, the theoretical wall is dened as the location from which the y
distances in Eq. 2.65 are measured (Fig. 2.12).
Now, for convenience, change the coordinate to y , the distance from the
base wall. Now suppose that the theoretical bed lies at the distance y1 from
the wall (Fig. 2.12), then the relationship between y and y :
y=y
y1
(2.69)
Figure 2.13:
39
40
y1 )
30(y
ks
(2.70)
For a given measured velocity prole u(y ), and taking A = 2.5, the
quantities Uf , ks and y1 (i.e., the location of the theoretical wall) can be
determined from Eq. 2.70. The procedure may possibly be best described
by reference to the following example:
In the laboratory, time-averaged velocity proles are measured with a
Laser Doppler Anemometer over a stone-covered bed in an open channel
(the stones the size of k = 3.8 cm, and the ow depth h = 40 cm from
the top of the stones). The measurements are made at four, equally spaced,
vertical sections between the crests of two neigbouring stones. Then, the
space-averaged velocity prole is obtained from the measured, four velocity distributions. Subsequently, the following procedure is adopted to get
the location of the theoretical wall (namely, y1 , or alternatively y, see the
denition sketch in Fig. 2.13), Uf and ks .
1. Plot u (here, u denotes the time- and space-averaged velocity) in a semilog graph for various values of y1 (or alternatively for various values of
y). (The measured velocity proles are plotted for eight values of
y, namely y/k = 0; 0.1; 0.15; 0.2; 0.25; 0.3; 0.35; 0.4. Only three
proles are displayed here, in Fig. 2.13, to keep the gure relatively
simple).
2. Identify the straight line portion of each curve in Fig. 2.13.
3. For this, look at the interval 0.2ks y 0.1h to start with. (The latter
interval can be taken as 0.2ks y (0.2 0.3)h with the upper bound
relaxed if necessary). This is the interval where the logarithmic layer is
supposed to lie. Here the upper boundary 0.1h (or (0.2 0.3)h in the
relaxed case), ensures that the y levels lie in the constant stress layer,
while the lower boundary, 0.2ks , ensures that the variation of u with
respect to y is not inuenced by the boundary roughness, two conditions
necessary for the velocity distribution to satisfy the logarithmic law.
4. Identify the case where the thickness of the logarithmic layer (where the
velocity is represented with a straight line) is largest. The previously
41
0
0.1
0.15
0.2
0.25
0.3
0.35
0.4
Logarithmic layer
Thickness of the
logarithmic layer
3.5
3.5
3.5
4.5
3.5
3.0
2.5
2.5
cm
cm
cm
cm
cm
cm
cm
cm
5. As seen from the preceding table, the case where y/k = 0.2 gives
the thickest logarithmic layer. Therefore, adopt this location as the
location of the theoretical wall (Fig. 2.13 b).
6. The straight line portion of this velocity prole (corresponding to y/k =
0.2, Fig. 2.13 b) has a slope equal to AUf , Eq. 2.70. From this information, nd Uf .
7. Extend the straight line portion of the velocity prole (corresponding
ks
to y/k = 0.2) to nd its y-intercept; this is equal to 30
(Fig. 2.13 b).
Returning to the theoretical wall, research shows that the origin of the
distance y(= y y1 ), namely the level of the theoretical bed, lies (0.15-0.35)k
below the top of the roughness elements (Bayazit, 1976, and 1983).
Turbulence quantities.
The turbulence quantities given in Eq. 2.45 in the case of the rough
wall should depend on the same independent quantities as in the case of
the smooth wall (Eq. 2.46) except that (1) will drop out (no inuence of
viscosity) and (2) ks will be added to the list of the independent quantities:
0,
y, , ks
(2.71)
42
Figure 2.14: Turbulence quantities. Rough wall. Averaging over both time and
space. Sumer et al. (2001).
Fig. 2.14 shows data regarding u 2 , v 2 , and u v . The data are plotted in a format slightly di erent from that in the above equation, namely,
the actual roughness height is used to normalize y, rather than Nikuradses equivalent sand roughness. Notice the large di erence in the roughness
Reynolds numbers of the two sets of data. Despite this, the good agreement
between these two data may imply that the functions g1, .., g4 are universal.
Fig. 2.15 compares the turbulence quantities obtained in another study;
in these tests, everything else is maintained unchanged, but the wall roughness is changed, to see the inuence of the wall roughness on the end results.
From the gure, the following two points are worth emphasizing:
1. The inuence of the wall roughness disappears with the distance from
the wall, as expected.
2. The distribution of turbulence across the depth is more uniform in the
case of the rough wall than in the case of the smooth wall.
43
2.2
44
(Fig. 2.2),
=
y
)
h
0 (1
(2.73)
, , h)
(2.74)
However, from Eqs. 2.73 and 2.74, the preceding equation may be written as
u = F ( 0 , y,
, , h)
(2.75)
du
, 0 , y, , h) = 0
dy
and from dimensional analysis one obtains
1(
du Uf
=
dy
h
1(
y
)
h
Integrating gives
u = Uf
1(
y
y
) d( ) + c
h
h
y=h:
one obtains
U0 = Uf
1(
y=h
y
y
) d( )
h
h
(2.77)
(2.78)
45
y
U0 = Uf f1 ( )
h
(2.79)
in which U0 is the velocity at the free surface in the case of the open-channel
ow, and the center line velocity in the case of the pipe ow. The preceding
equation is known as the velocity-defect law. This velocity distribution is
satised in the outer-ow region, Fig. 2.16.
The explicit form of the function f1 ( hy ) can be found in the following way.
As sketched in Fig. 2.16, the outer region and the logarithmic layer is
expected to overlap over a certain y interval. Therefore, in the overlapping
region, the two velocity distributions, namely the logarithmic law (Eq. 2.42)
and the velocity-defect law (Eq. 2.79), should be identical; therefore from
Eqs. 2.42 and 2.79,
AUf ln(
yUf
) + BUf
y
U0 + Uf f1 ( )
h
(2.80)
(2.81)
y
U0 = Uf (A ln( ) + B )
h
(2.82)
46
Figure 2.17: Velocity distribution over the entire section of the pipe.
1. A is the same universal constant as in the analysis for the near-wall
ow, i.e., A = 2.5.
2. B , on the other hand, is found from experiments as B = 0.8 for pipe
ows (Monin and Yaglom, 1971). However, if Eq. 2.82 is assumed to
be applicable for the entire cross section, then B has to be B = 0.
3. Experiments show that the velocity-defect law begins to deviate from
the measured velocity proles (not very signicantly, however) for hy >
0.2 0.3 (Fig. 2.17).
4. When the general form of the equation for the velocity is considered
(Eq. 2.76), it will be seen that U0 is a function of the Reynolds number.
From Eq. 2.80,
U0 = AUf ln(
hUf
) + Uf (B
B)
(2.83)
u(y) dy
y=0
and
V = U0
Uf (A
B)
(2.84)
47
30h
)
ks
Uf B
(2.86)
48
Turbulence quantities.
The functional forms of the turbulence quantities depicted in Eq. 2.47
hU
will include one additional parameter in the present case, namely f , or
alternatively hV , the Reynolds number.
Figs. 2.18 and 2.19 illustrates how this latter parameter inuences the
turbulence quantities u 2 and v 2 (Wei and Willmarth, 1989). As seen,
near the wall, the e ect of Re disappears, as expected (see Section 2.1.2).
However, away from the wall, the Reynolds number signicantly inuences
2
2.3
Turbulence-modelling approach
The preceding sections give a considerable insight into the process of turbulent boundary layers in channels/pipes, and over walls. However, the method
used for the analysis does not enable us to do a parametric study, a study
49
where the inuence of various parameters on the boundary layer process can
be studied in a systematic manner. This can be achieved with the aid of
turbulence modelling. A detailed account of turbulence modelling is given in
Chapter 6. We restrict ourselves here only with the so-called mixing-length
model and use it to address the problem of determining the mean velocity
distribution as function of the distance from the wall.
2.3.1
Smooth wall.
Our objective is to obtain the velocity distribution, u, near the wall (in
the constant-stress layer). The x component of the Reynolds equation leads
to the following equation (Eq. 2.7):
(=
du
+(
dy
u v )) =
(2.87)
du
dy
(2.88)
(2.89)
50
2
m
du
dy
(2.92)
The third step is to determine the mixing length in the preceding equation. van Driest (1956) gives the mixing length as follows
m
= y[1
exp(
y+
)]
Ad
(2.93)
in which is the von Karman constant (=0.4), and Ad is called the damping
coe cient, and taken as 25. As seen,
m
(2.94)
while
0, as y +
(2.95)
The latter implies that there will be no turbulence mixing as the wall is
approached.
Eqs. 2.89, 2.92 and 2.93 are the model equations (three equations); so,
there are four equations altogether (along with the Reynolds equation, Eq.
2.87), and four unknowns, namely, u,
u v , t , and m (the latter already
given by Eq. 2.93). Hence, the system is closed. Therefore the unknowns can
be determined. Let us focus on u.
From Eqs. 2.87, 2.89 and 2.92, one gets
2 du 2
)
m(
dy
du
=
dy
(2.96)
51
dy +
u = 2Uf
0
1+ 1+4
2 y +2 [1
exp(
y+ 2
)]
Ad
1/2
(2.97)
(2.99)
2. Furthermore, the van Driest prole agrees remarkably well with the
measured velocity prole including the bu er layer (5 < y + < 70).
52
Figure 2.21: van Driest velocity proles for di erent values of roughness.
Regarding the other two unknowns,
u v and t , inserting Eqs. 2.93
and 2.97 in Eq.2.92, t is obtained, and from this and Eqs. 2.97 and 2.89,
u v is obtained.
Transitional and rough walls.
In this case, the mixing length may be given as (Cebeci and Chang, 1978)
m
= (y +
y)[1
exp(
(y + + y + )
)]
Ad
(2.100)
ks+ exp(
ks+
)];
6
(2.101)
53
ks Uf
(2.102)
dy +
u = 2Uf
0
1+ 1+4
2 (y +
y + )2 [1
exp(
(y+ + y + ) 2
)]
Ad
1/2
(2.103)
Fig. 2.21 displays the velocity proles obtained from Eq. 2.103 including the
smooth-wall prole. In Fig. 2.21, a represents the linear prole in the viscous
sublayer (Eq. 2.98), b the logarithmic prole in the case of the hydraulically
smooth wall (Eq. 2.99), and c the logarithmic prole in the case of the
completely rough wall (Eq. 2.65), namely, by putting z = y + y,
30z
)
ks
z being measured from the theoretical wall.
(2.104)
u = AUf ln(
y.
ln y + + c(0)]
(2.106)
Now, in the case of the smooth wall, we know that there are three distinct
regions, namely,
54
55
ln(y + +
y + ) + c(0)]
(2.107)
ln(y + ) + c(0)]
U ( y+)
(2.109)
On the other hand, this velocity is given independently in Eq. 2.105. So,
setting these two equations equal,
Uf [
ln(y + ) + c(0)]
U ( y+)
one gets
c(ks+ )
c(0) +
Uf [
ln y + + c(ks+ )]
U ( y+)
=0
Uf
(2.110)
(2.111)
56
dy +
U ( y ) = 2Uf
0
1+ 1+4
y +2 [1
exp(
y+ 2
)]
Ad
1/2
the coordinate shift y + can be calculated from Eq. 2.111. Rotta (1962)
carried out this calculation. Fig. 2.25 shows the result.
It may be noted that the variation depicted in Fig. 2.25 agrees fairly well
with the expression given by Cebeci and Chang (1978) in Eq. 2.101.
As a nal remark, the coordinate shift normalized by the roughness
height, ky , can be found from Eq. 2.101 (where ks may be taken as 2k).
The result of this exercise is plotted in Fig. 2.26. As seen, the ky values
exhibited in the gure agree rather well with the experimental observations
that the level of the theoretical bed lies (0.15-0.35)k below the top of the
roughness elements, given in conjunction with the theoretical bed in Section
57
2.3.2
The velocity distribution given in Eq. 2.103 is obviously valid near the wall.
When applied for the entire depth, the velocities predicted from this equation are apparently smaller than the measured velocities. This di erence is
corrected by the so-called Coles wake function (see Coleman and Alanso,
1983)
y+
y+
( ) ( + ) = ( ) 2 sin2 ( + )
(2.112)
h
2h
in which is a wake strength coe cient, h is the ow-depth/pipe-radius, or
the boundary-layer thickness.
When this equation is used together with Eq. 2.103, it does not produce
equal to zero at the center line of a pipe, or at
a prole with a derivative du
dy
the free-stream edge of the boundary layer over a wall (clearly, in the case of
58
pipe ows and boundary-layer ows, this derivative has to be zero). To avoid
this drawback, Coles wake function has been corrected. The nal form of
the velocity expression (including the corrected wake function) is as follows
(Coleman and Alanso, 1983):
u
Uf
y+
2 dy +
=
0
1+ 1+4
y+ 2
+( + ) (1
h
2 (y +
y + )2 [1
y+ 2
y+
2
) + ( ) ( + ) [3
h+
h
(2.113)
1
exp(
(y + + y+ ) 2
)]
Ad
y+
2 +]
h
2.4
59
Flow resistance
The question here is that, given the mean ow velocity V, how we can calculate the friction velocity Uf , i.e.,
Uf =
(2.114)
(2.115)
(2.118)
3.75Uf
Combining Eq. 2.116 with Eq. 2.118, we obtain the following equation
(Schlichting, 1979, p. 610)
1 V
VD
Uf
= 2.035 log(
8 )
V
8 Uf
0.91
(2.119)
in which D is the pipe diameter. (Note that log in the above equation should
not be confused with the natural logarithm). The coe cients can be adjusted
slightly so that the preceding equation agrees with the experimental data.
This latter exercise gives the following relation
1 V
VD
Uf
= 2.0 log(
8 )
V
8 Uf
0.8
(2.120)
60
This is Prandtls universal law of friction for smooth pipes (Schlichting, 1979,
p.611).
In the case of a rough circular pipe, we can do the same exercise, but this
time with the velocity distribution given by Eq. 2.65
u = AUf ln(
30y
)
ks
(2.121)
30R
)
ks
(2.122)
and combining Eqs. 2.122 and 2.118, we get (Schlichting, 1979, p. 621)
Uf
=
V
1
8(2 log kRs + 1.68)
(2.123)
A comparison with Nikuradses experimental results shows that closer agreement can be obtained, if the constant 1.68 is replaced by 1.74 (Schlichting,
1979, p.621):
Uf
1
=
(2.124)
V
8(2 log kRs + 1.74)
In the case of a transitional-wall-category circular pipe, Colebrook and
White obtains the following resistance relation (Schlichting, 1979, p.621)
1 V
= 1.74
8 Uf
2 log(
18.7
ks
)
+ VD
R ( )( 8 Uf )
V
(2.125)
Given the mean-ow velocity, the friction velocity can easily be calculated
from Eqs. 2.120, 2.124 or 2.125.
In the case of a non-circular pipe ow or an open ow, the preceding
resistance relations can be used provided that the pipe diameter D should
be replaced by 4rh in which rh is the hydraulic radius, equal to the crosssectional area divided by the wetted perimeter (Schlichting, 1979, p. 622).
For example, in the case of an open channel with a rough bed, the resistance
relation can be found as
V
14.8rh
= 5.75 log(
)
Uf
ks
(2.126)
61
V
14.8rh
= 2.5 ln(
)
Uf
ks
(2.127)
It may be noted that the ratio V /Uf may be traditionally written in terms
of the friction coe cient, f, as
Uf =
f
V
2
(2.128)
2.5
Bursting process
Experimental work conducted in 1960s and 70s has shown that the nature
of the ow pattern near the wall in a turbulent boundary layer is repetitive;
the ow near the wall occurs in the form of a quasi-cyclic process, called the
bursting process. Reviews of the subject can be found in Laufer (1975), Hinze
(1975, pp. 659-668), Cantwell (1981), Grass (1983) and Nezu and Nakagawa
(1993).
62
63
64
Figure 2.30: (a): ejection; and (b): in-rush events. Grass (1971).
ejection event (Fig. 2.30, a), and the in-rush event (Fig. 2.30 b).
The bursting process has been studied in the case of rough walls, and it
was found that practically the same kind of quasi-cyclic process takes place
(Grass, 1971, and Grass et al., 1991).
The rest of this section is concerned with quantitative information on the
bursting process.
1. The mean spacing of the low-speed wall streaks (Fig. 2.29, Frame 1) is
+
Uf
100
(2.129)
(2.130)
(Nychas et al., 1973). Corino and Brodkey (1969), on the other hand,
report that lower half of the viscous sublayer y + < 2.5 is essentially
65
passive and the rest (2.5 < y + < 5) active, being inuenced by the
quasi-cyclic, bursting events occurring in 5 < y + < 70.
3. The x and z extents of ejected low-speed uid structures involve
dimensions of the order of
20 to 40 of x+ , and 15 to 20 of z +
(2.131)
100
(2.132)
(Nychas et al., 1973). Praturi and Brodkey (1978) report that, in some
rare cases, the ejected uid elements travel up to 300 y + ).
5. The mean streamwise distance from the onset of lift-up of a low-speed
wall streak to the break-up of any sign of coherency is about
1300 in x+
(2.133)
(2.134)
(2.135)
(Jackson, 1976). Note that this, too, scales with the outer ow parameters.
66
2.6
We may write the unit of any hydrodynamic quantity in the following form
(2.136)
[A] = L T K
in which [A] is the unit of A, and L, T, and K are the fundamental units,
namely, L is the length, T the time, and K is the force. , and are
positive or negative integers, or they may be zero.
Now, Buckinghams Pi Theorem states the following.
A dimensional function
f (A1 , A2 , ...., An ) = 0
(2.137)
2 , ...... n r )
(2.138)
=0
n r
i)
(2.139)
should be selected so
67
(2.140)
, )
or
f ( 0,
(2.141)
, , u, y) = 0
2)
(2.142)
=0
and
x1 y1
0
z1
(2.143)
x2 y2
0
z2
(2.144)
and
The left and right hand sides of Eq. 2.143 should have the same units.
Writing the units of the quantities on both sides of the equation according
to Eq. 2.136,
L0 T 0 K 0 = (KL 2 )x1 (KL 2 T )y1 (KL 4 T 2 )z1 (LT
(2.145)
(2.146)
or
L0 T 0 K 0 = L
68
which requires
(2.147)
2y1 4z1 + 1 = 0,
y1 + 2z1 1 = 0,
x1 + y1 + z1 = 0
2x1
This gives
1
1
, y1 = 0, and z1 =
2
2
Hence the rst nondimensional quantity
x1 =
1/2 0 1/2
0
u=
u
u
=
1/2
( )
Uf
0
(2.148)
(2.149)
(2.150)
and
2x2
(2.151)
2y2
4z2 + 1 = 0,
y2 + 2z2 = 0,
x2 + y2 + z2 = 0
and
1
x2 = , y2 = 1, and
2
Then, the second nondimensional quantity
2
1/2
0
1 1/2
y=
z2 =
y( 0 )1/2
1
2
yUf
(2.152)
(2.153)
u yUf
)=0
,
Uf
(2.154)
(2.155)
2.7. REFERENCES
2.7
69
References
70
2.7. REFERENCES
71
24. Praturi, A.K. and Brodkey, R.S. (1978): A stereoscopic visual study
of coherent structures in turbulent shear ow. J. Fluid Mech., vol. 89,
p.251.
25. Rotta, J.C. (1962). Turbulent boundary layers in incompressible ow.
Progress in Aerospace Science, vol. 2, p.1.
26. Schlichting, H. (1979). Boundary-Layer Theory. McGraw-Hill.
27. Sumer, B.M. (1986): Recent Developments on the Mechanics of sediment suspension, General-Lecture Paper, in the book: Euromech 192:
Transport of suspended solids in open channels, A.A. Balkema Publishers, Rotterdam, 1986.
28. Sumer, B.M., Cokgor, S. and Fredse, J. (2001): Suction removal of
sediment from between armour blocks. Journal of Hydraulic Engineering, ASCE, vol. 127, No. 4, pp. 293-306, 2001.
29. Sumer, B.M. and Deigaard, R. (1981): Particle motions near the bottom in turbulent ow in an open channel - Part 2. Journal of Fluid
Mechanics, Vol. 109, p. 311.
30. Sumer, B.M. and Oguz, B. (1978): Particle motions near the bottom
in turbulent ow in an open channel. Journal of Fluid Mechanics, Vol.
86, p. 109, 1978.
31. Sumer, B.M., Kozakiewicz, A., Fredse, J. and Deigaard, R. (1996):
Velocity and concentration proles in the sheet ow layer of movable
bed. Journal of Hydraulic Engineering, ASCE, vol. 122, No. 10,
549-558.
32. van Driest, E.R. (1956): On turbulent ow near a wall. J. Aeronautical Sciences, vol. 23, p.1007.
33. Wei, T. and Willmarth, W.W. (1989): Reynolds-number e ects on
the structure of a turbulent channel ow. J. Fluid Mech., vol. 204, p.
57.
72
Chapter 3
Statistical analysis
As has been seen in the previous chapters, hydrodynamic quantities such
as the velocity components appear to be random variables. Although recent research has shown that the near-wall turbulent-boundary-layer process
is repetitive in nature (Chapter 3, Bursting Process), unless special pattern recognition techniques are used to detect these repetitive, quasi-cyclic
events, the turbulence signals even near the wall appear to be largely random. Therefore, it is imperative to use statistical techniques to study the
turbulence.
Three topics will be discussed in the present chapter. First we will briey
study the probability density function of the turbulent velocity, then we will
concentrate on the correlation and subsequently spectral analysis of turbulence. In the entire analysis, it will be assumed that the turbulence is isotropic
unless otherwise stated.
3.1
Consider, for example, the x component of the velocity, u (Fig. 3.1 a). The
velocity u (or its uctuating component u ) is a random variable. Thus, u
(or u ) must have a probability density function (p.d.f.) such that
p(u ) du = Pr[u < U < u + du ]
(3.1)
and
p(u ) du = 1
73
(3.2)
74
(u
u)2 =
u2
(3.4)
75
(3.5)
u p(u) du
(u
u)2 p(u) du =
u 2 p(u ) du
(3.6)
u)3 p(u) du =
u 3 p(u ) du
(3.7)
(u
76
u3
(
(3.8)
2)2
u
(u
u)4 p(u) du =
u 4 p(u ) du
(3.9)
3.2
u4
( 2u )2
Correlation analysis
There are two kinds of correlations: (1) Space correlations, and (2) time
correlations. We will rst focus on the space correlations.
3.2.1
Space correlations
Denitions.
Fig. 3.3 gives the denition sketch. Consider the i component of the
uctuating velocity at Point A, (ui )A , and j component of the uctuating velocity at Point B, (uj )B . Suppose that these velocities are measured
simultaneously.
Two-point double correlation is dened by
(Oij )AB = (ui )A (uj )B
(3.10)
in which the overbar indicates the time averaging (Eq. 1.1). There are
other correlations as well: three-point, four-point,..., and triple, quadruple,...correlations.
Two most important two-point double correlations (Fig. 3.4) are
u (x) u (x + r)
(3.11)
v (x) v (x + r)
(3.12)
and
77
(3.15)
78
(3.16)
1, and g(r)
(3.17)
0, then u 2
u (x)u (x + r)
4. The correlation coe cients tend to zero as the separation distance goes
to innity
f (r)
0, and g(r)
0, as r
(3.18)
This is because when r is very large, there will be as many u (x)u (x+r)
products with the negative sign as with the positive sign, therefore,
when averaged, u (x)u (x + r) will be zero. Physically, r is so large
that the two points will not lie in the same coherent structure, therefore, their motions will be uncorrelated, meaning that the correlation
coe cients should be zero for such large r distances.
79
(3.19)
1
rg(r)dr = [r2 f (r)]0 = 0
2
(3.20)
80
Scales of turbulence.
Micro scales. Expanding f (r) as a Taylor series, and for small values
of r, gives
1 2f
f (r) = 1 + ( 2 )r=0 r2 + ...
(3.22)
2 r
Note that the terms with rn (n being an odd number, n = 1, 3, 5, ..) will drop
because f (r) = f ( r).
Now, the formal denition of the micro time scale is
1
2
f
1 2f
(
)r=0
2 r2
(3.23)
r2
(3.24)
2
f
As seen from Fig. 3.5, this is a parabola t to the correlation coe cient for
small values of r , and f is the x intercept of the tted parabola.
Likewise,
r2
(3.25)
g(r) = 1
2
g
1
v 2
)
(
2
r
2v
(3.27)
g(r))r=0 = 2(
2
r2
f (r))r=0
(3.28)
81
(3.29)
f (r)dr
(3.30)
g(r)dr
(3.31)
and
g
=
0
This length scale is indicated in Fig. 3.6 where the area of the rectangle
1 is equal to the area under the correlation curve.
g(r)dr =
0
1
f (r)dr + [rf (r) |0
2
1
2
f (r)dr
(3.32)
which gives
f
3.2.2
=2
(3.33)
Time correlations
There are two kinds of time correlations: Eulerian time correlations, and
Lagrangian time correlations.
82
u ( )u ( + t)
u2
(3.34)
This is for the x component of the velocity. Likewise, other Eulerian correlation coe cient can also be dened for the y and z components of the
velocity.
Similar to space correlations, one has
RE (0) = 1
(3.35)
RE (t)
(3.36)
0, as t
(3.37)
and
RE (t)
1
(
2
RE (t)
)t=0
t2
(3.38)
(cf. Eq.3.23), or
RE (t) = 1
t2
2
E
(3.39)
RE (t)dt
(3.40)
The macro time scale can be interpreted as the time during which a coherent
structure (or an eddy) passes the measurement point.
We expect that the macro time scale and the macro length scale are related. To nd this relation, G.I. Taylor made the so-called frozen-turbulence
approximation. Consider the coherent structure at time t in Fig. 3.7. The
size of this coherent structure is f , the macro length scale of turbulence.
83
= uTE
(3.41)
This relationship enables the length scale to be calculated from the knowledge
of the time scale. (The latter is easier to measure than the former).
The Lagrangian time correlation is dened in the same way as in the case
of the Eulerian time correlation. However, the velocity is now measured by
tracking the uid particle, as the particle travels. Therefore, the Eulerian
time series (which is the basis for the Eulerian time correlation) is to be
replaced by the time series from the Lagrangian particle tracking. Then, all
the equations given for the Eulerian correlation (Eqs. 3.34 - 3.41) are equally
valid for the Lagrangian correlation.
84
V ( )V ( + t)
V 2
(3.42)
in which V is the Lagrangian uctuating velocity, and, for later use, the
Lagrangian macro time scale
TL =
RL (t) dt
(3.43)
3.3
Spectrum analysis
85
3.3.1
General considerations
The Fourier transform of two-point double correlation Qij (see Eq. 3.10 for
the denition of Qij ) is
ij (k)
1
(2 )3
Qij (r) e
ikr
dr
(3.44)
in which r is the separation vector (Fig. 3.8), dr =dr1 dr2 dr3 , k is the wave
number vector and i is the imaginary unit, i =
1. The inverse transform,
on the other hand, is
Qij (r) =
ij (k)
eikr dk
(3.45)
ij (k)
dk
Qii (0) =
ii (k)
dk
and setting i = j,
(3.46)
86
(3.47)
ii (k1, k2 , k3 )
(3.48)
Since the left hand side of the above equation is proportional to the turbulent
energy, then ii (k1, k2 , k3 ) will be like the energy density in the wave number
space. We shall return to this in the following paragraphs.
In the case of isotropic turbulence, there will be spherical symmetry.
Therefore, the spherical coordinates can be used in the wave number space.
Eq. 3.48 in this case will read
Qii (0) =
ii (k)
k=0
ii (k)
=
k=0
dk dA(k)
A(k)
dk
(4 k 2 )
dA(k) =
A(k)
k=0
ii (k)
dk
87
or
1
Qii (0) =
2
E(k) dk
(3.49)
k=0
in which
E(k) = 2 k2
ii (k)
(3.50)
3.3.2
88
rk rk
Qij (r)
(3.52)
in which
Sij (r) =
rk
(3.53)
(Sik,j + Skj,i )
in which
(Sik,j )AB = (ui )A (uk )A (uj )B
(3.54)
(3.55)
and
It should be noted that Eq. 3.52 stems from the N.-S. equation.
The Fourier transform of Eq. 3.52
t
ij (k)
in which
ij (k)
2 k2
ij (k)
1
(2 )3
ij (k)
(3.56)
ikr
(3.57)
Sij (r) e
dr
ij (k)
ikr
dk
(3.58)
ii (k)
(3.59)
ii (k)
2 k2
E(k) = T (k)
2 k2 E(k)
(3.60)
in which
T (k) = 2 k2
ii (k)
(3.61)
89
Figure 3.10: Energy spectrum and dissipation spectrum. (Note that peaks are
well apart).
E(k)dk =
0
T (k)dk
k 2 E(k)dk
(3.62)
It can be shown that the rst term on the right hand side of the preceding
equation is zero (Hinze, 1959, p. 178). Hence Eq. 3.62 will be
t
E(k)dk =
0
k2 E(k)dk
(3.63)
Recall that 0 E(k)dk is the turbulent energy per unit mass (Eq. 3.49);
hence, the left hand side of the above equation will represent the rate of
change of the turbulent energy. Now, the integral on the right hand side of
the equation, 0 k2 E(k)dk, is always positive. This means that
t
E(k)dk < 0
(3.64)
90
E(k)dk =
0
T (k)dk
0
k 2 E(k)dk
(3.65)
Consider the interval in the wave number space from k = 0 to k (Fig. 3.10).
The rst term on the left hand side of the above equation represents the
change in the turbulent energy in this interval per unit time. The third term
represents the viscous dissipation in this interval per unit time (Fig. 3.11).
k
Experiments show that the sign of the term 0 T (k)dk is the same as
k
that of the viscous dissipation. Therefore, 0 T (k)dk should be an e ux,
rather than an inux. This e ux cannot take place at the k = 0 end of the
considered interval, because k = 0 forms a closed end (there is no negative
91
3.3.3
In the process of turbulent energy transfer from small wave numbers (large
scale motion) to large wave numbers (small-scale motion), a point is reached
beyond which the turbulence will forget the inuence of the initial largescale motion. From this point to the wave numbers where the viscous dissipation occurs, there is a range of wave numbers in which the character of
turbulence will be the same in all ows (in streams, in atmospheric boundary-
92
layer turbulence, etc.). This range is called the universal equilibrium range.
It is called universal because the character of turbulence in this range is independent of the ow environment; and it is called equilibrium because the
energy input is in balance with the energy dissipation, as will be detailed
later in the section. Kolmogoro s theory is concerned with this range.
Kolmogoro s rst hypothesis states that, irrespective of the large-scale
motion, the turbulence in the universal equilibrium range is isotropic. Now,
the energy spectrum function in this range is a function of the following
quantities
E = f ( , , k)
(3.66)
in which
time:
is the dissipation of turbulent energy per unit mass and per unit
= (
ui
u
u
+
) i
x
xi x
(3.67)
(Eq. 1.32).
is involved because it represents the dissipation of turbulent energy.
should be involved because the energy is dissipated into heat through viscosity. The wave number, k, is involved because the turbulent energy process is
primarily dependent on the wave number (Eq. 3.60). (k is taken instead of
k1 , k2 and k3 because of the spherical symmetry; isotropic turbulence!). Note
that E (rather than ii ) is adopted here to represent the energy spectrum
because we assume that turbulence is isotropic.
From dimensional analysis, one gets
E(k) = v2
in which
(3.68)
( k)
(3.69)
v = ( )4
and
is
3
=(
2
(3.70)
)4
1
cm 2 4
Here the unit of v is [v] = ( cms cms ( s.cm
) ) = cm
, and that of is
s
1
2
1
cm 3
[ ] = (( s ) cm2 cm 2 ) 4 = cm. The quantity v is called the characteristic
s
( s.cm )
= ( )2
(3.71)
93
The above characteristic length, , and the characteristic time, , are involved in the large wave number range (small scale motion); they represent
the smallest scales of turbulence, namely the smallest length scale ( known
as the Kolmogoro scale), and the smallest time scale. These are the smallest scales of turbulence because scales smaller than these involve molecular
motion.
The nondimensional equation obtained for the spectrum function (Eq.
3.68) can be developed further, following Kolmogoro s second hypothesis.
The latter states that the peaks of the energy spectrum and that of the
dissipation spectrum are su ciently apart. This is illustrated in Fig. 3.10.
The latter implies that there is a dominating wave number range in the
energy spectrum (small wave numbers, or alternatively, large eddies) which
contain most of the turbulent energy, while there is another dominating wave
number range (large wave number, or alternatively small eddies) which are
associated with the viscous dissipation. According to Kolmogoro s second
hypothesis, these wave number ranges are su ciently apart, meaning that
there must be a wave number range between these two ranges where there is
no signicant energy, and there is no signicant dissipation.
First of all, this range should be a part of the universal equilibrium range,
because it contains no signicant energy. (Recall the denition of the universal equilibrium range at the beginning of this subsection: In the universal
equilibrium range, the energy input is in balance with the energy dissipation;
no signicant energy presence).
Secondly, in this range, the only signicant activity is that the turbulent
energy is transferred from small wave numbers to large wave numbers. This
process is governed by the inertia terms in the N.-S. equation, as described
in Example 7 above. Therefore, this particular range of wave numbers is
called the inertial subrange. (inertial because of the previously mentioned
inertia terms; and subrange because this range is a part of the universal
equilibrium range, as pointed out in the preceding paragraphs).
Now, in the inertial subrange, the energy spectrum should be independent
of the viscosity because no signicant dissipation takes place in this range,
as mentioned previously. This is an important statement. Because, for the
viscosity to drop out in Eq. 3.68, the function ( k) has to be given in the
following form
( k) = Constant (k )
5
3
(3.72)
94
2
3
5
3
(3.73)
3.3.4
One-dimensional spectrum
F (k) =
Q11 (r) e
u 2 f (r) e
ikr
dr
ikr
(3.74)
dr
u2
f (r) cos(kr) dr
0
(3.75)
95
F (k) cos(kr) dk
(3.76)
It can be shown that, in isotropic turbulence, the one-dimensional spectrum F (k) is related to the energy spectrum function E(k)
E(k) = k2 F (k)
(3.77)
kF (k)
5
3
= k 2 F (k)
kF (k)
(3.78)
k2
C1 + C2
2
(3.79)
9
55
2
3
5
3
The second and third terms on the right-hand-side of the preceding equation
should drop because F (k) should go to zero for large values of k, and therefore
F (k) = Constant
2
3
5
3
(3.80)
u2
f (r) cos(kr) dr =
0
For k = 0:
k = 0 : F (0) =
u2
u2
(1 + (k f )2 )
f
96
Figure 3.13: 1-D and 3-D energy spectra for the numerical example.
Hence,
F (k) =
F (0)
1 + (k f )2
(k f )4
(1 + (k f )2 )3
97
Figure 3.14: 1-D energy spectra. (a): measured in an open channel. Raichlen
(1967). (b): measured in a jet. Champagne (1978).
98
5
3
revealing the previously mentioned Kolmogoro s -5/3 law. The same holds
true for the second example (Fig. 3.14 b) for the range 4 10 2 cm 1 < k1 <
10 cm 1 .
3.4
References
Chapter 4
Di usion and dispersion
Di usion/dispersion is another important property of turbulent ows. Some
examples are: (1) Waste products released into a recipient environment such
as a stream or the ocean undergo rapid di usion; (2) Smoke (or exhaust gases)
released into the atmosphere di uses rapidly when there is strong wind; (3)
Likewise, dredged material (sand, silt, etc.), when dumped into the sea, is
subjected to large di usion.
The objective of this chapter is to describe the di usion and dispersion
processes in turbulent ows. First, we will study the di usion process in a
turbulent ow eld (a general analysis), and then, we will turn our attention to the so-called longitudinal dispersion process, a process caused by the
combined action of the turbulent di usion and the velocity variation in the
ow.
4.1
One-Particle analysis
100
101
dx
dt
(4.1)
102
Integrating gives
t
(4.2)
V (t ) dt
x(t) =
0
Now, ensemble averaging gives the mean particle position (the position of
the centroid of the particle cloud):
t
x(t) =
V (t ) dt =
(4.3)
V (t ) dt
(4.4)
in which is any hydrodynamic quantity, n the value of in one realization, and N the total number of realizations. (In the present context, one
realization is one release of particle).
The variance, on the other hand, can be calculated in the following way.
Let
y(t) = x(t) x(t)
(4.5)
Then, from Eq. 4.2
t
y(t) = x(t)
x(t) =
V (t ) dt
0
V (t ) dt
(4.6)
[V (t )
V (t )] dt
(4.7)
V (t ) dt
(4.8)
V =V
Hence,
t
y(t) =
0
(4.9)
103
(4.10)
V (t ) dt ] V (t)
0
t
V (t ) V (t) dt
= 2
0
V (t )V (t)
= RL (t
V 2
Then,
d 2
y (t) = 2 V
dt
To summarize at this point,
t) V
(4.11)
t
2
RL (t
t ) dt
(4.12)
RL (t) dt
(4.13)
(Eqs. 3.42 and 3.43). Recall that Lagrangian macro time scale, TL , may be
interpreted as the time during which a coherent structure keeps its identity
/ coherence (Section 3.2.2).
4.1.1
V (t = 0)
u(x = 0)
(4.14)
104
x(t) =
V (t ) dt =
0
u(x = 0) dt = u(x = 0)
0
dt = u(x = 0) t
0
(4.15)
Regarding the variance, for small times (t
TL ), rst of all, V 2
u 2 (x = 0), and secondly, RL (t t ) 1. Therefore, Eq. 4.12
d 2
y (t) = 2 u 2 (x = 0) t
dt
and integrating gives
y 2 (t) = u 2 (x = 0) t2
(4.16)
Now, assume that the ow/turbulence is homogenous. Then the statistical properties of velocity should be invariant with respect to the position:
u(x = 0)
u(x) = u = constant
(4.17)
u 2 (x) = u 2 = constant
(4.18)
and
u 2 (x = 0)
Thus, the mean particle position and the variance will be (from Eqs. 4.15,
4.16, 4.17, 4.18)
x(t) = u t
(4.19)
and
y 2 (t) = u 2 t2
(4.20)
The p.d.f. of the particle position, on the other hand, will be the same
as that of the Eulerian velocity provided that u should be replaced with xt .
4.1.2
p[V (t + )]
(4.21)
105
meaning that the statistical properties of the particle velocity such as the
mean V and the variance (V V )2 = V 2 are time invariant. Homogeneity
in space is equivalent to being stationary in time).
First, consider the mean particle position. From Eq. 4.3 (taking V (t )
outside the integral because it is constant),
t
x(t) =
(4.22)
V (t ) dt = V t
0
The mean particle velocity should be equal to the Eulerian mean velocity,
u. (The fact that the Lagrangian velocity is a stationary random function
of time requires the homogeneity in the Eulerian sense; therefore, V
u).
Hence,
x(t) = u t
(4.23)
Next, consider the variance of the particle position. Eq. 4.12:
d 2
y (t) = 2 V
dt
t
2
RL (t
(4.24)
t ) dt
(4.25)
t
2
(4.26)
RL ( ) d
=0
y 2 (t) = 2 V 2 (
RL ( ) d ) t
=0
2V
RL ( ) d
(4.27)
=0
y 2 (t) = 2 V 2 (
RL ( ) d ) t
=0
(4.28)
106
RL ( ) d
RL ( ) d = TL
=0
(4.29)
=0
(from Eq.4.13). Inserting the latter equation in Eq. 4.28, and also considering
that V 2 is identically equal to u 2 (due to the fact that V (t) is a stationary
random function of time), the variance is obtained as
y 2 (t) = 2 u 2 TL t
(4.30)
TL )
(4.31)
y 2 (t)
TL )
(4.32)
and
This implies that the spreading occurs very fast in the beginning of the
process (y 2 (t) t2 ), and then the rate at which the spreading occurs reaches
an equilibrium stage where y 2 (t) t, as sketched in Fig. 4.4.
P.d.f. of particle position.
The particle displacement (Eq. 4.2)
t
V (t ) dt
x(t) =
0
can be considered as the sum of a series integral over a xed interval, say ,
t
x(t) =
(n+1)
V (t ) dt =
0
V (t ) dt
n
(4.33)
such that each integral has only a partial statistical connection with each
other. For times t
TL , the number of such integrals becomes su ciently
large so that we have a situation like that covered by the central limit theorem.
107
p(x) =
2
exp(
x)2
(x
(x
x)2
2(x
x)2
p(y) =
2
exp(
y2
(4.34)
x,
y2
)
2 y2
(4.35)
p(y) =
2
2 u 2 TL t
exp(
y2
)
2 (2 u 2 TL t)
(4.36)
(4.37)
108
in which D is the di usion coe cient, y the space coordinate, and t the time.
Consider the initial condition
t=0:
(4.38)
c = (y)
1
2
2Dt
exp(
y2
)
2 (2 D t)
(4.39)
Now, compare the above equation with Eq. 4.36. From this comparison,
the following deductions can be made:
1. The p.d.f. of the particle position for large times (t
di usion equation (cf., 4.37)
2
p
p
=D 2
t
y
TL ) satises the
(4.40)
(4.41)
1 dy 2 (t)
2 dt
(4.42)
109
m
M
(4.43)
1 m
M dy
(4.45)
m
, the number of particles per unit length, is the concentration c in
Here, dy
the one-dimensional case. So,
c(y)
M
p(y) =
(4.46)
, y)
is
m1
M
in which m1 is the number of particles occupying the interval (
wise, this probability can be written as
Pr[
< Y < y] =
(4.47)
, y). Like-
Pr[
p(y) dy
< Y < y] =
(4.48)
m1 = M
p(y) dy
(4.49)
p(y) dy
(4.50)
110
y
m1
=M
p(y)
t
2t
(4.51)
y
p(y)
2Dt
(4.52)
p
y
MD
or
m1
(Mp)
= D
t
y
From Eq. 4.46, c = Mp. Inserting this in the above equation gives
m1
=
t
c
y
(4.53)
(4.54)
111
Changing notation
=
m1
t
c
y
(4.55)
The quantity , the number of particles passing through the section y per
unit time (Fig. 4.5), is called the ux ( the ux due to turbulent di usion).
Notice the similarity between the ux due to turbulent di usion and that
due to molecular di usion, namely mol = Dmol yc (Ficks law) in which
Dmol is the molecular di usion coe cient.
Example 10 Conservation of mass in turbulent ows
c
dy dz
x
(due to convection)
(4.56)
(4.57)
(u c) dx] dy dz
(due to convection)
(4.58)
112
and
[ Dx
c
+ ( Dx
x
x
c
) dx] dy dz
x
direction is
(u c) dx dy dz
and z
(v c) dx dy dz
( Dx
c
) dx dy dz
x
(4.60)
directions:
y
( Dy
c
) dx dy dz
y
(4.61)
c
) dx dy dz
(4.62)
z
z
z
On the other hand, the rate of change of mass in the volume dx dy dz
(w c) dx dy dz
( Dz
(4.63)
(c dx dy dz)
The conservation of mass requires that the rate of change of mass should
be equal to the net inux. Hence, from Eqs. 4.60-4.63, and using the continuity equation (Eq. 1.6), namely
u
v
w
+
+
=0
x
y
z
(4.64)
one gets
c
c
c
c
+u
+v
+w
=
(Dx
t
x
y
z
x
c
) + (Dy
x
y
c
) + (Dz
y
z
c
) (4.65)
z
c
)
xi
(4.66)
Di .
(4.67)
4.2
4.2.1
113
Longitudinal dispersion
Mechanism of longitudinal dispersion
114
Particles near the bottom will experience relatively smaller velocities u(y),
and therefore will be retarded (Path 1 in Fig. 4.7 a) while particles near the
free surface will experience relatively larger velocities u(y), and therefore
will be moved large distances (Path 2 in Fig. 4.7 a). This implies that the
particles are prone to disperse farther and farther apart in the streamwise
direction. Thus, the end result will be the progressive dispersion of the
dispersant in the streamwise direction, as sketched in Fig. 4.7 b. This process
is called the longitudinal dispersion.
When compared with the ordinary di usion process studied in the previous section (cf., Fig. 4.1 and 4.2), the key e ect here is that the dispersion is
caused mainly by the velocity variation u(y) across the depth, whereas the
dispersion in the ordinary di usion process is caused by the random paths of
particles; no spatial variation in the mean velocity exists in the latter process.
For the above reason, the present di usion process is sometimes called
the mechanical dispersion.
From the preceding considerations, it can easily be seen that
1. there will be no longitudinal dispersion, when the velocity distribution
across the depth is uniform; and
2. likewise, there will be no longitudinal dispersion when there is no turbulent motion across the depth (i.e., when there is no turbulent di usion
across the depth). (Note, however, that the existing molecular di usion will cause longitudinal dispersion although the degree of dispersion
in this case is several orders of magnitude smaller than the turbulent
longitudinal dispersion, Taylor, 1953, 1954, Elder, 1959).
4.2.2
Fig. 4.8 shows the path of a particle. Here, V (t) is the streamwise velocity
of the particle. V (t) is actually the sum of two velocities:
1. the velocity u(y) at the position where the particle is at that particular
time t, and
2. the x component of the uctuating velocity u , which occurs at that
position, and at that particular time, t.
115
Figure 4.8: Particle velocity V samples from the velocity prole u(y).
The uctuating velocity is small compared with u(y), and therefore can
be neglected, as already stated in the preceding paragraphs. Hence, the
particle velocity can be taken as V (t) = u(y(t)).
Now, release many particles at time t = 0, and determine their velocities,
V (t), at time t. Let the probability density function (p.d.f.) of these velocities be p(V (t)). Next, do the same exercise, but this time determine their
velocities at time t + . Let the p.d.f. of the latter velocities be p(V (t + )).
Now, if t is su ciently large, then the two p.d.f.s will be identical
p(V (t))
p(V (t + ))
(4.68)
This implies that the velocity V (t) is a stationary random function of time
(Section 4.1.2).
The immediate implication of this is that the theory given in the previous section in conjunction with the ordinary di usion process for large times
(Section 4.1.2) can be applied to the present process because, as in the previous case, the particle velocity is a stationary random function of time. By
application of the one-particle analysis, we reach the following conclusions.
1. The mean particle position (or the position of the centroid of the dispersant) will be at (Fig. 4.7 b)
x(t) = Um t
(4.69)
(from Eq. 4.23). Here, Um is the mean Eulerian velocity, and equal to
the cross-sectional average velocity (Fig. 4.8)
Um =
1
h
u(y) dy
0
(4.70)
116
x)2 = 2 V 2 TL1 t
(4.71)
(from Eq. 4.30). Here TL1 is the Lagrangian time scale dened by
TL1 =
(4.72)
RL1 (t) dt
0
V ( )V ( + t)
V 2
(4.73)
in which
V =V
(4.74)
Um
p(x) =
2
(x
exp(
x)2
(x
x)2
2(x
x)2
(4.75)
1 d(x x)2
2
dt
(4.76)
(from Eq. 4.42). Here, D1 is called the longitudinal dispersion coe cient.
5. The p.d.f. of particle position satises the di usion equation
2
p
p
= D1
t
(x x)2
(4.77)
117
(from Eq. 4.40). Here, t is the time derivative when considering the
moving coordinate x x(t). Changing to the xed coordinate x,
t
x
x(t)
=
+
=
+ Um
x t
t
t x
t
x
(4.78)
(4.79)
<c>
x2
(4.80)
1
h
(4.81)
c dy
0
c
) + (Dy
x
y
c
)
y
(4.82)
Clearly, the one-dimensional equation (Eq. 4.80) is much simpler than the
two-dimensional representation (Eq. 4.82). (This is the advantage of the
longitudinal-dispersion concept). Obviously, this is at the expense of the
calculated concentration. In the one-dimensional representation, the crosssectional average concentration is calculated, whereas, in the two-dimensional
representation, the point-concentration is calculated. However, this is not a
problem, because, in longitudinal-dispersion problems, the calculation with
the cross-sectional average concentration is normally good enough (the concentration variations in the longitudinal direction are important, not the
variations over the cross section).
118
For the application of the previously mentioned one-dimensional dispersion equation, two quantities need to be known: the cross-sectional average
velocity, Um , and the longitudinal dispersion coe cient, D1 .
Um is obtained simply by integrating the velocity distribution over the
cross-section, and dividing it by the cross-sectional area. This is done if
the velocity distribution is known theoretically (for ex., in the form of the
logarithmic law). Otherwise, the velocity distribution over the cross-section
is measured, for example, as in the case of a natural stream, and the mean
velocity calculation is based on the measured velocity.
The longitudinal dispersion coe cient, on the other hand, can be calculated theoretically (see the next section), or it can be calculated numerically,
based on the measured velocity distribution over the cross section, as will be
detailed in Section 4.4.
G.I. Taylor (1953, 1954) introduced the concept of the longitudinal dispersion. In his study, the ow environment was a circular pipe. His theoretical
work and his experiments showed that the longitudinal dispersion coe cient
D1
in a turbulent pipe ow (Taylor, 1954) is aU
= 10 in which a is the radius of
f
the pipe, and Uf is the friction velocity. Taylors work laid the groundwork
for the later research in 1960s and 70s in the area of dispersion in natural
streams, estuaries, etc. and also in industrial ows (such as pipe ows).
4.3
In this section, we shall calculate the longitudinal dispersion coe cient for
an open channel ow (Fig. 4.7). We shall do this exercise for heavy particles
(for example, sand, or silt in water). This feature will be represented by the
fall velocity of the particles, w. However, the case of the neutrally-buoyant
particles/dispersant will be captured by setting w equal to zero in the end
solutions.
The analysis presented in the following paragraphs is mainly based on the
work of Sumer (1974). It may be noted that the method used in the analysis
was rst introduced by Aris (1956) and also employed by Sayre (1968).
4.3.1
119
Formulation
c
c
c
=
( (y) ) + ( (y) )
y
x
x
y
y
(4.83)
in which w is the fall velocity of the particles, and (y) is the turbulent
di usion coe cient. It is assumed that (y) is the same for both x and y
directions.
The boundary conditions:
y = 0 and h:
(y)
c
+ wc = 0
y
(4.84)
c(x, y, 0) =
M
(x)
h
(4.85)
(4.86)
in which Um and Dm are the cross-sectional average velocity and the crosssectional average di usion coe cient, respectively. We shall return to these
quantities and the nondimensional functions (y) and (y) later in the section.
Now, dene the pth moment of concentration
p
Cp =
in which
Cd
(4.87)
Us t
h
(4.88)
in which Us is the mean particle velocity, and C is the normalized concentration dened by
c
(4.89)
C= M
h2
120
Now, multiply the two sides of the above equations (Eqs. 4.83-4.85) by
, and integrate the equations from
to + :
Cp
)pCp
Cp
1
= p(p
Cp
= 0 and 1:
1) Cp
+
s Cp
Cp
=0
) (4.90)
(4.91)
and
= 0:
(4.92)
Here,
Dm t
y
, = 2
(4.93)
h
h
are the normalized distance from the bottom of the channel and the normalized time, respectively. Also, , s and s are the normalized mean ow
velocity, the normalized mean particle velocity and the normalized mean fall
velocity, respectively:
=
Um h
,
Dm
Us h
,
Dm
wh
Dm
(4.94)
Regarding the ow velocity, u(y), and turbulent di usion coe cient, (y),
distributions across the depth, rst of all, from Eq. 2.85,
u
Um =
y
Uf [ln( ) + 1]
h
(4.95)
(the logarithmic law). The distribution of the turbulent di usion coe cient,
on the other hand, is found from the analogy between the momentum ux
(Eq. 2.89) and the ux due to turbulent di usion
=
dc
dy
uv
du
dy
(4.96)
Taking
uv
Uf2 (1 hy ) (see Eq. 2.21), and using the above analogy
(Eq. 4.96), the di usion coe cient
(y) = Uf y(1
y
)
h
(4.97)
121
and Dm is
Dm =
1
h
(y) dy =
0
1
hUf
6
(4.98)
4.3.2
6
2
(1 + ln ), and
= 6 (1
(4.99)
) 1
(
sin(
) +
aK (
) F ( K, 1+K; 1
K=1
in which aK
aK =
(2K + 1) (K + 1
)
3 F2 ( K, K + 1, 1; 1
2 (1
) (K + 1 + )
, 2; 1)
(4.101)
w
Uf
(4.102)
sin(
) 1
(
as
(4.103)
Sumer (1974) argues that, for the case of neutrally buoyant particles ( = 0),
1. Indeed the above equation reveals this.
C0 should go to unity, C0
122
w
Uf
4.3.3
123
mp =
(4.104)
Cp d
0
s+
)Cp
p(p
1)
Cp
d =0
= 0:
(4.105)
(4.106)
dm1
d
)C0 d = 0
(4.107)
=0
(4.108)
and
m1 (0) = 0 at
)C0 ( ,
)d =0
C0 ( ,
0
)d =
sin(
) 1
(
) d
(4.109)
This is the mean particle velocity relative to the mean ow velocity. Fig.
4.10 illustrates how the mean particle velocity varies with the fall velocity
parameter. Note that for
= 0, the case of neutrally-buoyant particles,
obviously this velocity must be equal to the mean ow velocity, meaning
that s
must be zero, as revealed by the gure.
124
Figure 4.10: Mean velocity of heavy particles relative to the mean ow velocity.
Sumer (1974).
4.3.4
The longitudinal dispersion coe cient can be written in the normalized form
as
D1
1 dm2
)
= (
(4.110)
hUf
6 2 d
(see Eqs. 4.87 and 4.104).
From Eq. 4.105,
1 dm2
=6
2 d
(1
0
)C0 d +
(
0
)C1 d
(4.111)
125
aK fK ( ) exp{ 6(K 2 + K) }
( )+
K=1
dK (
(4.112)
; ) exp{ 6(K 2 + K) }
) F ( K, 1 + K; 1
K=1
in which
= A (
sin(
) 1
(
)(
1
6 (1
d (4.113)
) d
as
.
in which < 1. It may be noted that C1
Now, from Eqs. 4.111, 4.113 and 4.103, the dispersion coe cient is obtained as
D1
hUf
(1
(
0
sin(
[(
36
)
s)
](
) d (4.114)
1
(1
[(
s)
](
) d
in which < 1.
For the case of neutrally buoyant particles/dispersant, putting
gets
1
D1
1
d
=
d
d
hUf
6
36 0
(1
)
0
0
= 0, one
(4.115)
The expression in Eq. 4.115 was rst obtained by Elder (1959). The integral
can be calculated numerically. The result is (taking , the von Karman
constant, 0.41)
D1
= 0.07 + 5.86 = 5.93
(4.116)
hUf
As seen from Eq. 4.115, the dispersion coe cient consists of two terms; the
rst term on the right hand side of the equation, 6 , (the numerical value of
which is 0.07, Eq. 4.116) represents the e ect of the turbulent di usion in
the streamwise direction; it stems from the rst term on the right hand side
of Eq. 4.83. The second term is actually the main portion of the longitudinal
126
Figure 4.11: Ratio of the longitudinal dispersion coe cient of heavy particles to
that of neutrally-buoyant particles. Sumer (1974).
dispersion coe cient, and induced by the combined action of the turbulent
di usion across the depth y and the variation of the ow velocity in the y
direction.
Returning to the case of heavy particles, the dispersion coe cient is plotted in Fig. 4.11. The gure shows that the dispersion coe cient increases
with increasing , the fall velocity parameter. This can be explained as follows. A heavy particle spends most of the time near the bottom where the
velocity gradient is largest. On the other hand, the larger the velocity gradient, the larger the dispersion coe cient. Therefore, a heavier particle (i.e.,
with larger ) should have a larger dispersion coe cient.
4.4
In the previous section, the longitudinal dispersion coe cient has been calcuD1
lated for an open-channel ow, and found to be hU
6. Field experiments in
f
127
128
the transverse velocity prole, i.e., the velocity prole across the river width.
(Fig. 4.12 displays an example illustrating the transverse velocity variation in
a river, the velocity isolines at a measurement section in the Green-Duwamish
River in Washington, and the corresponding transverse velocity prole).
Fisher obtained the following expression for the longitudinal dispersion
coe cient
D1 =
1
A
(u(y)
0
1
dy
t d(y)
Um ) d(y) dy
0
(u(y)
Um ) d(y) dy
(4.117)
in which A is the cross-sectional area of the river, W width at the surface,
Um mean ow velocity, y the transverse distance measured from one bank
of the river (Fig. 4.12) (not to be confused with the vertical distance used
in the previous section), d(y) the depth (the depth is a function of y, Fig.
4.12), t is the turbulent di usion coe cient in the transverse direction (it
may or may not be a function of y), and u(y) is the depth-averaged velocity
dened by (Fig. 4.12) :
0
1
u(y) =
d(y)
(4.118)
u(y, z) dz
d(y)
D1
=
hUf
36
d
0
(1
d
0
1
h
(u(y)
Um )dy
1
dy
(y)
(u(y)
Um )dy
or alternatively
D1
=
hUf
1
h2
(u(y)
0
Um )hdy
0
1
dy
(y)h
(u(y)
Um )hdy
(4.119)
4.5. REFERENCES
129
4.5
References
130
9. Pedersen, F.B. (1977): Prediction of Longitudinal Dispersion in Natural Streams. Series Paper No. 14, Technical University of Denmark,
Institute of Hydrodyn. and Hyd. Eng. (ISVA).
10. Sayre, W.W. (1968): Hydraulic Papers, No. 3, Colorado State University, Fort Collins.
11. Sumer, B.M. (1974): Mean velocity and longitudinal dispersion of
heavy particles in turbulent open-channel ow. J. Fluid Mech., vol.
65, p.11.
12. Taylor, G.I. (1921): Di usion by continuous movements. Proc. London Math. Soc., vol. 20, p. 196.
13. Taylor, G.I. (1953): Dispersion of soluble matter in solvent owing
slowly through a tube. Proc. Roy. Soc. London, A, vol. CCXIX,
p.186.
14. Taylor, G.I. (1954): The dispersion of matter in turbulent ow through
a pipe. Proc. Roy. Soc. London, vol. A223, p.446.
Chapter 5
Wave boundary layers
In shallow waters, the orbital motion of water particles under a progressive
wave degenerates into a bottom-parallel, straight line, oscillatory motion at
the seabed (Fig. 5.1). For each half cycle of this oscillatory motion at the
bed, a new time-dependent boundary layer develops. This cyclic boundary
layer is called the wave boundary layer.
The wave boundary layers are of great importance for several processes:
Turbulent mixing of mass and momentum takes place in the wave boundary layers; Likewise, frictional dissipation takes place in the wave boundary
layers; The wave boundary layers are one of the key components of the sediment transport processes; Exchanges of chemicals and organisms between
seabed and the main body of the water take place in the wave boundary layers. Therefore, understanding of wave-boundary layers is essential to develop
predictive tools and capabilities for the previously mentioned processes.
Fredse and Deigaard (1992, Chapters 2 and 3) can be consulted for a
detailed treatment of the subject including the mathematical modelling.
The ow in the wave boundary layers may be in the laminar regime, it
may be in the transitional, or it may be in the turbulent regime.
We will start o with the laminar-regime wave boundary layers, and then
we will focus on the turbulent wave boundary layers. The latter will be
mainly based on the work of Jensen et al. (1989), an extensive experimental work conducted in an oscillating water tunnel, covering both smoothand rough-boundary ows, with very broad range of the Reynolds number
including very high Reynolds numbers.
131
132
5.1
5.1.1
The streamwise component of the N.-S. equation (Eqs. 1.8 and 1.9) in the
boundary layer (Fig. 5.2):
u
u
u
+u
+v
=
t
x
y
1 p 1
+
x
y
(5.1)
First of all, v in the boundary layer is practically zero, therefore, the third
term on the left hand side of the equation drops out. Secondly, the x distances
are large compared to the y distances (the wave length in the laboratory
L = O(2-3 m) while the boundary layer thickness = O(1 mm) O(1 cm);
the same is also true for the turbulent wave boundary layers in the sea,
L = O(100 m) while the boundary layer thickness = O(10 cm)); therefore,
the x dependence in the above equation can be neglected, hence the second
term on the right hand side of the equation also drops out. Therefore, Eq.
5.1:
u
1 p 1
=
+
(5.2)
t
x
y
Now, the boundary layer is actually driven by the pressure gradient xp .
This term can be calculated from the equation of motion outside the boundary layer (in the free-stream region). This equation, by setting = 0 in the
133
Figure 5.2: The boundary-layer thickness, , is small compared with the wave
length, L.
above equation,
U0
=
t
1 p
x
(5.3)
in which U0 is given as
(5.4)
U0 = U0m sin( t)
in which is the angular frequency of the wave ( =
period.
From Eqs. 5.2 and 5.3, one gets
u
U0 1
=
+
t
t
y
in which
2
T
(5.5)
u
y
(5.6)
u
y2
(5.7)
U0m sin( t)
(5.8)
134
and
y = 0:
(5.9)
u=0
U0m exp(
y
1
in which
1
1
=(
)2
) sin( t
(5.10)
(5.11)
representing the thickness of the boundary layer, called the Stokes length.
The time development of the velocity distribution is plotted from Eq.
5.10 in Fig. 5.3. In the gure, is the boundary layer thickness, slightly
di erent from that in Eq. 5.11; namely, is the distance from the bed where
the velocity becomes maximum at the phase value t = 900 ; this distance is
found from the above solution as
=
3 2 1
( )2
4
(5.12)
3
2 1
( )2
4 Re
in which
Re =
aU0m
(5.13)
(5.14)
135
Figure 5.3: Time development of velocity proles in laminar wave boundary layer.
respond to the pressure gradient earlier, therefore will reverse earlier than the
outer ow, as revealed by the velocity proles in Fig. 5.3.
It may be noted that the preceding description may also explain the
overshooting exhibited by the plotted velocity proles in Fig. 5.3.
5.1.2
Flow resistance
= (
u
U0m
[sin( t) + cos( t)]
)y=0 =
y
1
(5.16)
or
2 U0m
)]
(5.17)
4
This equation shows that the bed shear stress leads over the free-stream ow
with a phase di erence (cf. Eq. 5.4)
0
[sin( t +
= 450
This also implies that the ow at the bed reverses
free stream.
(5.18)
= 450 earlier than the
136
Figure 5.4: Two di erent pressure gradients over one half cycle.
Finally, the friction coe cient in wave boundary layers is dened by
fw =
0m /
2
U0m
(5.19)
or
fw
U0m
(5.20)
2
in which fw is the wave friction coe cient (cf. Eq. 2.128), and Uf m is the
maximum value of the friction velocity dened by
Uf m =
Uf m =
0m
(5.21)
Here 0m is the maximum value of the bed shear stress. From Eqs. 5.16 and
5.20, fw
2
fw =
(5.22)
Re
5.2
Laminar-to-turbulent transition
137
The gure also includes, for convenience, the phase di erence (lead) between
the bed shear stress and the free-stream velocity (Fig. 5.5 b) plotted versus
the Reynolds number.
Figure 5.5: (a): Wave friction coe cient, and (b): phase di erence. Jensen et al.
(1989).
138
3. The phase lead decreases with the transition. This can be explained
as follows. In the transitional/turbulent ows, the momentum exchange
between the layers of uid will be enhanced. Therefore, the near-bed
ow will be supplied with the new, momentum-rich uid from the upper layers of the boundary layer, and as a result, it will dissipate its
momentum not as fast as in the case of the laminar ow. Hence, its
response to the adverse pressure gradient will be delayed, leading to
the decrease in the phase lead, as indicated in Fig. 5.5 b. In the fullydeveloped turbulent case (very large Re), the phase lead is only about
50 .
Figure 5.6: Normalized friction coe cient at di erent phase values. Jensen et al.
(1989).
139
Note that the bed shear stress above is not the maximum value of the shear
stress, but rather the phase-resolved bed shear stress, 0 = f ( t). The advantage of this representation is that, when divided by cos( t 4 ), the bed
shear stress will be independent of the phase t in the case of the laminar
ow (from Eq. 5.17):
2
U0m
=
)
(5.24)
0
1 sin( t +
4
Re 2
but will depend only on Re:
2
(5.25)
fw =
1
Re 2
which represents a straight line when plotted on a log-log paper.
Now, this enables us to identify the ow regime in the following way.
1. Measure the bed shear stress,
= f ( t).
t as a
140
5.3
5.3.1
Fig. 5.7 displays time series of the horizontal and vertical velocity components of the velocity in the boundary layer (the bottom two time series). The
top time series is the free-stream velocity. This is from an experiment with
Re = 6 106 .
The so-called ensemble averaging is used to calculate the mean values:
u(y, t) =
1
N
u[y,
(t + (j
1)T )]
(5.26)
j=1
141
Figure 5.8: Mean velocity distributions. Smooth bed. Re = 6106 . Jensen et al.
(1989).
velocity u = u
u 2 (y, t) =
5.3.2
u,
1
2
1
1
[u[y,
(t + (j
1)T )]
u(y, t)]2
(5.27)
j=1
Mean ow velocity
Fig. 5.8 shows the time development of the velocity proles for Re = 6 106 .
The way in which the velocity proles evolve with time is similar to that in
the case of the laminar wave boundary layer (Fig. 5.3). Features such as the
early reversal of the ow near the bed, and the overshooting in the velocity
proles are observed here, too.
The velocity proles in the previous gure are plotted in wall units on
semi-logarithmic plots in Fig. 5.9. The gure also includes the van Driest
prole (Eq. 2.97, the velocity prole obtained for steady boundary layers) as
a reference line in each panel. Note that Uf used in the normalized quantities
yU
y + = f and Uuf is the temporal value (the phase resolved value) of the
friction velocity. Recall that the van Driest prole tends to the logarithmic
142
(5.28)
143
Finally, note the following. Fig. 5.10 shows the semi-log plots of the
velocity distributions obtained in another research (Sumer et al., 1987), while
Fig. 5.11 shows the corresponding shear stress distributions. As seen from
Fig. 5.11, the constant stress layer is displaced to higher elevations in the
later phases (starting from t = 912 ) of the deceleration stage. This means
that the logarithmic layer should also be displaced to these higher elevations.
The velocity distributions may still be represented by the logarithmic law.
However, Uf in such a representation cannot be the wall shear velocity, but
1
a velocity dened by ( ) 2 where is the shear stress occurring at these
elevations rather than at the wall.
144
5.3.3
Flow resistance
The friction coe cient, dened by Eq. 5.19 or Eq. 5.20, has been determined from experiments in oscillatory ows and plotted in Fig. 5.5 for both
transitional and fully developed turbulent oscillatory boundary layers.
The variation of the friction coe cient with Re in Fig. 5.5 in the fully
developed turbulent regime (Re
106 ) can be approximated by (Fredse
and Deigaard, 1992, p. 29)
0.035
fw =
(5.29)
Re 0.16
5.3.4
Turbulence quantities
v2
Uf
w2
,
Uf
uv
Uf2
(5.30)
versus y + . Also plotted in the gures are the corresponding steady boundarylayer distributions (see, for example, Fig. 2.7).
Except at very early phases of the acceleration stage, and also late phases
of the deceleration stage, the oscillatory-ow proles appear to be in reasonable accord with the steady boundary-layer distributions throughout the
phase space. See Jensen et al. (1989) for further details.
5.3.5
Fig. 5.14 gives the mean velocity proles for di erent Reynolds number.
The characteristic features of the time development of the velocity proles
shown in the preceding gure are as follows.
1. The velocity prole eventually reaches a state where the logarithmic
layer comes into existence.
2. This logarithmic layer grows in thickness, as one proceeds further in
the phase space t.
3. This continues to be the case until practically the point of reversal near
the bed is reached.
145
et al. (1989).
146
Figure 5.13: Reynolds stresses. Smooth bed. Re = 6 106 . Jensen et al. (1989).
From the gure, it is seen that the higher Re, the earlier the logarithmic
layer comes into existence. This can be explained by reference to Fig. 5.6;
From the latter gure, it is seen that, for the establishment of the logarithmic
layer, the ow has to reach the fully developed turbulent state.
Jensen et al. (1989) also give the e ect of Re on the turbulence quantities.
5.3.6
3
2 1
( )2
4 Re
(5.31)
From the gure, the boundary-layer thickness for storm conditions (very
large Re)
a
(5.32)
= O(0.01)
= O(10 cm).
147
148
Figure 5.16: Mean velocity u. Circles: Smooth bed. Triangles: Rough bed. Re
= 6 106 . Jensen et al. (1989).
5.3.7
Figs. 5.16-5.19 compare the rough-bed results with the corresponding smoothbed ones in the tests of Jensen et al. (1989). Re = 6 106 and ks+ = 84 in
k U
which ks+ = s f m where ks is Nikuradses equivalent sand roughness. The
parameter kas , on the other hand, is 3700.
From these gures, the following points appear to be noteworthy.
1. The description given in the preceding paragraphs of how the ow and
turbulence evolve during one half-cycle is qualitatively the same for the
rough-bed ows, too.
2. The introduction of the roughness, however, results in a decrease in the
149
Figure 5.17: R.m.s. value of u . Circles: Smooth bed. Triangles: Rough bed. Re
= 6 106 . Jensen et al. (1989).
Figure 5.18: R.m.s. value of v . Circles: Smooth bed. Triangles: Rough bed. Re
= 6 106 . Jensen et al. (1989).
150
Figure 5.19: Reynolds stress u v . Circles: Smooth bed. Triangles: Rough bed.
Re = 6 106 . Jensen et al. (1989).
mean ow velocity and an increase in the turbulence quantities.
3. However, this e ect tends to disappear with the distance from the bed;
the mean ow velocity is practically the same for the two cases for
distances ay
0.01. The same is also true for the turbulence quantities, although this time the e ect of roughness disappears at somewhat
larger distances from the bed, namely for ya 0.03. The decrease in the
mean velocity is due to the retarding e ect of the rough boundary while
the increase in the turbulence quantities is due to the enhancement of
the momentum transfer by the presence of the roughness elements..
4. The ow begins to feel the e ect of the roughness only after t reaches
the value of about 150 . This is because the temporal value of the roughness Reynolds number is still far too small for these phases for the
bed to act as a completely rough boundary and, as a result, the ow
quantities for this short interval of time remain practically unchanged,
regardless of the boundary category.
The rough turbulent boundary layers are governed solely by the parameter
This parameter plays a similar role to that of Re in the case of the smooth
turbulent wave boundary layers. Jensen et al. (1989) discussed the e ect of
this parameter on the mean and turbulence ow quantities.
The friction coe cient in the case of the rough bed can be represented
by the following empirical expressions (Fredse and Deigaard, 1992, p. 25
a
.
ks
151
a 1
a
) 4
when
> 50
(5.33)
ks
ks
a 3
a
fw = 0.4( ) 4
when
< 50
(5.34)
ks
ks
The latter expression is due to Kamphuis (1975) (Fredse and Deigaard,
1992, p. 26).
In the case of the transitional-category-bed wave boundary layers, the friction coe cient can be obtained from Kamphuis (1975) empirical diagrams.
Also consult Fredse and Deigaard (1992, p. 33).
Finally, Fig. 5.20 depicts the boundary-layer thickness. The gure shows
that, the boundary-layer thickness for storm conditions (very large kas )
fw = 0.04(
5.4
(5.35)
= O(0.02)
= O(20 cm).
Other aspects
Several other aspects of wave boundary layers have been studied in recent
years. The following is a partial list of these studies undertaken at ISVA.
Fredse, Sumer, Laursen and Pedersen (1993) have studied the nonuniform wave boundary layers, where the non-uniformity is due to a sudden
change in the bed roughness, while Sumer, Laursen and Fredse (1993) have
152
5.5
References
5.5. REFERENCES
153
6. Fredse, J., Sumer, B.M., Laursen, T. S.and Pedersen, C. (1993): Experimental investigations of wave boundary layers with a sudden change
in roughness. Journal of Fluid Mechanics, Vol. 252, pp.117-145.
7. Fredse, J., Sumer, B.M., Kozakkiewicz, A., Chua, L. H. C. and Deigaard,
R. (2003): E ect of externally generated turbulence on wave boundary
layer. Coastal Engineering, vol. 49, pp. 155-183.
8. Jensen, B. L., Sumer, B. M. and Fredse, J.: Turbulent oscillatory
boundary layers at high Reynolds numbers. Journal of Fluid Mechanics, Vol. 206, p. 265, 1989.
9. Kamphuis, J. W. (1975): Friction factor under oscillatory waves.
J. Waterways, Port, Coastal and Ocean Eng. Div., ASCE, Vol. 101
(WW2), pp. 135-144.
10. Lodahl, C. R., Sumer, B. M. and Fredse, J. (1998): Turbulent combined oscillatory ow and current in a pipe, J. Fluid Mech., vol. 373,
313-348.
11. Sleath, J. F. A. (1984): Sea Bed Mechanics, John Wiley & Sons.
12. Sumer, B. M., Jensen, B. L. and Fredse, J: Turbulence in oscillatory boundary layers. Advances in Turbulence. Proc. 1st European
Conference, Lyon, France, published by Springer, p. 556, 1987.
13. Sumer, B. M., Laursen, T. S. and Fredse, J. (1993): Wave boundary
layers in a convergent tunnel. Coastal Engineering, Vol. 20, pp. 317342.
14. Tanaka, H., Sumer, B. M. and Lodahl, C. (1998): Theoretical and
experimental investigation on laminar boundary layers under cnoidal
wave motion, Coastal Engineering Journal, vol. 40, No. 1, 81-98.
154
Chapter 6
Turbulence modelling
6.1
ui
ui
+ uj
) = gi +
(
t
xj
xj
ij ,
(6.1)
ij
ui uj )
(6.2)
ij
+ (
ui
uj
+
)
xj
xi
(6.3)
u1 u1
u2 u1
u3 u1
u1 u2
u2 u2
u3 u2
u1 u3
u2 u3
u3 u3
(6.4)
in which there are, due to symmetry, six independent Reynolds stresses (three
diagonal components and three o -diagonal ones).
As has already been discussed (Chapter 1), there are four equations for
the mean ow, i.e., the continuity equation (Eq. 1.6) and the three Reynolds
155
156
equations (Eq. 1.14) whereas there are ten unknowns (namely, three components of the velocity, ui , and the pressure, p, and six components of the
Reynolds stress,
ui uj ), hence the system is not closed. This problem is
known as the closure problem of turbulence.
To close the system, we must nd enough number of equations to solve
for the unknowns. We do this through turbulence modelling.
(In the case of laminar ow, there are four unknowns (namely, three
components of the velocity, ui , and the pressure, p), and four equations, i.e.,
the continuity equation (Eq. 1.4) and three N.-S. equation (Eqs. 1.8 and
1.9). Therefore, the system is closed.)
6.2
Turbulence models
There are several kinds of turbulence models the most important of which
are
1. Algebraic models;
2. Turbulence-energy equation models; and
3. Simulation models such as Direct Numerical Simulation (DNS), Large
Eddy Simulation (LES) and Detached Eddy Simulation (DES).
Algebraic models, also known as the zero-equation models, are the simplest of all turbulent models. Here, (1) the Reynolds stress is expressed as
the product of a turbulence viscosity (eddy viscosity) and the mean strain
rate, the so-called Boussinesq eddy-viscosity approximation, and (2) the turbulence viscosity, in turn, is often computed in terms of a mixing length. Of
the algebraic models, the mixing-length model is the simplest and the most
popular one. This model will be described in the next section.
Turbulence-energy equation models are advanced models. Here, too, the
Reynolds stress is expressed as the product of a turbulence viscosity (eddy
viscosity) and the mean strain rate. However, the turbulence viscosity is not
computed in terms of a mixing length, but rather in terms of the turbulence
kinetic energy, k = 12 (ui ui ). There are two kinds of turbulence-energy equation models: (1) One-equation models, and (2) Two-equations models. As an
example, the so-called k
model, a two-equations model, will be described
in the following paragraphs.
157
Simulation models are most advanced models. These models directly address the turbulence process itself, and therefore they do not involve the afore
mentioned closure problem. In a DNS study, a complete time-dependent solution of the Navier-Stokes equations and the continuity equation is sought.
Therefore the solution furnishes the time series of (1) the three components
of the velocity and (2) the pressure in 3-D space. Up until now, only low
Reynolds number-ows are calculated with the DNS method. For large
Reynolds numbers, the scales of the dissipative part of turbulence motion
are so small that this kind of small scale motion can not be resolved in a
numerical calculation (the number of grid points required to resolve this motion increases approximately with Re3 , Rodi, 1992). However, the upper
limit of the Reynolds number range is pushed to higher and higher values as
more and more computer power becomes available. Rogallo and Moin (1984)
gives a review of the DNS method where the principal/early publications are
cited. Regarding the LES model, here the large eddies (large-scale motions)
are computed while the smallest eddies are modelled. The details of the LES
model along with an application example are given in the chapter.
Wilcoxs (1994) book can be consulted for a detailed account of turbulence
models.
6.3
Mixing-length model
6.3.1
158
Figure 6.2: Travel of uid element A from one side of the area (dx,dz) considered
in the previous gure to the other side.
regime.)
The mass of the uid transported from one side of the unit area to the
other is
(1 1 dy)
and therefore the momentum transported in the y direction from one side
of the unit area to the other (x component):
x
or, inserting u = u + u
x
159
Fx =
d(dy)
dt
Fx = v (u + u )
and by time averaging
Fx = v (u + u ) = v u + u v ) = 0 + u v = u v
(6.5)
(6.6)
These forces, Fx and Fy , are exerted on the unit area indicated in Fig. 6.1,
and therefore they are, respectively, the shear and normal stresses. (Notice
that these stresses are induced by turbulence (Fig. 6.3.)
So the above
analysis demonstrates that
1. The turbulent transport of momentum in the y direction generates
stresses in the uid;
160
xy
xy
yy
u2
uv
uv
v2
(6.7)
6.3.2
Mixing length
161
m,
from Level I to
uv;
3. Newtons second law implies that the change in the momentum over
the time from I to II should be equal to the shear stress
u v , i.e.,
u v = Momentum II
Momentum I
(6.8)
Momentum I = ( q)(
du
)
dy
(6.10)
162
Figure 6.6: Volume associated with Fluid Parcel A passing through per unit area.
in which q is the volume associated with the uid parcel (Fig. 6.5) passing
through per unit area per unit time (Fig. 6.6). The quantity q is
q = Vt (1 1) = Vt
and therefore the gained momentum (x component) will be
Momentum II
Momentum I = Vt
du
dy
(6.11)
Momentum I =
du
dy
(6.12)
(6.13)
du
dy
(6.14)
163
the turbulence shear stress can be written as (as was rst suggested by Boussinesq in 1877)
du
uv = t
(6.15)
dy
in which t is called turbulence viscosity (or eddy viscosity). Hence, from
Eqs. 6.13 and 6.15, the turbulence viscosity is
t
m Vt
(6.16)
164
(6.17)
while the amount of uid entering the control volume in the horizontal direction per unit time at the instant I:
uI
(6.18)
and that leaving the control volume in the horizontal direction per unit time
at the instant II:
uII m
(6.19)
Continuity requires:
Vt dx + uI
= uII
and therefore
Vt = (uII
and since O(
m)
uI )
dx
uI
du
dy
165
2
m
du
dy
(6.21)
yUf
70
166
Now, we should expect that the mixing length (for y values in the range
y
h and y + > 30 70) increases with increasing distance from the wall,
and the simplest form of this dependence is
m
(6.22)
= y
in which
is a constant. From the solution for the velocity (Eq. 6.28),
turns out to be the von Karman constant ( = A1 = 0.4) introduced in
Chapter 2. Eq. 6.22 is valid for the ranges 30 70 < y + , and y
h.
van Driest (1956) revised Eq. 6.22 so as to accommodate the inuence of
viscosity in the transport of momentum. This led to
m
= y[1
exp(
y+
)]
Ad
(6.23)
in which Ad is called the damping coe cient, and taken as 25, and this
equation for the mixing length is valid for the range 0 y
h. From Eq.
6.23, it is seen that
y, for large values of y +
(6.24)
while
0, as y +
(6.25)
= y[1
2
m
du
dy
du
dy
exp(
y+
)]
Ad
167
So, for the problem we consider, i.e., the steady boundary-layer ow close
to a smooth wall (in the constant stress layer, see Section 2.1), there are three
model equations (the preceding equations), and one Reynolds equation, Eq.
2.87, i.e., altogether four equations, and four unknowns, namely, u,
uv,
t , and m . Hence, the system is closed. Therefore the unknowns can be
determined. Of particular interest is the velocity, u. The solution for the
latter quantity is obtained in Section 2.3.1 (Eq. 2.97):
y+
dy +
u = 2Uf
0
1+ 1+4
2 y +2 [1
exp(
y+ 2
)]
Ad
1/2
(6.26)
(the van Driest velocity distribution, Fig. 2.20). This equation is valid in
the constant stress layer, namely for the range 0 y
h.)
+
For small values of y , the van Driest prole reduces to the linear velocity
distribution in the viscous sublayer (Eq. 2.20),
u
= y+
Uf
(6.27)
(6.28)
It is seen from Fig. 2.20 that the van Driest prole agrees remarkably well
with the experiment.
6) Mixing length in the case of transitional and rough walls and solution
In this case, the mixing length may be given as (Cebeci and Chang, 1978)
m
= (y +
y)[1
exp(
(y + + y + )
)]
Ad
(6.29)
y in the above equation is measured from the level where the velocity u is
zero. This level lies almost at the top of the roughness elements, as discussed
in Section 2.3.1 (y here should not be confused with the distance measured
from the theoretical wall. y+ y is the distance measured from the theoretical
wall. See the detailed discussion in Section 2.3.1).
168
ks+ exp(
ks+
)];
6
(6.30)
in which ks+ is the roughness Reynolds number (ks being Nikuradses equivalent sand roughness):
ks Uf
ks+ =
(6.31)
Here the distance from the wall y should be in the interval 0 y
h where
the upper boundary ensures that the y levels lie in the constant stress layer.
The solution for u for the set of the model equations along with the
Reynolds equation is obtained in Section 2.3.1 (Eq. 2.103):
y+
dy +
u = 2Uf
0
1+ 1+4
2 (y +
y + )2 [1
exp(
(y+ + y + ) 2
)]
Ad
1/2
(6.32)
Fig. 2.21 displays the velocity proles obtained from the preceding equation
including the smooth-wall prole. In Fig. 2.21, a represents the linear prole
in the viscous sublayer (Eq. 6.27), b the logarithmic prole (Eq. 6.28) in the
case of the hydraulically smooth wall, and c the logarithmic prole in the
case of the completely rough wall (Eq. 2.65), namely, by putting z = y + y,
u = AUf ln(
30z
)
ks
(6.33)
6.4
169
k-omega model
This model is a turbulence-energy equation model. As mentioned in the preceding paragraphs, there are two kinds of turbulence-energy equation model,
one-equation model and two-equations model. The k
model is a twoequations model.
There are three versions of the k
model in the literature: (1) the original k
model, which is due to Wilcox (1994), (2) the k
, BSL (baseline)
model, and (3) the k
, SST (shear-stress transport) model. The latter two
models were developed by Menter (1993) to improve Wilcoxs original model
so that an even higher sensitivity could be obtained for adverse-pressuregradient ows. In this section, we will describe the k
, SST model.
Incidentally, the following note is worth mentioning. Menter (1993) made
an extensive comparison between (1) the classic k
model; (2) the original
k
model; (3) the k
, BSL model; and (4) the k
, SST model for
various well documented ows. The tested ows were, among others, two
kinds of adverse pressure gradient ow (one having a very strong adverse
pressure gradient, so strong that separation occurs); the backward-facingstep ow; and the ow past a NACA 4412 airfoil at an angle of attack near
maximum lift condition. The latter two ows also have substantial adverse
pressure gradient e ects. The main conclusion from this inter-comparison
exercise was that the k
, SST model gave the most accurate results while
the k
model did not yield as accurate results as the other three for the
tested adverse-pressure-gradient ow cases.
6.4.1
Model equations
uj k
xj
xj
( +
k t)
k
=
xj
ij
ui
xj
(6.34)
(It should be noted that this equation is in a form slightly di erent from that
170
uj
xj
( +
xj
+2 (1
t)
F1 )
xj
ij
t
ui
xj
k
xj xj
(6.35)
Above, k is:
1
k = ui ui
2
(6.36)
and :
=
in which
(6.37)
=
In the preceding equations,
ij
ij
(6.38)
ui uj =
2
k
3
ui
uj
+
xj
xi
ij
ui uj ),
(6.39)
= F1
k1
+ (1
F1 )
k2
(6.40)
= F1
+ (1
F1 )
(6.41)
171
= F1
+ (1
F1 )
(6.42)
= F1
+ (1
F1 )
(6.43)
in which k1 , k2 ,
1,
2,
1, 2,
1 and
2 , the model constants, are
combined as a function of the distance z to the nearest wall with the socalled blending function , F1 .
6.4.2
Model constants
0.567
6.4.3
0.85
0.5
0.463
1.0
0.856
Blending function F1
(6.44)
and
k 500
;
0.09 z z 2
4 k 2
(CDk )z 2
(6.45)
with
(CDk ) = max 2
k
; 10
xj xj
20
(6.46)
172
6.4.4
In the k
0.3k
max (0.3 ; F2 )
(6.47)
where is the absolute value of the vorticity. The function F2 in Eq. 6.47
is another blending function, dened by
F2 = tanh arg22
(6.48)
in which
arg2 = max 2
k 500
;
0.09 z z 2
(6.49)
Both F1 and F2 take the value unity at the wall and gradually decrease and
approach to zero as the distance from the wall is increased.
6.4.5
Boundary conditions
The boundaries of the computational domain (Fig. 6.8) are (1) Inlet, (2)
Outlet, (3) Symmetry Boundaries and (4) Walls (e.g., the bed in the example
173
given in Fig. 6.8 and the surface of a rigid body present in the computational
domain, if any).
1) Inlet and outlet boundary conditions
At the inlet, zero transverse, and vertical velocities are specied. At
the outlet, zero-gradient conditions ( / n = 0), Neumann conditions, are
applied for all quantities.
2) Symmetry boundaries
At the symmetry boundaries , Neumann conditions are applied for k and
and for the three components of the velocity (u, v and w). This is (1) at
the sides and (2) at the top surface of the computational domain where a
lid is placed in the case when the model cannot handle the free surface,
Fig. 6.8).
3) Walls (e.g., the bed and/or the surface of a rigid body)
1. Zero velocity is specied for u , v and w.
2. Zero turbulent energy is specied for k when the wall is smooth, while
the Neumann condition is applied when the wall is rough and transitional.
Regarding the latter condition, experiments reveal that the Neumann condition is satised ( k/ n = 0) at the wall for rough boundaries, n being the
direction normal to the wall (Nezu, 1977, Sumer et al., 2001 and 2003). A
validation exercise for the case of a steady boundary layer in an open channel ow revealed a much better agreement with the experiments (Nezu, 1977,
and Sumer et al., 2001) when the boundary condition was taken in the form
of the Neumann condition (Roulund, 2000, p. 25).
3. The Dirichlet condition is applied for , namely
= Sr
Uf2
at the wall
(6.50)
where Uf =
o / is the friction velocity based on the wall shear stress o .
The quantity Sr is a tuning parameter, and it is used to account for the bed
roughness:
Sr =
3
40
+
ks
100
0.85
(ks+ )
(6.51)
174
where ks+ = (ks Uf )/ is the wall roughness in wall units, and ks is Nikuradses
equivalent sand roughness. Wilcox (1994) was the rst to introduce the
boundary condition in Eq. 6.50. The tuning parameter Sr was given by
Wilcox (1994) as in Eq. 6.51, but with constants slightly di erent from
those in Eq. 6.51 due to the special implementation of the velocity boundary
conditions. With the present constants (Eq. 6.51), the friction velocity
determined for a wide range of velocities for the case of the undisturbed ow
was found to agree rather well with that obtained from the classic Colebrook
and White equation (the di erence being less than 6%, depending on the
wall roughness, Roulund, 2000, p. 23).
4. For the implementation of the preceding boundary condition, the wall
shear stress, 0 (= Uf2 ), is needed. Traditionally, 0 is calculated from 0 =
u
( + t )( xuji + xji )W all . However, it may also be calculated from the van
Driest (1956) velocity prole (Eq. 6.32):
u
=2
Uf
y+
0
dy +
1+
1+4
(y + +
y + )2
exp
y+ +
y+
(6.52)
This improves the stability of the model and allows a coarser mesh resolution
at the bed, as was revealed by the work of Roulund et al. (2005). In the
implementation of Eq. 6.52, u is taken as the tangential velocity in the cell
adjacent to the wall.
Similar implementation of the bed shear stress in the wall boundary conditions is reported in Srensen (1995) for the k
model.
As a nal remark in conjunction with the previous description, it may be
emphasized that the k
model is a well tested model and has been used
successfully in various contents in computational uid dynamics (Wilcox,
1994, Menter, 1993). The only change made in the present implementation
is the values of the constants in Eq. 6.51 which are slightly di erent from
those originally given by Wilcox (1994) (item 3 above), plus the way in which
the wall shear stress is calculated is through the van Driest equation (item 4
above).
6.4.6
Numerical computation
175
6.4.7
An application example
Under this subsection, we will present the results of a small study where the
k
model was employed.
The CFD code used in this study, EllipSys3D, is a three-dimensional
general purpose ow solver (an incompressible general purpose Navier-Stokes
solver). It is a multiblock nite-volume numerical model that solves the
incompressible Reynolds equations (Eqs. 6.2-6.2). The k
, SST model is
used as the turbulence closure. EllipSys3D has been developed at the Ris
National Laboratory, Denmark and at the Technical University of Denmark,
Department of Fluid Mechanics. The code o ers a variety of turbulencemodel options, including the k
, SST model. The basic principles of the
model are described in Michelsen (1992) and Srensen (1995).
176
177
178
Figure 6.11: Turbulence kinetic energy versus distance from the wall. Smooth
wall.
Figure 6.12: Turbulence kinetik energy versus distance from the wall, in terms of
inner ow parameters. Smooth wall.
179
180
6.5
6.5.1
181
The principal idea in the method of LES is to average the N.-S. equations over
a small volume of space. (Recall that the N.-S. equations are averaged over
time to obtain the Reynolds equations, i.e., Eqs. 6.1 and 6.2). Normally, the
size of the averaging volume is chosen such that the small-scale component
of turbulence is, for convenience, not resolved. The aforementioned spaceaveraging gives the following set of equations: (1) the continuity equation,
ui
=0
xi
(6.53)
1 p
+
xi
xj
ui
uj
+
)
xj
xi
ui uj
xj
(6.54)
1 p
( +
+
xi
xj
t )(
ui
uj
+
)
xj
xi
(6.56)
(It may be noted that the turbulence viscosity t here is di erent from that
in the k
model dened in the previous section. As emphasized in the
182
= Cs2
(6.57)
(6.58)
(6.59)
1 ui
uj
Sij = (
+
)
2 xj
xi
(6.60)
in which Sij
ui
uj ui
+
)
xj
xi xj
1/2
(6.61)
183
6.5.2
An application example
In this subsection, we will present the results of a study where the LES model
was used to simulate the wave boundary layer in a U-shaped oscillating tunnel
(Fig. 6.17); The study is essentially a part of a Ph.D. work conducted at the
Technical University of Denmark, Lohmann et al. (2006). The CFD code
employed in the study, called NS3, a 3-D numerical ow solver, was made
available by DHI Water and Environment.
The Smagorinski subgrid-scale model was used in the calculations. Regarding the value chosen for the Smagorinski constant,Cs , experience shows
that Cs equals 0.2 in isotropic turbulence and 0.1 in channel ow. In Lohmann
et al.s (2006) study, the latter value was found to be most appropriate, based
on comparison with experimental data. The van Driest damping function
y+
f = 1 exp( A
) (see Eq. 6.23) is generally used as a standard part of
d
the Smagorinski subgrid-scale formation; it was also used in the study of
Lohmann et al. (2006) (see p. 3 of the latter publication).
Note that the free-stream velocity varies according to
U0 = U0m sin( t)
184
1
N
1
2
[u[y,
(t + (j
1)T )]
u(y, t)]2
(6.63)
j=1
Lohmann et al. (2006) found that, for the chosen phase values, averaging
185
186
Figure 6.19: Sequence of velocity proles over half period of the motion.
Figure 6.20: Sequence of velocity proles over half period of the motion. Semi-log
plot.
187
188
6.6. REFERENCES
189
turbulence fairly well. Yet, one must note that the model near the bed fails
to some extent to predict the turbulence accurately. This may be a major
shortcoming when processes (to be studied by the model) have signicant
elements related to the turbulence near the bed, particularly related to the
y and z components of turbulence.
6.6
References
190
9. Michelsen, J. A. 1992 Basis3D - a platform for development of multiblock PDE solvers. Technical report, Dept. of Fluid Mechanics, Technical University of Denmark, AFM 92-05, ISSN 0590-8809.
10. Nezu, I. 1977 Turbulent Structure in Open Channel Flows. Ph.D.
thesis, Kyoto University, Japan.
11. Nielsen, P. and Teakle, A.L. (2004). Turbulent di usion of momentum
and suspended particles: A nite-mixing-length theory. Physics of
Fluids, vol. 16, Number 7, p. 2342.
12. Rodi, W. (1992): On the simulation of turbulent ows past blu bodies. J. Wind Eng., No. 52, August, pp.1-16.
13. Rogallo, R.S. and Moin, P. (1984): Numerical simulation of turbulent
ows. Annual Review of Fluid Mechanics, vol. 16, pp. 99-137.
14. Roulund, A. 2000 Three-dimensional modelling of ow around a bottommounted pile and its application to scour. Ph.D. thesis, Department
of Hydrodynamics and Water Resources, Technical University of Denmark, Supervisors B.M. Sumer, and J. Fredse.
15. Roulund, A., Sumer, B.M., Fredse, J. and Michelsen, J.: Numerical
and experimental investigation of ow and scour around a circular pile.
J. Fluid Mechanics, vol. 534, 351-401, 2005.
16. Srensen, N. N. 1995 General purpose ow solver applied to ow over
hills. Ph.D. thesis, Ris National Laboratory, Roskilde, Denmark, RisR-827(EN).
17. Sumer, B. M., Chua, L. H. C., Cheng N.-S. & Fredse, J. 2003 The
inuence of turbulence on bedload sediment transport, J. Hydraulic
Engineering ASCE, 129, No. 8, August Issue.
18. Sumer, B. M., Cokgor, S., & Fredse, J. 2001 Suction removal of sediment from between armour blocks. Journal of Hydraulic Engineering,
ASCE, 127, No. 4, pp. 293-306, 2001.
19. van Driest, E.R. (1956): On turbulent ow near a wall. J. Aeronautical Sciences, vol. 23, p.1007.
6.6. REFERENCES
191
20. Wilcox, D.C. (1994): Turbulence Modelling for CFD. DCW Industries,
Inc., La Canada, California, 460 p.
.