Está en la página 1de 20

Acta Geophysica

vol. 59, no. 5, Oct. 2011, pp. 993-1012


DOI: 10.2478/s11600-011-0025-1

Power-Law Velocity Profile


in Turbulent Boundary Layers:
An Integral Reynolds-Number Dependent Solution
Oscar CASTRO-ORGAZ1 and Subhasish DEY2
1

IAS-CSIC, National Research Council, Cordoba, Spain


e-mail: oscarcastro@ias.csic.es

Department of Civil Engineering, Indian Institute of Technology, Kharagpur,


West Bengal, India; e-mail: sdey@iitkgp.ac.in (corresponding author)

Abstract
Geophysical flows of practical interest encompass turbulent boundary
layer flows. The velocity profile in turbulent flows is generally described by
a log- or a power-law applicable to certain zones of the boundary layer, or
by wall-wake law for the entire zone of the boundary layer. In this study,
a novel theory is proposed from which the power-law velocity profile is obtained for the turbulent boundary layer flow. The new power-law profile is
based on the conservation of mass and the skin friction within the boundary layer. From the proposed theory, analytical expressions for the powerlaw velocity profile are presented, and their Reynolds-number dependency
is highlighted. The velocity profile, skin friction coefficient and boundary
layer thickness obtained from the proposed theory are validated by the reliable experimental data for zero-pressure gradient turbulent boundary layers.
The expressions for Reynolds shear stress and eddy viscosity distributions
across the boundary layer are also obtained and validated by the experimental data.
Key words: boundary layer, power-law, Reynolds stress, stream flows,
turbulence.

2011 Institute of Geophysics, Polish Academy of Sciences

Unauthenticated
Download Date | 2/22/15 5:23 PM

994

O. CASTRO-ORGAZ and S. DEY

1. INTRODUCTION
Several geophysical flows of practical relevance are boundary layer turbulent
flows, such as the atmospheric wind boundary layer or the free surface flow in
a river (Meyer 2009) and erosion gullies. Turbulent flows are required to be
described in terms of the velocity profile and flow resistance (Rowiski et al.
2005). For a velocity profile in turbulent flows, the flow conditions in the vicinity of the wall are described by the so-called logarithmic law of the wall (henceforth log-law for brevity). However, it has been extensively verified that the
log-law does not apply in the outer region of the boundary layer. For instance,
in free surface flows, the law of the wall holds only for 20 percent of the flow
depth from the wall (White 1991). Better agreement from the theoretical validity of the law of the wall could be achieved simply by changing the parameters
of the law of the wall while fitting the velocity profile across the entire boundary layer (Chen 1991). This method, however, suffers from a weak physical
basis. It has been recently shown that the constants of the law of the wall could
vary considerably from their classical values of = 0.41 and B = 5, where
is the von Krmn constant, and B is the constant of integration for the law
of the wall (sterlund 1999, sterlund et al. 2000, Koll 2006, George 2007).
Coles (1956) made an important advancement and argued that away from the
wall, the deviations of the profiles of measured velocity from those obtained
from the law of the wall could be explained by another universal law, called the
law of the wake. Coupling both the laws (wall and wake), a complete approximation to the time-averaged velocity profile in turbulent flows is then feasible
(White 1991, Krogstad et al. 1992, Guo et al. 2005). Note that joint wall and
wake laws are only feasible for the flow above the roughness elements, but not
for the flow within.
The deviation of the log-law from the measured velocity profiles indicates that the velocity profiles could be explained by power-laws, allowing for
a no-slip boundary condition. Extensive research has resulted in both physical and experimental support for the power-law (Barenblatt 1993, George and
Castillo 1997, Wosnik et al. 2000, Barenblatt 2000a, b), against the widely
used log-law model (Zagarola 1996, Zagarola et al. 1997, Zagarola and Smits
1998, Buschmann and Gad-el-Hak 2003a, b, Buschmann and Gad-el-Hak
2006). In a series of papers by Afzal (1996, 2001a, b, 2005, 2007, 2008)
and Afzal et al. (2006), it was shown that in power-law theory the outer layer is
also a wake-like structure, presenting composite power law-wake law velocity
profiles for boundary layer flows.
In general, the velocity profile in turbulent flows is often described by a
log- or a power-law in certain zones of the boundary layer or by wall-wake law
to describe the entire velocity profile. While the wall-wake law describes the
entire velocity profile in turbulent flows, there is a lack of generalization in the

Unauthenticated
Download Date | 2/22/15 5:23 PM

POWER-LAW VELOCITY PROFILE

995

power-law. The power-law is currently a topic of debate (Balachandar et al.


2002) between power- and log-law advocates. In this study, a novel power-law
velocity profile is developed based on the boundary layer theory considering
the results of the wall-wake law velocity profile. The profile obtained from the
new power-law is in good agreement with the measured velocity data within the
entire boundary layer. The proposed power-law produces the same skin friction
coefficient and the displacement thickness that are obtained from the log-wake
model. Thus, this profile simplifies the computation of boundary layers using
integral methods based on von Krmns integral equation (White 1991). New
generalized Reynolds-number dependent power law parameters derived from
the theory are presented here. The zero-pressure gradient turbulent boundary
layer (henceforth ZPGTBL for brevity) is one of the fundamental problems on
turbulent shear flows. It is important in mechanical, aeronautical and hydraulic
engineering (Rotta 1962, Hinze 1975, White 1991). Thus, the ZPGTBL is used
as a test case to validate the proposed power-law theory, yet it may be applied
for variable pressure gradients within the boundary layer. This may be achieved
by providing closure, as a function of the pressure gradient, to the strength of
the wake in the wall-wake model. This is also a parameter in the proposed
power-law.

2.

INTEGRAL POWER-LAW THEORY FOR


TURBULENT BOUNDARY LAYERS
The time-averaged velocity profile for the entire thickness of a turbulent boundary layer for smooth-bed flows can be expressed as a combination of logarithmic
and wake terms (Dean 1976)
1  yu 
u
= ln
+B+
u

1  yu 
= ln
+B+

 y 2 1
 y 3
1
(1 + 6)
(1 + 4)
,

(1)

where u is the time-averaged streamwise velocity component (parallel to the


wall) at a distance y , y is the vertical distance from the wall, u is the shear
velocity, that is, (0 /)0.5 , 0 is the boundary shear stress, is the mass density
of fluid, is the kinematic viscosity of fluid, is the boundary layer thickness,
W is the wake function, and is the wake parameter. In equation (1), the wake
function proposed by Dean (1976) and Granville (1976) was considered. The
velocity defect law for the wall-wake model is obtained from eq. (1) as

 y 2
 y 3 
U u
1 y 1
= ln
+
2 (1 + 6)
+ (1 + 4)
u

Unauthenticated
Download Date | 2/22/15 5:23 PM

(2)

996

O. CASTRO-ORGAZ and S. DEY

implying a complete similarity. The skin-friction coefficient Cf is related to the


shear stress by
0 =

1
Cf U 2 .
2

(3)

Other approximations, such as Coles (1956), White (1991) or Afzal (1996),


with other wake terms, can be considered. However, for the integral treatment
of boundary layer flow using the momentum equation, Deans approach is simple, satisfying the boundary conditions at the outer edge of the velocity profile,
that is, du/dy = 0. The analysis is then limited to this approach. The first two
terms of eq. (1) refer to the law of the wall, and the terms involving correspond to the law of the wake. Equation (1) accounts for a wake term that
is a result of a best fit of the experimental data over the entire boundary layer
thickness in the turbulent smooth flow regime. The simple polynomial for the
wake effect in eq. (1) satisfies the boundary conditions for the velocity profile,
including a zero-slope at the outer edge of the boundary layer. This condition is
not considered in the classic cosine law of the wake proposed by Coles (1956).
For the turbulent smooth flow regime, the law of the wall only agrees with the
experimental data up to y/ 0.2. For y/ > 0.2, the wake term in eq. (1) is
added to the law of the wall to match with the mean velocity profile in turbulent
flows. There is a strong debate concerning the mean velocity profile in turbulent boundary layers (George 2007, Buschmann and Gad-el-Hak 2003a, b).
Recently, Buschmann and Gad-el-Hak (2006) argued that it is not viable to
propose either a pure log-law or a power-law to describe the entire velocity
profile in a turbulent boundary layer. This indicates that within the boundary
layer, there are zones that can better be described by log-law equations, whereas
power-laws appear to fit reasonably well in the other zones. Then, even when
there exists a strong physical reasoning from the viewpoints of both the laws,
use of only a log- or a power-law to fit the entire velocity profile is not yet a
solved problem.
Considering the method of Coles (1956), given in eq. (1) with a polynomial
for wake, we can describe the entire mean velocity profile. Equation (1) includes
the free parameters as , B and , determined from experimental data. The
power-law model for smooth-wall flows is defined as (Castro-Orgaz 2009, 2010,
2011)
 yu 1/n
u

=
u

(4)

with a coefficient and n an exponent, which can be proposed for the turbulent
boundary layer velocity profile. Afzal (1996) proposed a composite power-law
profile
 yu 1/n
u

=
+ E ,
u

Unauthenticated
Download Date | 2/22/15 5:23 PM

(5)

997

POWER-LAW VELOCITY PROFILE

where is the power law wake function and E is a wake constant (Afzal et al.
2011). Note that eq. (4) is a simplified form of eq. (5) for = 0. George
and Castillo (1997) developed a two-layer (inner and outer) power law theory
in turbulent boundary layer, in contrast to one inner power-law (eq. (2)) velocity
profile. For an accurate description of the entire velocity profile this is one of
the current theories. Another option is the wall-wake law (White 1991). For
integral computations of boundary layer flows using the momentum equation


1
d 1
= C 1+ H
dx 2 f
2

dU 2
,
U 2 dx

(6)

where is the boundary layer momentum thickness, U is the freestream velocity,


x is the streamwise distance along the wall, and H is /, that is, the shape
factor of the boundary layer, it is standard use of eq. (1) coupled with eq. (6)
(White 1991, Schlichting and Gersten 2000). However, an alternative approach
to simplify integral momentum computations of boundary layers is to use a
single power-law within the boundary layer, eq. (1). White (1991) considered
a single power law with n = 7. Thus, in the present work, the inner and outer
regions are not considered for the power law, keeping in mind the associated loss
of accuracy in a point-by-point description of the velocity profile. Instead, an
approximate power law, equivalent to the wall-wake model in terms of integral
momentum modeling, is proposed using eq. (4). The purpose is to simplify the
use of von Krmns equation, eq. (6), in practical applications of boundary
layer flows. The functions for Cf and are needed in eq. (3) for a closure, and,
therefore, some physical conditions need to be imposed to constrain the values
given by eq. (4) based on those given by eq. (1). Note that a one-term power law
(Barenblatt 1993) would not suffice for boundary layer flows. For this reason,
the wall-wake law is considered here as the real velocity profile to propose an
approximate power-law model to be used into eq. (6).
A theory is thus required for the parameters n and , and their dependency
on the Reynolds number. Note that eq. (1) describes a developing shear layer
of skin friction Cf and displacement thickness . Therefore, the power-law
model, of necessity, should produce the same Cf and implicit in eq. (1) for
an accurate description of the boundary layer in an integral sense using eq. (6).
The displacement thickness of the boundary layer is obtained by integrating
eq. (1) as

w 
0

u
u
dy =
1
U
U

11
+
12


.

(7)

Using eq. (4), the displacement thickness is


=

.
1+n
Unauthenticated
Download Date | 2/22/15 5:23 PM

(8)

O. CASTRO-ORGAZ and S. DEY

998

By using the condition for the conservation of mass within the boundary layer,
it results in
U
n=
u

11
+
12

1

2
Cf

1/2 

1/n


1=

11
+
12

1
1,

(9)

from which eq. (4) yields



=

2
Cf

1/2 

(10)

Using eq. (1) at the outer edge of the boundary layer, where u(y = ) = U ,
yields
U
1
= ln
u

2
+B+
=

2
Cf

1/2
.

(11)

Using a conservation of shear-forces at the wall, one may obtain from


eqs. (9)-(11)
1
 


u
11
+
1,
n = ln
+ B + 2

12



1/n

u
2
u
1
ln
+B+
,
=

(12)
(13)

which allows the computation of the power-law velocity profile parameters n


and . Now, assume that is computed using eq. (5), and equate it with eq. (9).
This would result in values of n and including the effect of the power-law wake
constant E in addition to . If we neglect both E and effects, we get
1/2

n = (2/Cf )

1,

(14)

which is an expression widely used in open channel flows (Castro-Orgaz 2009).


Equations (12) and (13) imply a Reynolds number dependence + = u /
of the power-law parameters, as earlier discussed by Barenblatt (1993) and
Barenblatt et al. (2000a, b) using a different theory. Their results are
2
ln + ,
3

(15)

5
1
+ ln + ,
2
3

(16)

n=
=

when taking + as the representative Reynolds number. Other choices for the
Reynolds number are possible, as were adopted by Barenblatt (1993) and Barenblatt et al. (2000a, b). Figure 1 compares the results of the new integral expressions for the power-law parameters n and , given by eqs. (12) and (13) as a
function of + , using = 0.41, B = 5, and = 0.55 for a ZPGTBL, with
Unauthenticated
Download Date | 2/22/15 5:23 PM

POWER-LAW VELOCITY PROFILE

999

those obtained from eqs. (15) and (16) according to Barenblatt. The approach
of Barenblatt yields smaller values of n and than those obtained by using the
integral power-law proposed in this study. It was found that when coefficients
in eqs. (15) and (16) were changed as
13
ln +
16

(17)

1
7
+ ln +
2
3

(18)

n=
=

the new integral theory results were accurately reproduced. Implications of


eq. (14) are included in Fig. 1 (standard approach), clearly indicating a systematic deviation from both the present approach and Barenblatt theory. Given the
dependency of n and on + , the defect profiles deduced from eq. (4) are not
fully similar. Figure 2 compares eq. (2) with the ZPGTBL experimental data
of sterlund (sterlund 1999, sterlund et al. 2000), and with the defect profiles obtained from eq. (4) for + = 9500. sterlund (1999) performed wind
tunnel experiments with flat plates. The wind profiles were measured using
hotwire techniques and the skin friction was measured independently of the velocity measurements using oil-film interferometry. Therefore, the experimental
values of Cf are not deduced from velocity profiles. Measurements were conducted in two different experimental facilities, the Minimum Turbulence Level

Fig. 1. Comparison of the proposed integral theory with the results obtained from the
approach of Barenblatt and eq. (14) (standard approach in open channel hydraulics)
for (a) , and (b) n. New Barenblatt type power-law parameters are fitted with the
analytical results of the integral model.
Unauthenticated
Download Date | 2/22/15 5:23 PM

1000

O. CASTRO-ORGAZ and S. DEY

Fig. 2. Comparison of velocity defect profiles obtained from the proposed integral theory with the wall-wake model and experimental data of sterlund: (a) in natural scale,
and (b) in semi-log scale.

(MTL) wind tunnel at the Royal Institute of Technology and the National Diagnostic Facility (NDF) wind tunnel at the Illinois Institute of Technology, using
5 measuring sections tested with 10 different wind speeds in the former, and
3 sections and 5 free-stream velocities in the latter.
Notably the mean values used herein, = 0.41 and = 0.55, permit
the simulation of the velocity defect profiles of ZPGTBL accurately from the
wall-wake model. It is also shown that, in general, the new power-law profiles derived from the integral theory are in good agreement with the wall-wake
model. No attempt was made to incorporate a gradual transition to the laminar
sub-layer near the wall.
Figure 3a shows the comparison of the defect profiles obtained from eq. (2)
with those from the power-law for + = 1000 and 10 000, displaying small deviations. The same power-law defect profiles are plotted in a semi-log scale in
Fig. 3b. There is a small deviation between the two profiles for y/ < 0.1
(in the wall-zone). A comparison of eq. (1) for the complete velocity profile
u/u [y/] with eq. (4) for the proposed power-law velocity profile is shown in
Fig. 3c for + = 1000 and 10 000, and = 0.41, B = 5, and = 0.55.
There exists a good agreement between the velocity profiles obtained from the
new power-law and the log-wake law, regardless of the values of + . The dimensionless velocity U/u at the outer edge of the boundary layer is the same
for both, as the conservation of Cf was imposed. The power-law profile underestimates the flow velocity roughly in the upper half (y/ > 0.5), whereas it
overestimates in the lower half. This slight asymmetry of the power-law proUnauthenticated
Download Date | 2/22/15 5:23 PM

POWER-LAW VELOCITY PROFILE

1001

Fig. 3. Comparison of the proposed integral power-law with the wall-wake model as
a function of + : (a) velocity defect profiles, (b) velocity defect profiles in semi-log
scale, and (c) velocity profiles.

duces the same displacement thickness as that of a complete velocity profile.


The differences, however, are not important for integral momentum computations, because values of Cf and are preserved between both profiles (Fig. 3c).
Thus, new theory for the power-law closely approximates the complete velocity
profile within the whole thickness of a turbulent boundary layer.
A comparison of eq. (1) for the complete velocity profile u/u [y/] with
eq. (4) for the proposed power-law theory is shown in Figs. 4 and 5, for various
Unauthenticated
Download Date | 2/22/15 5:23 PM

1002

O. CASTRO-ORGAZ and S. DEY

Fig. 4. Comparison of velocity profiles obtained from the proposed integral power-law
theory with those from the wall-wake model and experimental data of sterlund (1999).

values of + , according to the experimental data sets of sterlund (1999) and


= 0.41, B = 5, and = 0.55. There exists a good agreement between the
velocity profiles obtained from the new power-law model, the wall-wake model
and experimental data, regardless of the values of + , supporting the universality of the new expressions for both n and . sterlund (1999) measured the wall
shear stress by an oil film interferometry, which provides information indepenUnauthenticated
Download Date | 2/22/15 5:23 PM

POWER-LAW VELOCITY PROFILE

1003

Fig. 5. Additional comparison of velocity profiles obtained from the proposed integral power-law theory with those from the wall-wake model and experimental data of
sterlund (1999).

dent of the log-law of the wall. Thus, this complete data set was considered
adequate for an impartial evaluation of the wall-wake and power-law profiles.
The power-law profile of eq. (1) may be written more generally as
u
=
u

y
di

1/n
,

(19)

where di = ks , that is, the roughness height for turbulent rough flow and di =
u / for smooth walls. Castro-Orgaz (2009, 2010) and Castro-Orgaz and Hager
(2010) applied eq. (19) to high speed turbulent rough flow in chute channels of
Unauthenticated
Download Date | 2/22/15 5:23 PM

O. CASTRO-ORGAZ and S. DEY

1004

high dams. Thus, the power-law integral parameters may be then formulated in
universal form

n=

2
Cf

1/2 

2
Cf

11
+
12

1/2 

di

1
1,

(20)

(21)

1/n

This result shows that the expressions derived are general, and only the scaling
variable changes between smooth- and rough-bed flows. Future applications of
the proposed integral power-law theory encompass rough flows, which is the
area of growing interest in the study of geophysical flows, i.e., canopy-flows
and river flows.
It is well-known from the turbulent flow theory that the strength of the
wake is not a universal constant. It varies for open channel flow, pipe flow
and boundary layer flow (White 1991). Thus, each experimental condition on
the data sets may result in different values. More importantly, depends on
the pressure gradient. Thus, the use of eqs. (20) and (21) requires an analysis of
the strength of the wake. Castro-Orgaz (2010, 2011) clearly stated that the value
= 0.2 only applies to developing chute flow, neglecting the small pressure
gradients, whereas for the ZPGTBL one has = 0.55.
3.

SKIN FRICTION LAW AND BOUNDARY LAYER PROFILE:


TESTING THE NEW THEORY WITH A ZPGTBL
The boundary layer profile for a ZPGTBL is given by
d
1
Cf =
2
dx

(22)

neglecting dU 2 /dx, e.g., the pressure gradient, in eq. (3). Using eq. (4), eq. (22)
yields
2/n

1 u
n
d
=
.
(23)
2

(n + 1) (n + 2) dx
Equation (23) is the equation developed by Castro-Orgaz (2009) for zero pressure gradient. Equation (21) after integration becomes

+
= x+
,
+
x

(24)

where x+ = xu / , and the new coefficients and are



=

2
+1
n

 2 +1 1
(n + 1)(n + 2) 2 ( n )

,
n
Unauthenticated
Download Date | 2/22/15 5:23 PM

(25)

POWER-LAW VELOCITY PROFILE

2
.
n+2

1005

(26)

The expressions for the skin friction coefficient can be given either by

Cf = 2

1
ln


+B+

2

(27)

or by
2
Cf = 2

2/n
.

(28)

Using the auxiliary relations


1/2
Cf
,
2
 1/2
2
Rex = x+
Cf

U
=

(29)
(30)

eqs. (22)-(30) with eqs. (12) and (13) provide the solution to the boundary layer
flow.
Figure 6a presents the plots of Cf against the streamwise Reynolds number
Rex = U x/ together with the experimental data of sterlund (1999) and the
semi-empirical equations proposed by White (1991) as given below

Cf = 0.02

1/6
,

U
= 0.16 Re6/7
.
x

(31)
(32)

The results obtained from the proposed approach are in good agreement
with data of sterlund (1999), who measured Cf by oil film interferometry,
which gave an independent result from that of any velocity profile measurement. The streamwise coordinate x was considered from the leading edge of
a flat plate, as given by sterlund (1999). No attempts were considered here
to estimate the laminar boundary layer thickness or a virtual origin for the turbulent boundary layer, given the tripping of the boundary layer in sterlunds
experiments (sterlund 1999). The results from the present model form an upper bound of the experimental data. The approach of White (1991) appears to
overestimate Cf in the entire range of Rex tested by sterlund (1999), but there
is an improvement in prediction by the new theory developed herein.
The results obtained from eq. (24) are plotted in Fig. 6b and compared with
the equation given by White (1991)

= 0.16 Re1/7
.
x
x
Unauthenticated
Download Date | 2/22/15 5:23 PM

(33)

1006

O. CASTRO-ORGAZ and S. DEY

Fig. 6. Streamwise evolution of (a) skin-friction coefficient, (b) boundary layer profile, and (c) power-law velocity exponent, obtained from the proposed integral theory,
wall-wake model and experimental data by sterlund; (d) effect of wake functions W
and on power-law velocity profiles.

The new solution is in agreement with Whites equation for high values of
Rex . sterlund (1999) and sterlund et al. (2000) defined the boundary layer
thickness as u(y = 95 ) = 0.95U . The data of sterlund (1999) are shown
in Fig. 6b. The boundary layer thickness, as defined alternatively by u(y =
999 ) = 0.999U , determined from sterlunds measured velocity profiles by
numerical interpolation, is also plotted in Fig. 6b. The prediction using eq. (24)
or (33) lies between 95 and 999 . However, the values of 95 are exceedingly
low, and therefore, the analytical prediction for is shown to be closer to the
values of 999 . Finally, the values of n obtained from the values reported by

sterlund for 95 and 95


are n95 = (95 /95
) 1. These values are compared
in Fig. 6c with the predictions from the integral power law theory. The values
of n95 are very low when compared with the integral theory values. Further,
Unauthenticated
Download Date | 2/22/15 5:23 PM

POWER-LAW VELOCITY PROFILE

1007

experimental velocity profiles by sterlund were used to compute 999 and 999
,

and from these results n999 = (999 /999 )1 was computed. The last values are
shown to be higher than the n values from the integral power law theory. This
analysis shows that the integral momentum method based on the power-law
profile yields intermediate values for n and /x, lying between those derived
from the experimental data defining the nominal limit of the boundary layer
thickness at distance y = 95 from the wall (where u = 0.95U ) and distance
y = 999 (where u = 0.999U ). Lastly, n99 and 99 were determined. Figure 6b, c
reveals that these values are very close to those theoretically computed with the
integral power-law theory proposed here. This comparison is used to show that
the agreement between the integral momentum model and experimental data
relies strongly on the choice of the nominal thickness of the boundary layer for
the analysis of experimental velocity profiles.
The basic issue Can the power-law wake function ( in eq. (5)) be neglected while the log-law wake function (W in eq. (1)) is necessary, can now
be highlighted. Run 981127M of sterlund (1999) is plotted in Fig. 6d. Experimental data is compared with the log-law with wake term W 6= 0 in eq. (1),
resulting in good agreement. Furthermore, the log-law with W = 0 is seen
to be a poor approach outside the wall-layer, as expected. Therefore, W is relevant for the log-law. Now, consider the power law proposed herein, (eq. (4);
= 0), and the power-law parameters n and determined from eqs. (12) and
(13) (W 6= 0). It is seen that the proposed model is in good agreement with
experimental data and the wall-wake law. For comparison purposes, consider
eq. (4) without wake law again ( = 0), but with the parameters n and determined assuming W = 0, e.g., no wake-law in the log-law, that is, assume
eq. (14). The results included in Fig. 6d indicate that the latter is not a good
approximation. Therefore, the wake term in eq. (5) may be neglected ( = 0),
e.g., eq. (4) assumed, whereas the power-law parameters can only be determined
while the contribution of W is retained in eq. (1). The same conclusions are derived from the other runs of sterlund (1999).

4.

NEW POWER-TYPE RELATION FOR SHEAR STRESS


AND EDDY VISCOSITY DISTRIBUTIONS
Once the solution u+ = u+ (y + , + ) is derived from the integral theory, the distributions of the Reynolds shear stress and the eddy viscosity may be obtained
using the continuity and momentum equations for the turbulent flow. Here, u+
is u/u , and y + is yu / . Steady two-dimensional incompressible flow in a
ZPGTBL is given by the continuity equation
u v
+
=0,
x x
Unauthenticated
Download Date | 2/22/15 5:23 PM

(34)

O. CASTRO-ORGAZ and S. DEY

1008

where v is the vertical velocity component. The momentum equation in


x-direction is
u
u
1
u
+v
=
(35)
x
y
y
with
= u0 v 0

(36)

as the Reynolds shear stress (denoted by ), when neglecting the viscous sublayer. Here, u0 and v 0 are the fluctuations in instantaneous streamwise and vertical velocities, respectively. Using eq. (34), integration of eq. (35) results in
w u2
w u
0
=
dy u
dy

x
x
y

(37)

for the Reynolds shear stress distribution. Equation (4) may be written in the
alternative form
 y 1/n
u=U
(38)

to determine u/x and u /x


u
u d
=
,
x
n dx

(39)

2u2 d
u2
=
,
x
n dx

(40)

where U/x = 0 was imposed. Inserting eqs. (39) and (40) into eq. (37)
yields after integration
u0 v 0
=1+
u2

2
Cf


  2
d
1
2
y 1+ n

.
dx 1 + n 2 + n

(41)

The streamwise momentum balance is from eq. (22)


1
n
d
Cf =
2
(n + 1) (n + 2) dx

(42)

which inserted into eq. (41) finally gives the Reynolds stress distribution as
 y 1+ n2
u0 v 0
=
1

.
u2

(43)

Equation (43) is the new power-type distribution of the Reynolds shear stress
across in ZPGTBL using the new integral power-law solution. Equation (43)
for + = 7630 (n = 7.24) is compared in Fig. 7a with the experimental data
of Klebanhoff (Hinze 1975), resulting in a good agreement for 0 < y/ < 0.7.
For y/ > 0.7, there is a slight departure of predictions from the observations
caused by the values of u/x and u2 /x obtained from the power-law model.
Unauthenticated
Download Date | 2/22/15 5:23 PM

POWER-LAW VELOCITY PROFILE

1009

Fig. 7. Comparison of experimental data with the proposed power-law theoretical distributions of (a) Reynolds shear stress, and (b) eddy viscosity.

At y/ = 1, u/y = 0, as given by eq. (1), but in contrast, the power-law


model yields u/y 6= 0.
The eddy viscosity t is defined by
t =

/
.
u/y

(44)

Using eqs. (38) and (43), the power-law model for t is



 y 1+ n2   y 1 n1
t
n
=
(1
+
n)C
1

.
f
U
2

(45)

Equation (45) for + = 7630 (Cf = 0.0023) is compared with the experimental data from Schlichting and Gersten (2000) in Fig. 7b. It is evident
that the order of magnitude of the maximum dimensionless eddy viscosity,
t /(U ) 0.022, corresponds closely with the experimental data. However,
the power-law model displaces the position of maximum dimensionless eddy
viscosity roughly from y/ 0.3 to 0.5. However, the results obtained from the
proposed power-law model are in good agreement with those from a wall-wake
model reported by Schlichting and Gersten (2000), which provides a slightly
better fit to experimental data.

5. CONCLUSIONS
Based on conservation of mass and skin friction in a turbulent boundary layer
with any wake term, a new theory for the power-law velocity profile is proposed. The new power-law produces the same skin friction coefficient and displacement thickness as those obtained from the wall-wake model. The new
Unauthenticated
Download Date | 2/22/15 5:23 PM

1010

O. CASTRO-ORGAZ and S. DEY

theoretical power-law parameters are Reynolds-number dependent, and their


generalized form allows their use for any pressure gradient within the boundary
layer. The present power-law theory allows advancement in integral momentum
computation of boundary layer flows, proposing a unique power-law relationship for the entire boundary layer. The present theory has been verified by the
experimental data for ZPGTBL, and a good agreement has been found for the
velocity profiles (both in absolute and defect forms) and skin friction coefficient.
The developed theory was used to critically review estimates of the boundary
layer thickness profiles, indicating the dependence on the nominal limit considered for the boundary layer thickness. From the proposed power-law model, the
Reynolds shear stress and eddy viscosity distributions are derived, and they are
also found to be in good agreement with the experimental data.

References
Afzal, N. (1996), Wake layer in turbulent boundary layer with pressure gradient: A new
approach. In: K. Gersten (ed.) Iutam Symposium on Asymptotic Methods for
Turbulent Shear Flows at High Reynolds Numbers: Proceedings of the Iutam
Symposium Hes, Springer, London, 95-118.
Afzal, N. (2001a), Power law and log law velocity profiles in fully developed turbulent
pipe flows: equivalent relations at large Reynolds numbers, Acta Mech. 151,
171-183, DOI: 10.1007/BF01246916.
Afzal, N. (2001b), Power law and log law velocity profiles in turbulent boundary layer
flows: equivalent relations at large Reynolds numbers, Acta Mech. 151, 195216, DOI: 10.1007/BF01246918.
Afzal, N. (2005), Analysis of power law and log law velocity profiles in overlap region of a turbulent wall jet, Proc. R. Soc. A 461, 2058, 1889-1910, DOI:
10.1098/rspa.2004.1400.
Afzal, N. (2007), Power law velocity profile in turbulent boundary layers
on transitional rough wall, J. Fluids Eng. 129, 8, 1083-1100, DOI:
10.1115/1.2746902.
Afzal, N. (2008), Power-law velocity profile in a turbulent Ekman layer on a transitional rough surface, Quart. J. Roy. Met. Soc. 134, 634, 1113-1125, DOI:
10.1002/qj.285.
Afzal, N., A. Seena, and A. Bushra (2006), Power law turbulent velocity profile
in transitional rough pipes, J. Fluids Eng. 128, 3, 548-558, DOI: 10.1115/
1.2175161.
Afzal, N., A. Seena, and A. Bushra (2011), Discussion to velocity profile and flow resistance models for developing chute flow, J. Hydraul. Eng. 137, 8 (in press).
Balachandar, R., D. Bakely, and J. Bugg (2002), Friction velocity and power law velocity profile in smooth and rough shallow open channel flows, Can. J. Civ.
Eng. 29, 256-266, DOI: 10.1139/L01-093.
Unauthenticated
Download Date | 2/22/15 5:23 PM

POWER-LAW VELOCITY PROFILE

1011

Barenblatt, G.I. (1993), Scaling laws for fully developed turbulent shear flows.
Part I: Basic hypothesis and analysis, J. Fluid Mech. 248, 513-520, DOI:
10.1017/S0022112093000874.
Barenblatt, G.I., A.J. Chorin, and V.M. Prostokishin (2000a), A note on the intermediate region in turbulent boundary layers, Phys. Fluids 12, 2159-2161, DOI:
10.1063/1.1287613.
Barenblatt, G.I., A.J. Chorin, and V.M. Prostokishin (2000b), Self-similar intermediate structures in turbulent boundary layers at large Reynolds numbers,
J. Fluid Mech. 410, 263-283.
Buschmann, M.H., and M. Gad-el-Hak (2003a), Generalized logarithmic law and its
consequences, AIAA J. 41, 1, 40-48, DOI: 10.2514/2.1911.
Buschmann, M.H., and M. Gad-el-Hak (2003b), Debate concerning the meanvelocity profile of a turbulent boundary layer, AIAA J. 41, 4, 565-572, DOI:
10.2514/2.1994.
Buschmann, M.H., and M. Gad-el-Hak (2006), Recent developments in scaling of wall-bounded flows, Prog. Aerosp. Sci. 42, 5-6, 419-467, DOI:
10.1016/j.paerosci.2007.01.001.
Castro-Orgaz, O. (2009), Hydraulics of developing chute flow, J. Hydraul. Res. 47, 2,
185-194, DOI: 10.3826/jhr.2009.3462.
Castro-Orgaz, O. (2010), Velocity profile and flow resistance models for developing chute flow, J. Hydraul. Eng. 136, 7, 447-452, DOI: 10.1061/
(ASCE)HY.1943-7900.0000190.
Castro-Orgaz, O. (2011), Closure to velocity profile and flow resistance models for
developing chute flow, J. Hydraul. Eng. 137, 8 (in press).
Castro-Orgaz, O., and W.H. Hager (2010), Drawdown curve and turbulent boundary layer development for chute flow, J. Hydraul. Res. 48, 5, 591-602, DOI:
10.1080/00221686.2010.507337.
Chen, C.L. (1991), Unified theory on power laws for resistance, J. Hydraul. Eng. 117,
3, 371-389, DOI: 10.1061/(ASCE)0733-9429(1991)117:3(371).
Coles, D.E. (1956), The law of the wake in the turbulent boundary layer, J. Fluid Mech.
1, 2, 191-226, DOI: 10.1017/S0022112056000135.
Dean, R.B. (1976), A single formula for the complete velocity profile in a turbulent
boundary layer, J. Fluids Eng. 98, 723-726.
George, W.K. (2007), Is there a universal log-law for turbulent wall-bounded flows?
Phil. Trans. Royal Soc. A 365, 789-806, DOI: 10.1098/rsta.2006.1941.
George, W.K., and L. Castillo (1997), The zero pressure-gradient turbulent boundary
layer, Appl. Mech. Rev. 50, 689-729, DOI: 10.1115/1.3101858.
Granville, P.S. (1976), Modified law of the wake for turbulent shear layers, J. Fluids
Eng. 98, 578-580.
Guo, J., P.Y. Julien, and R.N. Meroney (2005), Modified log-wake law for zero pressure gradient turbulent boundary layers, J. Hydraul. Res. 43, 421-430.
Hinze, J.O. (1975), Turbulence, McGraw-Hill, New York.
Koll, K. (2006), Parameterisation of the vertical velocity profile in the wall region over
rough surfaces. In: E.C.T.L. Alves et al. (eds.), Proceedings of the InternaUnauthenticated
Download Date | 2/22/15 5:23 PM

1012

O. CASTRO-ORGAZ and S. DEY

tional Conference on Fluvial Hydraulics, River Flow 2006, 6-8 September


2006, Lisbon, Portugal, Vol. 1, 163-171, Taylor & Francis, London, DOI:
10.1201/9781439833865.ch15.
Krogstad, P.A., R.A. Antonia, and L.W.B. Browne (1992), Comparison between
rough- and smooth-wall turbulent boundary layers, J. Fluid Mech. 245, 599617, DOI: 10.1017/S0022112092000594.
Meyer, Z. (2009), Modified logarithmic tachoida applied to sediment transport in a
river, Acta Geophys. 57, 2, 743-759, DOI: 10.2478/s11600-009-0064-z.
sterlund, J.M. (1999), Experimental studies of zero pressure-gradient turbulent
boundary layer flow, Ph.D. Thesis, Royal Inst. of Tech., Stockholm, Sweden.
sterlund, J.M., A.V. Johansson, H.M. Nagib, and M.H. Hites (2000), A note on the
overlap region in turbulent boundary layers, Phys. Fluids 12, 1, 1-4, DOI:
10.1063/1.870250.
Rotta, J.C. (1962), Turbulent boundary layers in incompressible flow, Prog. Aerosp.
Sci. 2, 1-220, DOI: 10.1016/0376-0421(62)90014-3.
Rowiski, P.M., J. Aberle, and A. Mazurczyk (2005), Shear velocity estimation in
hydraulic research, Acta Geophys. 53, 553-565.
Schlichting, H., and K. Gersten (2000), Boundary-Layer Theory, Springer, Berlin.
White, F.M. (1991), Viscous Fluid Flow, McGraw-Hill, New York.
Wosnik, M., L. Castillo, and W.K. George (2000), A theory for turbulent pipe and
channel flows, J. Fluid Mech. 421, 115-145.
Zagarola, M.V. (1996), Mean Flow Scaling of Turbulent Pipe Flow, Ph.D. Thesis,
Dept. of Mechanical and Aerospace Eng., Princeton Univ., Princeton, NJ.
Zagarola, M.V., and A.J. Smits (1998), Mean-flow scaling of turbulent pipe flow,
J. Fluid Mech. 373, 33-79.
Zagarola, M.V., A.E. Perry, and A.J. Smits (1997), Log laws or power laws: The scaling in the overlap region, Phys. Fluids 9, 2094-2100, DOI: 10.1063/1.869328.
Received 27 July 2010
Received in revised form 18 March 2011
Accepted 11 April 2011

Unauthenticated
Download Date | 2/22/15 5:23 PM

También podría gustarte