Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

See

discussions, stats, and author profiles for this publication at:


https://www.researchgate.net/publication/285955226

On the definition of the shear


velocity in rough bed open channel
flows

Article · August 2006


DOI: 10.1201/9781439833865.ch7

CITATIONS READS

22 154

5 authors, including:

Dubravka Pokrajac Costantino Manes


University of Aberdeen Politecnico di Torino
62 PUBLICATIONS 1,016 CITATIONS 51 PUBLICATIONS 645 CITATIONS

SEE PROFILE SEE PROFILE

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

Ureteric Stents View project

FRAMAB - Flood Risk Assessment and mitigation for Masonry Arch Bridges
View project

All content following this page was uploaded by Vladimir Nikora on 09 December 2015.

The user has requested enhancement of the downloaded file.


River Flow 2006 – Ferreira, Alves, Leal & Cardoso (eds)
© 2006 Taylor & Francis Group, London, ISBN 0-415-40815-6

On the definition of the shear velocity in rough bed open channel flows

D. Pokrajac
Department of Engineering, University of Aberdeen, Aberdeen, UK

J.J. Finnigan
CSIRO, Marine and Atmospheric Research, Canberra, Australia

C. Manes, I. McEwan & V. Nikora


Department of Engineering, University of Aberdeen, Aberdeen, UK

ABSTRACT: Shear velocity u∗ is a well-known fundamental velocity scale of the wall turbulence, widely
used for scaling various flow quantities such as mean velocity and turbulent intensities. The definition of shear
velocity is closely related to the universal Logarithmic law of the wall, which was originally derived for constant
shear stress boundary layers. Because the shear stress is constant, the momentum flux ρu∗2 is equal to the wall
shear stress, so u∗ is often understood as some kind of synonym for the wall shear stress. The universal Log-law is
also used for the flows with linear shear stress such as pipe flows and uniform open channel flows. In such flows
the turbulent momentum flux delivered to the roughness differs from the wall shear stress, but the difference is
negligible for the rough walls where boundary layer thickness (flow depth in case of uniform open channel flow),
is much larger than roughness height. In rough wall flows where the flow depth and the roughness height are of
the similar order, the difference between the turbulent momentum flux representative of the free outer flow and
the wall shear stress may become significant. For such flows it is important to distinguish between the wall shear
stress and the shear stress used for evaluation of the shear velocity. This paper highlights the problems associated
with the definition of the shear velocity in shallow open channel flows and its use for evaluation of the bed shear
stress. The influence of the definition of u∗ on the data analysis is illustrated using experimental results.

1 INTRODUCTION channel flow. For instance, although local flow prop-


erties near rough bed always vary in space, globally
Open channel flows over rough beds are of consid- uniform flow is defined by the requirement that
erable engineering interest, but there are still some the double-average flow properties do not change
uncertainties about them. One of the difficulties in in the streamwise direction (Nikora et al., submit-
investigating such flows is that flow properties near ted). Furthermore DAM provides strict definitions
the bed are spatially inhomogeneous. Instead of study- of the spatially-averaged or macroscopic momentum
ing time-averaged flow properties which vary in balance, fluid shear stress and bed shear stress.
space, it is more convenient to investigate double- However, there is one more question about the
averaged flow properties, using the double-averaged rough bed flows that has so far been neglected. Rough
(in time/ensemble and spatial domains) Navier-Stokes beds have complex geometry such that the bed level
equations. The idea of spatial flow averaging was first (height above an arbitrary bed-parallel datum) varies
introduced in hydraulics by Smith and McLean (1977). in space. In DAM this variation is expressed by a
The development of a new methodology based on single parameter – the roughness porosity function,
spatially averaged equations was initiated by Wilson equal to volume of water/total volume within the spa-
and Shaw (1977) for describing atmospheric flows tial averaging domain (thin bed-parallel slices). The
within vegetation canopies. The atmospheric physi- porosity function varies between one, above the rough-
cists Raupach and Shaw (1982), Finnigan (1985), ness crests, and zero (for impermeable bed), below the
Raupach et al. (1991), provided the mathematical basis lowest roughness troughs. At the level of the roughness
for a new set of equations. crest this variation can be abrupt or gradual, depend-
Double-averaging methodology provides clear def- ing on the roughness geometry. In either case it is not
initions of some basic concepts in rough bed open clear a priori where exactly, between roughness crests

89
and troughs, is the appropriate position of an equiv-
alent imaginary bed i.e. the idealised sharp interface
between turbulent flow and the rough wall. This affects
the definition of basic concepts such as flow depth
and the shear velocity, especially for flows with small
relative submergence (flow depth/roughness height
ratio).
The shear or friction velocity is usually defined
as the velocity scale of the momentum absorbed by
the bed. It can be evaluated either at the roughness
crests, where the bed starts extracting momentum from
the flow, or at the troughs, where the total momen-
tum is absorbed, or somewhere between them. Various
versions are present in the literature and yet called Figure 1. Definition sketch.
the same name. It introduces confusion and makes it
difficult to compare published data.
In this paper we present a case for defining the water and the grains. The double-averaged momentum
universal velocity scale u∗ in rough bed flows from equation in x direction is (Nikora et al. 2004)
the shear stress at the roughness crests, rather than
from the total momentum extracted from the flow i.e.
from the bed shear stress. In order to clearly define the
bed shear stress we first use DAM to summarise the
momentum balance in uniform flows over rough beds. where g = gravitational acceleration, ρ = density,
We then revisit the original derivations of the univer- α = angle between the bed and the horizontal datum,
sal Log-law where u∗ was first introduced, and point φ = porosity of the rough bed=volume of water/total
out that the Log-law was first developed for constant volume dV , τxz is the fluid shear stress and fx is the
shear stress boundary layers, so it does not answer the total drag force exerted by the fluid on the roughness
question about the appropriate definition of the shear per unit plan area and unit height dz. For an averaging
velocity. In case of rough bed open channel flows with volume above the roughness crests dV contains only
linear shear stress only the shear velocity associated water, so φ = 1 and fx vanishes from the equation. The
with the roughness crest has a potential to be univer- fluid shear stress is
sal, so it is adopted as the appropriate velocity scale
for scaling flow properties. The consequences of using
the velocity scale based on the bed shear stress instead
are illustrated using a case from the literature and our
own experiments. where the three terms are the spatially averaged vis-
cous stress, the spatially averaged turbulent stress and
the form-induced stress, respectively. The over-bar
denotes time average,  = fluctuation,   = spatial
2 DOUBLE-AVERAGED OPEN CHANNEL average and tilde = the deviation of a time-averaged
FLOW quantity from its spatial average. The drag term is
obtained by integrating local pressure and viscous
We consider steady uniform two-dimensional open stress over the interface between water and grains, S,
channel flow over a rough bed. A right-hand within the averaging volume dV :
co-ordinate system is used with x = streamwise,
y = lateral and z = bed-normal coordinates, and the
velocity components in x, y, z direction are denoted
by u, v, w respectively (Fig. 1). Roughness height is
denoted by k. and the origin of z is placed at the where nx , ny , nz are the components of the unit vector
roughness crest. For simplicity, roughness elements n, normal to the surface dS, pointing into the fluid
have identical height. In natural sediments, where this (Fig. 1).
is clearly not the case, some statistical property of the Spatial averaging smoothes spatial heterogeneity of
bed level e.g. z99% can be used as a global level of the flow near the rough bed in (x, y) planes, so that all
roughness crest. flow variables are represented as profiles in z. The
The double-averaged momentum equations are lowest point in such profiles is the lowest point of the
obtained by averaging Reynolds equations over thin bed roughness. Time-averaged flow depth measured
bed-parallel volumes dV (Fig. 1). These volumes may from that point is denoted by H , and time-averaged
intersect with the bed so in general they contain both flow depth measured from the roughness crests is h.

90
stress diminishes and the total momentum from the
flow above is transferred to the roughness:

The total momentum flux from the fluid flow is there-


fore in balance with the total drag. The total drag
exerted by the fluid on all roughness elements, per
unit plan area of the bed, is the spatially averaged or
macroscopic bed shear stress,

Figure 2. Momentum balance in 2d rough bed flow. Due to the additional momentum supply below the
roughness top, the bed shear stress τbed differs from
the fluid shear stress at the roughness top τ0
2.1 Momentum balance
In boundary layer terminology open channel flow 2.2 Flow layers
belongs to ‘constant pressure gradient’ boundary layer In ‘deep’ rough bed flows, where the influence of the
flow, because its driving force, gravity, plays a role in bed roughness is significant only over a small part
the momentum balance analogous to the pressure gra- of the flow depth (less than around 0.15–0.2 h), the
dient. In steady uniform flow over a rough bed the following layers can be distinguished:
terms of momentum balance can be obtained by inte-
grating the momentum equation (1) over z. (Nikora (i) The Outer Layer, similar to the outer layer
et al. 2001, Nikora et al., submitted). For any bed- for flows over hydraulically smooth beds, with
parallel plane above the roughness top (Fig. 2) the total the characteristic scales: the shear velocity, u∗ ,
driving force or momentum supply (per unit plan area) the maximum velocity, U∞ (assumed to occur
between the plane and the free water surface is at the free surface), flow depth, h, and the dis-
tance above the zero-plane. (zero-plane= origin
of the Log-law).
(ii) The Logarithmic layer, with the scales u∗ , and
the distance above the zero-plane. A single length
scale exists only over a region far away from both
where z = height of the plane above the roughness top the water surface and the rough bed so that it is
and S = bed slope (Fig. 2 left). It is in balance with not influenced by their respective length scales.
the momentum flux through the plane at the height z, (iii) The Roughness layer, also called the Roughness
τxz (z), so τxz increases linearly, starting from the zero sub-layer, is strongly influenced by the set of sur-
at the free surface. Far away from the roughness both face length scales Li that describe bed geometry,
viscous stress and form-induced stress are negligible including characteristic roughness height, k, so
so τxz is turbulent shear stress. Near the roughness its characteristic scales are u∗ , and Li . This layer
crest persistent vortices behind roughness elements may extend up to 4k above the roughness crest
cause form-induced momentum flux so momentum (Raupach et al. 1991). Flow within the Rough-
is transferred from the flow above by both turbulent ness sub-layer is strongly affected by coherent
and form-induced shear stress (Fig. 2 right). At the eddies generated at the top of the roughness. In
roughness crest, z = 0, the momentum flux from the plant canopy flows the relevant length scale of the
flow above the roughness is just sufficient to balance Roughness sub-layer is determined by the vor-
ρghS, so the fluid shear stress is ticity thickness δw = u(0)/(du/dz)(0) (Raupach
et al. 1996). This was also observed in flows
over relatively dense aquatic vegetation canopy
(Nepf & Vivoni, 2000, Poggi et al. 2004).
Below the roughness crest, between the plane z = 0
and a bed-parallel plane at level z there is an additional For relatively shallow open channel flows some of
momentum supply of φρg(0 − z)S. The correspond- the layers may be absent, because the Roughness sub-
ing total momentum supply of ρgS[h + φ(0 − z)] is layer covers the part of the flow profile ‘reserved’
partly transferred further down via fluid shear stress for the Log-layer, or even extends to the free surface,
and partly transferred to the roughness via form drag obliterating the Outer layer as well.
and viscous drag , which act as momentum sinks. When all three flow layers are present they have a
At the level of the roughness troughs fluid shear single velocity scale u∗ .The Logarithmic layer has only

91
one spatial scale, so it offers the simplest framework
for studying u∗ . For that reason the definition of the
shear velocity and its application in boundary layer
theory are closely related to the universal Logarithmic
law of the wall.

3 THE UNIVERSAL LOG-LAW

The universal Log-law was originally derived for con-


stant shear stress (zero pressure gradient) boundary
layers over smooth walls. It can be derived in several
ways. The two well-known approaches are:
(i) The overlap approach where the dimensional
analysis is combined with asymptotic match-
ing (Millikan, 1939, Yaglom, 1979). Asymptotic
matching requires a clear scale separation between
the boundary layer thickness and the thickness of Figure 3. Logarithmic velocity profile above rough bed.
the Roughness sub-layer. This derivation does not
rely on the constant shear stress: the only assump- where κ is the universal von Karman constant and
tion is that a single velocity scale u∗ is responsible d defines the origin of the logarithmic velocity pro-
for momentum transfer from the outer flow to the file (Fig. 3). Due to the complex wall geometry and
wall. This velocity scale is usually called the fric- flow interactions, the position of this origin is not a
tion velocity and calculated from the wall shear priory known, so it has to be found by fitting. Experi-
stress. Due to the scale separation, the difference mental evidence shows that the origin of the Log-law,
between the momentum absorbed at the level of called zero-plane, is typically located between rough-
roughness crests and troughs, i.e. between τ0 and ness crests and troughs and that it moves towards the
τbed , is negligible. troughs for more energetic flow and sparse roughness
(ii) The momentum flux approach where the momen- elements (Nikora et al. 2002). The zero-plane position
tum balance equation (τxz = −ρu w = τwall ) is at z = −d implies that the length scale relevant within
combined with an assumption about the rela- the Log-layer is z + d. If the Log-law is derived for a
tionship between turbulent momentum flux (tur- constant shear stress boundary layer, with the mixing
bulent shear stress) and velocity gradient either length assumed proportional to the distance from the
directly (Landau, 1944, Monin and Yaglom 1971) origin,  = κ(z + d), the zero plane defines the position
or through an assumption on the mixing length or where the mixing length is zero.
turbulent viscosity. This approach requires that the The roughness length z0 is formally an integration
fluid shear stress and the related velocity scale u∗ constant of the Log-law (height above the zero-plane
are constant along z. In such case it clearly makes no where the velocity obtained from the Log-law is zero,
difference whether u∗ is evaluated at the roughness Fig. 3). It depends on the roughness characteristics
crests or troughs. and is a measure of the efficiency of the surface in
In either of the two approaches the condition for absorbing momentum.
the existence and use of the Logarithmic layer is that The Log layer has to be outside the direct influ-
it occurs far enough from both boundaries (at suffi- ence of the surface roughness, so it must be located
cient distance from the wall, in terms of wall units z + , above the Roughness sub-layer. On the other hand,
and at the same time at height much smaller than the to ensure that the Log-layer is sufficiently far away
boundary layer thickness), so that the only remaining from the outer boundary of the boundary layer, it is
relevant length scale is the distance from the wall z. assumed to be limited to some small portion, usu-
In the ‘overlap’ approach this assumption is explicit. ally 0.15–0.2 of the boundary layer thickness. In cases
In the ‘momentum flux’ approach the assumed rela- where roughness generates a significant Roughness
tionship between shear stress and velocity gradient is sub-layer which spreads higher than 0.15–0.2 of the
valid only in the vicinity of the wall where the turbulent boundary layer the Log-layer does not have enough
momentum flux is not affected by the outer flow scales. space to develop.
The Log-law for boundary layers over rough walls The Log-law is also widely used for boundary lay-
can be stated as ers with linear shear stress such as pipe flows and open
channel flows with h, H  k. The ‘overlap’ derivation
of Log-law explains why it works well with variable
shear stress (provided that h, H  k). It can also be

92
argued that the shear stress is still approximately con- (with the assumed κ = 0.4) to the logarithmic for-
stant between the wall and the Log-layer because the mula (8). (Wang et al. 1993, Wang & Dong 1996).
Log-layer has to be close to the wall (Monin &Yaglom, It is not clear where exactly is τlog in the shear stress
1971), profile.
In shallow open channel flows over rough bed the
The first option clearly cannot yield the universal
roughness height is comparable to water depth. Neither
velocity scale because the bed shear stress depends
the ‘overlap’ approach nor the ‘momentum balance’
on the roughness porosity function φ, which is not
approach is applicable, since there is no scale sep-
universal. To illustrate this let us assume flow over a
aration and the shear stress is not constant. There
bed composed of closely packed rods forming a porous
is no theoretical basis for the universal Log-law and
medium with very low permeability, so that only a thin
the velocity scale responsible for momentum trans-
uppermost layer of the porous bed interacts with the
fer between the Outer layer and the bed becomes less
mean velocity profile above, while below this layer
obvious and Despite that, many investigators present
flow is driven solely by the momentum supplied by
experimental data fitted to the logarithmic equation
gravity within the bed. Increasing height of the rods
(8) and scaled with the shear velocity u∗ . The choice
would not affect the mean velocity profile of the free
of u∗ in shallow open channel flows is discussed in the
surface flow above them, they would remain identical.
remaining part of the paper.
However if we choose to scale them with the shear
velocity calculated from the bed shear stress the scaled
velocity profiles would differ by a factor proportional
4 DEFINITION OF THE SHEAR VELOCITY to the increase in the height of the rods. Thus we would
detect a non-physical suppression of velocities, which
The shear velocity u∗ discussed here is a velocity scale is an artefact of the selection of the shear velocity.
expected to universally collapse data for rough bed Two other options, zero-plane and the Log-layer,
open channel flows including flows with low rela- rely on the existence of the Log-layer and the uni-
tive submergence (with and without the-log layer). In versal Log-law. This makes it impossible to use them
other contexts e.g. sediment entrainment and sediment for shallow flows where the existence of the universal
transport different definitions of u∗ are required. Log-layer is questionable.
Although the u∗ was first introduced in the con- The remaining option of defining the u∗ from the
text of the Log-law, the original derivations of the fluid shear stress at the roughness crest is the sim-
Log-law do not provide a general definition of the plest and the least ambiguous one. In order to achieve
u∗ which would be applicable to the boundary layers uniform flow, flow depth, the velocity profile and the
with variable shear stress where the roughness height turbulence structure must adjust to deliver a momen-
is comparable with the boundary layer thickness. They tum flux to the roughness that just balances the gravity
associate u∗ with the momentum absorbed by the wall. forcing of the free fluid. This momentum flux is τ0 and
However this definition does not specify where exactly the corresponding velocity scale is u∗0 . At the rough-
(along the roughness height k) is this momentum to be ness crest form drag and viscous drag start affecting
evaluated. This is the reason why there is no general the flow introducing a major change in flow physics.
consensus in the literature on the appropriate choice Below the roughness crest flow adjusts to the micro-
of the u∗ . The following options exist: physics of drag and porosity in the roughness, so that
√ the bed shear stress and the corresponding velocity
• u∗bed = τbed , where τbed is the bed shear stress. scale reflect them, rather than solely the characteris-
This classical approach, which assumes that u∗ is an tics of free fluid. For these reasons we have adopted
equivalent way of expressing the bed shear stress, the shear velocity associated with the shear stress at
is widely used in the open channel flow literature the roughness crest as the appropriate velocity scale
(e.g. Bayazit 1976, Ditrich & Koll, 1997, Katul et al. for scaling turbulent flow quantities across the whole
2002,√Rowinski et al.). flow profile.
• u∗0 = τ0 , where τ0 is the total fluid shear stress
at the roughness crest (at z = 0 in Figure 2). This
approach is consistently used in the canopy flow
literature (e.g. Finnigan 2000, Poggi et al. 2004, 5 SHEAR VELOCITY AND THE DATA
Jarvela, 2005), but may be found in open channel INTERPRETATION
flow papers
√ (e.g. Nakagawa 1991).
• u∗d = τd , where τd is the total fluid shear stress Shear velocity is widely used as a universal velocity
extrapolated to the level of zero plane, z = −d scale, often in the context of the Log-law, but also for
(Nepf & Vivoni, 2000, Ditrich & Koll, 1997). scaling the data on shallow rough bed open channel
log √
• u∗ = τlog , where τlog is the shear stress that cor- flows where the Log-layer is not present (e.g Katul
responds to u∗ obtained by fitting experimental data et al. 2002). In this section we first analyse how the

93
choice of the shear velocity affects the parameters A given data set has a given set of velocity gradients.
of the Log-law in its original form (8) and present The slope of the best-fit line through their reciprocal
an example from the literature. We then demonstrate, values gives the ratio κ/u∗ and the intercept gives the
using our own measurements of turbulent intensities, position of the zero-plane d.
the general effect of the shear velocity definition on The shear velocity evaluated from the bed shear
interpretation of data on shallow flows over rough bed. stress is over-estimated by the factor shown in Fig-
Only the conventional ‘bed shear stress’ shear ure 4, so that κ is over-estimated by the same factor.
velocity is compared with the adopted ‘roughness This means that for flows with very small relative sub-
crest’ shear velocity. The effect of other versions of the mergence using the ‘bed shear stress’ shear velocity
shear velocity definition, listed in the previous section, instead of the ‘roughness crest’ velocity can have sig-
can be easily evaluated using an analogous procedure. nificant influence on the reported values of the von
From the momentum balance explained in the section Karman constant.
2.1 (Fig 2) it follows that the bed shear stress and the Figure 5 shows the results of Bayazit, 1976, who
fluid shear stress at the roughness crests are related as reported gradually increasing von Karman constant
as the relative submergence decreased. The values of
κ were evaluated by fitting the velocity data to the
Log-law, with the position of the zero-plane d = 0.33 k
below the roughness crest for all experiments. In the
so that same paper Bayazit made a comment that these large
values of κ for small relative submergences may be due
to the wrong choice of the velocity scale u∗ , but did

Obviously using u∗bed as approximation of u∗0 always


overestimates it. The ratios τ∗bed /τ∗0 and u∗bed /u∗0 are
plotted in Figure 4 as functions of the relative sub-
mergence H /k for three typical values of porosity φ
Porosity 0.3–0.5 is approximately the range that can
be expected in natural gravel beds, while 0.9 or higher
may occur in aquatic vegetation. In the former case the
difference between u∗bed and u∗0 becomes significant
(larger than 10%) for the relative submergences less
than 2.5-3.5 while in the latter case it happens for those
less than 5. The graphs evaluate the effect of using u∗bed
instead of u∗0 on the quantities scaled with the shear
velocity. Two examples are presented in the following
text to illustrate the consequences of this effect. Figure 4. Ratio between the bed shear stress and the shear
stress at the roughness crest (white symbols) and between the
corresponding shear velocities (black symbols), for a range
5.1 Mean velocity and Log-law parameters of relative submergences H/k and 3 different porosities φ.
Applicability of the universal Log-law to shallow open
channel flows and the alternative formulations of the 1.60
log-law (adjusted to linear shear stress and flows Original
1.40
without clear scale separation) are beyond the scope Re-scaled

of this paper. Here we analyse solely a case, often 1.20


found in the literature, of fitting the conventional
logarithmic velocity profile given by (8) to the experi- 1.00

mental velocity data for shallow flows with low relative 0.80
submergence H /k.
For a given set of experimental data, fitting to the 0.60
log-law produces a single ratio κ/u∗ , provided that
0.40
κ/u∗ and d wee both determined by fitting, The most
obvious way to demonstrate this is by differentiating 0.20
the log formula for u with respect to z, which yields 1.0 1.5 2.0 2.5 3.0 3.5 4.0
(Nikora et al. 2002) h/k

Figure 5. Von Karman constant reported in Bayazit, 1976


and re-calculated with the ‘roughness crest’ shear velocity.

94
not attempt to correct it. These values of κ are here re- 5.2 Turbulent intensities
calculated with the ‘roughness crest’ shear velocity u∗0
The difference between the ‘bed shear stress’ and
instead of the ‘bed shear stress’ u∗bed from the original ‘roughness crest’shear velocity has an important influ-
paper. Figure 5 shows that the correction significantly ence on the turbulence characteristics reported for the
changes κ values. rough bed shallow open channel flows. For instance
This discussion is not aimed at listing all potential turbulent intensities are usually scaled with the u∗ , and
problems with fitting the velocity data to the logarith- compared with the corresponding dimensionless tur-
mic law. The aim is to point out just one particular bulent intensities measured in deep rough bed flows or
source of error. The literature on shallow rough bed smooth bed flows. Scaling the turbulent intensities for
flows contains a range of values of von Karman con- deep flows is not sensitive to the difference between u∗0
stant. In cases where the reported values of κ are
and u∗bed , but for shallow flows the difference becomes
less than 0.4 the correction for u∗bed /u∗0 would further
reduce them. In flows with small to intermediate rela- significant. Using u∗bed instead of u∗0 for scaling intro-
tive submergence the universal Log-layer is often not duces an artificial suppression of turbulent intensities,
present, so there is no reason for κ obtained by fitting caused by the choice of the velocity scale rather than
the velocity data to have the universal value. How- the real physics. The ‘suppression factor’progressively
ever it is important to realise that without a physically increases with decreasing relative submergence. The
based and consistent velocity scale u∗ it is very diffi- maximum turbulent intensity (scaled with u∗ ) is known
cult or impossible to compare different experimental to decrease when the relative submergence decreases
data. Correcting the data for bias introduced by using (Wang et al. 1993, Ditrich & Koll 1997). Part of this
u∗bed would clarify the genuine influence of the relative finding may be the artefact of the choice of the velocity
submergence and other parameters that affect velocity scale.
profiles. To explore this possibility we first re-calculated the
An inverse situation happens when the u∗ , deter- results from Bayazit (1976), where turbulent intensi-
mined by fitting the log-law to the measured velocity ties were fitted to the formula
profile (assuming κ = 0.4), is used for evaluation of the
bed shear stress or the friction factor. This is frequent
in field measurements, where using u∗ is a simpler and
more robust alternative to the direct measurement of In Bayazit (1976), turbulent intensities scaled with u∗
the shear stress. In such cases the result of fitting is and plotted against ln u∗ z/ν do not collapse for dif-
the ‘roughness crest’ shear velocity u∗0 , but it is used ferent relative submergences. Instead, they follow a
as if it were ‘bed shear stress’ velocity u∗bed . Since u∗0 set of parallel lines, indicating that only the coeffi-
is always less than u∗bed , this means that, even if the cient A from (12) is responsible for the reduction of
assumed κ = 0.4 is correct, the bed shear stress and the turbulent intensities as the relative submergence (h/k)
friction factor are underestimated by factors shown in decreases. Figure 6 shows the values of A from the
Figure 4. original paper and those corresponding to ‘roughness
The above analysis applies to the cases where κ/u∗ crest’ u∗ as a velocity scale. The re-calculated A val-
and d were both found by fitting. However the position ues vary much less, indicating very small variation of
of the zero-plane is sometimes found using the centre- turbulent intensities with relative submergence. It is
of-pressure method (Jackson, 1981). This method is
not applicable to ‘constant pressure gradient’ bound- 6.0
ary layers because the momentum supply below the
roughness crest affects the position of the centre of 5.0

drag. If d was evaluated using this method the error


4.0
would combine with the error in u∗ and further affect
the value of κ. 3.0
A

For rough bed open channel flows with small


to intermediate relative submergence the use of the 2.0
Log-law in its original form is questionable and
alternative options should be explored. In any case, 1.0
Original
in future efforts to compile available experimental Re-scaled
0.0
data it is very important to make a clear distinction 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
between the non-universal ‘bed shear stress’ velocity h/k
scale u∗bed , which depends on the particular rough-
ness geometry and the ‘roughness crest’ velocity scale Figure 6. The constant A (eq. 5) defining maximum tur-
u∗0 , which has a potential to universally collapse the bulent intensities σu reported in Bayazit, 1976 and re-scaled
experimental data. with the ‘roughness crest’ shear velocity.

95
clear that turbulent intensities profiles would collapse out under uniform, fully turbulent rough flow condi-
much better when scaled with the ‘roughness crest’ tions. The aspect ratio varied from 18-5.4 for which
shear velocity. we can assume that the flow was two-dimensional
Further analysis was performed with our own mea- (Kironoto and Graf 1994). The main hydraulic param-
surements of flow over a single layer of spheres packed eters for the experiments are shown in Table 1.
in the cubic pattern. All the experiments were carried Measurements were carried out using Particle Image
Velocimetry along two vertical planes along the flume
Table 1. Experimental conditions. axis, one plane through the centres of the spheres and
another through their connections. The results pre-
Experiment E1 E2 E3 sented below were obtained by averaging the results
for these two planes. More detailed description of the
Bed slope 1:400 1:400 1:200 experimental procedures can be found in (Manes et
Sphere diameter, k(mm) 12 12 12 al. submitted). Figure
Depth h(m)1.3 22 42 21   7 shows turbulent intensities
Maximum depth H (mm) 34 54 33 σu = u2 and σw = w2 measured in the experiments
Relative submergence (H /k) 2.83 4.5 2.75 E1, E2, E3, once scaled with the ‘roughness crest’
Shear velocity u∗ (m/s) 0.0235 0.0300 0.0301 shear velocity u∗0 and once with the ‘bed shear stress
Roughness Reynolds number 141 180 181 shear velocity u∗bed . The values of the shear stress
Average profile velocity (m/s) 0.210 0.340 0.286 at the roughness crest and the bed shear stress were
ubed /u∗ 1.13 1.07 1.14 determined by the extrapolation of the Reynolds stress

1 1

0.9 0.9

0.8 0.8

0.7 0.7

0.6 0.6

0.5 0.5
z/h

z/h

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 0.5 1 1.5 2 2.5 0 0.5 1 1.5
u`/u* w`/u*

Figure 7. Turbulent intensities σu and σw scaled with the ‘roughness crest’ shear velocity (grey symbols) and the ‘bed shear
stress’ shear velocity (white symbols) in experiments E1 (circles), E2 (squares) and E3 (triangles); σu is compared with
turbulent intensities in flow over smooth bed (Nezu & Rodi 1986, line).

96
profiles to the appropriate level (the roughness crest shown that scaling with the velocity scale based on
and the mean bed level z = −φk consequently). The the bed shear stress produces artificially increased von
dimensionless profiles of turbulent intensity collapse Karman constant and artificially suppressed turbulent
reasonably well between the experiments when scaled intensities. For shallow flows this can significantly
with the ‘roughness crest’ shear velocity. When scaled affect data interpretation.
with the ‘bed shear stress’ shear velocity, the turbulent
intensities for the two shallower experiments (circles
and triangles in Fig. 7) clearly stand out, showing the ACKNOWLEDGEMENTS
suppression of the turbulent intensities which is solely
the artefact of the choice of the velocity scale u∗ . It This investigation was carried out as part of an exper-
is interesting to compare the scaled turbulent inten- imental programme funded by the United Kingdom
sities σu with those for the smooth bed flow (Nezu Engineering and Physical Sciences Research Council
& Rodi, 1986), shown in the same figure. Judging (Grant GR/R51865/01).
by the data scaled with the ‘bed shear stress’ veloc-
ity scale, the turbulent intensities are suppressed in
shallow flow over rough bed compared to smooth bed, REFERENCES
while the data scaled with the ‘roughness crest’ veloc-
ity scale indicate that they are similar. As expected, in Bayazit, M. 1976. Free surface flow in a channel of large
the experiment E2 with deeper flow than in other two relative roughness, J. Hydraul. Res., 14(2): 115–126.
experiments, this difference vanishes. Dittrich, A., & Koll, K. 1997. Velocity field and resistance
These results illustrate how we can reach a differ- of flow over rough surface with large and small relative
submergence. Int. J. Sediment Res., 12(3): 21–33.
ent conclusion about turbulent intensities in rough bed Finnigan, J.J. 1985. Turbulent transport in flexible plant
shallow flows, solely because of the choice of the shear canopies. The forest-atmosphere interactions, B.A.
velocity with which to scale experimental data. For Hutchinson and B.B. Hicks, eds., Reidel, Dodrecht, The
flow over spheres with uniform diameter using the cor- Netherlands, pp. 443–480.
rect velocity scale reveals the similarity of turbulent Finnigan, J.J. 2000. Turbulence in plant canopies. Annu. Rev.
intensities between smooth-bed and rough-bed flows, Fluid. Mech., 32: 519–571.
while the inappropriate scale causes the non-existent, Jackson, P.S. 1981. On the displacement height in the
artificial, suppression of turbulent intensities. logarithmic velocity profile, J. Fluid Mech., 111:15–26.
Jarvela, J. 2005. Effect of submerged flexible vegetation on
flow structure and resistance. Jour. Hyd., 307: 233–241.
Katul, G., Wiberg, P., Albertson, J., & Hornberger, G. 2002. A
6 CONCLUSIONS mixing layer theory for flow resistance in shallow streams.
Water Resour. Res., 38(11):1250–1257.
The paper is focussed on the definition of the shear Kironoto, B.A., & Graf, W.H. 1994. Turbulence characteris-
tics in rough uniform open-channel flow. Proc., Instn. Civ.
velocity which is to be used, as universal velocity scale, Enrg. Water, Maritime, and Energy, 106: 333–344.
for scaling properties of free-surface turbulent flows Manes, C., Pokrajac, D., & McEwan, I. Double averaged open
over rough bed. In the literature there are several dif- channel flows with small relative submergence. Submitted
ferent definitions of the shear velocity, each with the to J. Hydraul. Eng., ASCE.
different associated shear stress. The two most com- Millikan, C.D. 1939. A critical discussion of turbulent flows
mon definitions are based on the shear stress at the in channels and circular tubes. in Proc. Fifth. Int. Conngr.
level of the roughness crest and the bed shear stress. Appl. Mech.: 386–392.
In deep flows, where the water depth is much larger Monin, A.S. & Yaglom, A.M. 1971. Statistical fluid mechan-
than the roughness height, the different definitions ics: Mechanics of turbulence, Vol. 1. MIT Press, Boston,
Mass.
give approximately equal values of the shear veloc- Nakagawa, H., Tsujimoto, T., & Shimizu, Y. 1991. Turbulent
ity. In shallow flows, however, the differences between flow with small relative submergence, in Lecture Notes in
these values can be significant. Earth Sciences, pp. 33–44.
This is the reason why for shallow rough bed flows Nepf, H.M., & Vivoni, E.R. 2000. Flow structure in depth-
the definition of the shear velocity requires closer limited, vegetated flow. J. Geophys. Res., 105(C12):
inspection. In this paper we present a case for adopt- 28547–28557
ing the shear velocity evaluated from the shear stress Nezu, I., & Rodi, W. 1986. Open-channel flow measure-
at the roughness crest as the universal velocity scale. ments with a laser Doppler anemometer, J. Hydraul. Eng.,
There are clear reasons against using the bed shear 112(5): 335–355.
Nikora, V.I., Goring, D.G., McEwan, I.K., & Griffiths, G.
stress, inasmuch as the shear velocity must be inter- 2001. Spatially averaged open-channel flow over rough
preted as the universal velocity scale of the free fluid bed. J. Hydraul. Eng., ASCE 127(2): 123–133.
flow in the outer layer. (The velocity scale calculated Nikora, V., Koll, K., McLean S., Dittrich, A., &
from the bed shear stress is the appropriate choice in Aberle, J. 2002. Zero-plane displacement for rough-bed
some other context e.g. sediment transport). It was open-channel flows. in Proceedings of the International

97
Conference on Fluvial Hydraulics River Flow 2002, Rowinski, P., Aberle, J., & Mazurczyk, A. 2005. Shear veloc-
September 4–6, 2002, Louvain-la-Neuve, Belgium, pp. ity estimation in hydraulic research. Acta Geophysica
83–92. Polonica 53(4): 567–583.
Nikora, V., Koll, K., McEwan, I., McLean, S., & Dittrich, A. Smith, J.D. & McLean, S.R. 1977. Spatially averaged flow
2004.Velocity distribution in the roughness layer of rough- over a wavy surface. J. Geophys. Res., 83(12): 1735–1746.
bed flows, J. Hydraul. Eng., ASCE, 130(7), 1036–1042. Wilson, N.R. & Shaw, R.H. 1977. A higher order clo-
Nikora, V., McEwan, I., McLean, S., Coleman, S., sure model for canopy flow. J. Appl. Meteorology, 16:
Pokrajac, D., & Walters, R. Double averaging concept for 1197–1205.
rough-bed open-channel and overland flows. Theoretical Wang, J., Dong, Z., Chen, C., & Xia, Z. 1993. The effects of
background, submitted to J. Hydraul. Eng., ASCE. bed roughness on the distribution of turbulent intensities
Poggi, D., Porporato, A., Ridolfi, L., Albertson, J.D., & Katul, in open channel flow, Jour. Hyd. Res. 31(1): 89–98.
G.G. 2004. The effect of vegetation density on canopy Wang, J., & Dong, Z. 1996. Open-channel turbulent flow over
sub-layer turbulence, Boundary-Layer Meteorology, 111: non-uniform gravel beds, Appl. Sci. Res 56: 243–254.
565–587. Yaglom,A.M. 1979. Similarity laws for constant-pressure and
Raupach, M.R. & Shaw, R.H. 1982. Averaging procedures pressure-gradient turbulent wall flows, Ann. Rev. Fluid
for flow within vegetation canopies. Boundary-Layer Mech., 11:505–540.
Meteorology, 22: 79–90.
Raupach, M.R., Antonia, R. A., & Rajagopalan, S. 1991
Rough-wall turbulent boundary layers.Applied Mech. Rev.
44: 1–24.

98

View publication stats

You might also like