Elementos de Turbulencia Etc.
Elementos de Turbulencia Etc.
Elementos de Turbulencia Etc.
ELEMENTS OF TURBULENCE
ALDO TAMBURRINO
Associate Professor Universidad de Chile
Visiting Professor Università della Campania Luigi Vanvitelli
INTRODUCTION
The motion of a fluid can be characterised by different regimes, depending
on the specific feature that we are interested in. Thus, if we are interested in the
flow dependency on the time, it could be steady (the flow does not change with
time) or unsteady (the flow depends on the time). If we are interested in the
changes in the space, it could be uniform (velocity and flow section do not
changes in space), or spatially varied. In the case of spatially varied flow, when
the variation is small, the flow is gradually varied. In this case, the curvature of
the streamlines is small. On the contrary, when the streamlines have strong
curvature, the flow is named rapidly varied. A property of the gradually varied
flows is that the pressure distribution can be considered as hydrostatic. There
are many other features of the flow that we could want to study, and we could
define another regimes. However, we will focus in the motion of the fluid
particles. That motion can be ordered, the flow moving in “layers”, and the flow
regime is called laminar. On the contrary, the motion of the particles can be
disorganised or random and the regime is named turbulent. Of course, there is
not an abrupt change from the laminar to the turbulent regime, but it goes
through a transition laminar-turbulent.
Osborne Reynolds published in 1883 an article in which he described the
motion of water in a tube in which he injected dye. The experimental facility
used by Reynolds and depicted in the article is shown in Fig. 1. As a curiosity,
Fig. 2 shows the original apparatus used by Reynolds, currently in the
University of Manchester.
For low water velocities, the dye injected in the upstream end of the tube
followed a well-defined linear trajectory, as shown in Fig.3 (taken from Reynold’s
paper). Reynolds called this motion “direct”. For higher velocities, after some
distance from the entrance of the tube, the dye “at once mix up with the
surrounding water, and fill the rest of the tube with a mass of coloured water”,as
shown in Fig. 4. But the most interesting feature was observed when the tube
:: 1 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
was illuminated with an “electric spark”: the dye showed “more or less distinct
curls, showing eddies”, as those sketched in Fig.5. Because of this eddy pattern,
Reynolds called this motion also “sinuous”. In the current language we call
:: 2 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
Presence of eddies in the flow was not something new when Reynolds
performed his experiment. The eddy motions in water flows were already
:: 3 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
two components, one given by the “impetus” of the main motion and another
associated to the eddies. Leonardo was the first one that used the term
turbolenza in about 1550.
Existence of eddies in the flow has an important effect in the flow
resistance. Before than Reynolds’ experiment was already known that for low
velocities (i.e., laminar regime), the flow resistance was proportional to the mean
velocity (or discharge), and for higher velocities (i.e., turbulent regime) it was
proportional to the square of the mean velocity.
WHAT IS TURBULENCE?
It is difficult to answer this question. As the velocity is continuously
changing (o fluctuating) in time, we can say that it is an unsteady phenomenon.
However, usually we are not interested in the instantaneous values of the
velocity (or pressure) of a turbulent flow, but we want to know some time
averaged values. Thus, the continuous variation in time of the flow properties
:: 4 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
Due to the complexity of the turbulent flows, most of what we know about
them comes from experiments. By means of experiments, using visualization
techniques or direct measurements of the variables that we are interested in
(usually, velocity and pressure), we can get a general description of the processes
involved in the turbulent flows. The use of the experimental approach does not
mean that theories have not been developed, or that are of lesser value than the
experiments. On the contrary, the theoretical analysis guide us regarding the
variables that need to be measured for a better understanding of the turbulence
phenomena. Obviously, as in all branches of science, the interaction between
theory and experiments is necessary, and both of them help to increase our
understanding of the turbulence. It is important to remark that turbulence still
is an open problem. That is to say that currently the problem has not been solved
starting from the principles of the physics. Thus, any theory on turbulence will
always rely on some experimental data.
Up to now, there is not a definition of turbulence that can describe it
completely. Von Karman (1937) defined it as “an irregular motion which in
general makes its appearance in fluids, gaseous or liquid, when they flow past
solid surfaces or even when neighbouring streams of the same fluid flow past or
over one another”. According to Hinze (1959) “turbulent fluid motion is an
irregular condition of the flow in which the various quantities show a random
variation with time and space coordinates, so that statistically distinct average
values can be discerned”. This definition has a couple of consequences: the
:: 5 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
- Italian
o Dizionario Italiano (www.dizionario-italiano.it): “Fenomeno per cui
in un fluido liquido o gassoso in moto, in determinate condizioni, il
moto delle particelle elementari cessa di essere regolare e diventa
soggetto a forti fluttuazioni della velocità e a moti vorticosi e a
mancanza di regolarità nella traiettoria delle particelle”
o Grande Dizionario Italiano di Gabrielli Aldo
(www.grandidizionari.it): “Moto irregolare generalmente rilevabile
nei fluidi”
o Dizionari di Italiano Corriere della Sera
(http://dizionari.corriere.it/dizionario_italiano): “Moto disordinato
di un fluido, con formazione di vortici”
- Spanish
o Diccionario de la Real Academia Española (www.rae.es): “Zona en
que se desarrolla un movimiento turbulento”.
o Turbulento: “Dicho del movimiento: Propio de un fluido en el
que la presión y la velocidad fluctúan muy irregularmente en
cada punto, con la consiguiente formación de remolinos.”
As it can be seen, even the definition of the word turbulence in different
languages and dictionaries is not unique. However, from in the definitions
presented before we can group words that transmit more or less the same
meaning. For example, we find:
- “Violent or unsteady movement”, “forti fluttuazioni della velocità”,
- “Il moto delle particelle elementari cessa di essere regolare”, “Moto
irregolare”, “Moto disordinato”, “la presión y la velocidad fluctúan muy
irregularmente”
:: 6 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
:: 7 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝜕𝑉
= 𝑄𝑖 − 𝑄𝑜 (1)
𝜕𝑡
In Eq. 1, 𝑉 represents the volume of the fluid inside the control volume, 𝑡
is the time, 𝑄 is the volume rate or discharge, the sub-index 𝑖 indicates the input
to the control volume and the sub-index 𝑜 the output.
If the discharge 𝑄 goes through a section with an area normal to the flow
equal to 𝐴, the average velocity is defined as
𝑄
𝑣= (2)
𝐴
In a steady state flow there is not variation in time, 𝜕𝑉⁄𝜕𝑡 = 0, and Eq. 1
is simplified to:
𝑄𝑖 = 𝑄𝑜 (3)
Eq. 3 indicates that the flow that enters to the control volume is equal to
the output flow.
:: 8 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑉𝐶
The classical example where Eq. 1 is applied corresponds to the flow through
and orifice in a tank. Referred to Fig. 7, the question is to determine the variation
of the water depth (ℎ) in the tank in function of the time, if initially the flow
depth is ℎ0 . The transverse area of the tank is 𝐴𝑇 and the area of the orifice is
𝐴𝑜 . This simple example is repeated here to stress two isues:
i) The volume of fluid not necessarily is equal to the control volume,𝑉𝐶 ,
depicted as the segmented line in Fig. 7.
ii) Eq. 1 by itself is not enough to solve the problem.
The volume of fluid (𝑉) in the control volume 𝑉𝐶 is 𝑉 = ℎ𝐴𝑇 . There is not
flow rate entering to 𝑉𝐶 , so 𝑄𝑖 = 0, and the flow rate exiting the control volume
is 𝑄𝑜 = 𝑣𝐴𝑜 . Thus, Eq. 1 can be written as:
𝜕
(ℎ𝐴𝑇 ) = −𝑣𝐴𝑜 (4)
𝜕𝑡
𝑣 = √2𝑔ℎ (5)
:: 9 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝜕ℎ 𝐴𝑜
= − √2𝑔ℎ (6)
𝜕𝑡 𝐴𝑇
∆𝑊
∆𝐸
∆𝑄̂
𝑑𝐸 𝑑𝑄̂ 𝑑𝑊
= − (7)
𝑑𝑡 𝑑𝑡 𝑑𝑡
𝐸 = 𝑈 + 𝐸𝑃 + 𝐸𝐶 (8)
where 𝑈 is the internal energy, 𝐸𝑃 is the potential energy and 𝐸𝐶 is the kinetic
energy. They are expressed as:
1
𝑈 = 𝑚𝑢 𝐸𝑃 = 𝑚𝑔𝑧 𝐸𝐶 = 𝑚𝑣 2 (9)
2
where 𝑚 is the mass, 𝑢 is the specific internal energy, 𝑔 is the acceleration due
to gravity, 𝑧 is the vertical distance from an arbitrary reference level, and 𝑣 the
:: 10 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
velocity. The specific energy (defined as energy per unit mass, 𝑒 = 𝐸 ⁄𝑚), of the
system is:
𝑣2
𝑒 = 𝑢 + 𝑔𝑧 + (10)
2
It can be shown that the variation of energy of the system can be expressed
as the change of energy per unit time of the fluid in the control volume, plus the
flow of energy through the surfaces𝑆𝐶 that define 𝑉𝐶 . Mathematically, it is
written as:
𝜕 𝑑𝑄̂ 𝑑𝑊
∫ 𝜌𝑒𝑑𝑉 + ∫ 𝜌𝑒𝑣⃗ ∙ 𝑛̂𝑑𝑆 = − (11)
𝜕𝑡 𝑉𝐶 𝑆𝐶 𝑑𝑡 𝑑𝑡
The work 𝑊 can be divided in two main components: 1) the work that the
fluid has to do when it flows (it has to overcome the forces arising from the
pressure and shear stresses), and 2) the external work (also called “shaft work”).
The external work is that done to move the shaft of a turbine (work towards the
exterior of the system, work done by the fluid, positive), or that supplied by a
pump (work towards the interior of the system, work done on the fluid,
negative). Thus:
𝑊 = 𝑊𝐸 + 𝑊𝑝 + 𝑊𝜏 (12)
where the sub-indices 𝐸, 𝑝 and 𝜏 stand for “external”, “pressure” and “shear
stresses”, respectively.
It is possible to work with the term associated to the pressure:
𝑑𝑊𝑝 𝑑𝑟⃗
= ∫ 𝑝𝑑𝑆𝑛̂ ∙ = ∫ 𝑝𝑣⃗ ∙ 𝑛̂𝑑𝑆 (14)
𝑑𝑡 𝑑𝑡 𝑆𝐶
Introducing the expression for the specific energy 𝑒 in the second term of the left
hand side of Eq. 15:
:: 11 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
We can see that the sum of Bernuilli was formed in the second term of Eq.
17, and it can be written as:
𝜕 𝑢 𝑑𝑄̂ 𝑑𝑊𝐸 𝑑𝑊𝜏
∫ 𝜌𝑒𝑑𝑉 + ∫ 𝜌𝑔 ( + 𝐵) 𝑣⃗ ∙ 𝑛̂𝑑𝑆 = − − (18)
𝜕𝑡 𝑉𝐶 𝑆𝐶 𝑔 𝑑𝑡 𝑑𝑡 𝑑𝑡
where
𝑝 𝑣2
𝐵=𝑧+ + (19)
𝜌𝑔 2𝑔
Eq. 17 (or Eq. 18) is called the “energy general equation”. It can be applied
to any flow regime. Its only limitation is given by the potential energy should
derive from a gravitational field (Eq. 9).
Let’s simplify Eq. 18 assuming steady flow and considering a stream tube
as control volume, as shown in Fig. 9. The control volume has three surfaces:
section 1 (entrance), section 2 (exit) and the mantle, with surface areas 𝑆1, 𝑆2 and
𝑆𝑚 , respectively. Thus, the surface 𝑆𝐶 that defines the control volume 𝑉𝐶 can be
written as 𝑆𝐶 = 𝑆1 + 𝑆2 + 𝑆𝑚 . As the mantle is tangent to the velocity vectors,
there is not flow through its surface, and the integral over the surface of Eq. 18
is reduced to:
𝑢 𝑢 𝑢
∫ 𝜌𝑔 ( + 𝐵) 𝑣⃗ ∙ 𝑛̂𝑑𝑆 = ∫ 𝜌𝑔 ( + 𝐵) 𝑣⃗ ∙ 𝑛̂𝑑𝑆 + ∫ 𝜌𝑔 ( + 𝐵) 𝑣⃗ ∙ 𝑛̂𝑑𝑆 (20)
𝑆𝐶 𝑔 𝑆1 𝑔 𝑆2 𝑔
For simplicity, we can take that the velocity vector at sections 1 and 2 is
normal to the surface, that is to say that in section 1: 𝑣⃗ = 𝑣1 (−𝑛̂1 ), and in section
2: 𝑣⃗ = 𝑣2 𝑛̂2 . Thus, Eq. 20 becomes:
𝑢 𝑢 𝑢
∫ 𝜌𝑔 ( + 𝐵) 𝑣⃗ ∙ 𝑛̂𝑑𝑆 = − ∫ 𝜌𝑔 ( + 𝐵) 𝑣1 𝑑𝑆 + ∫ 𝜌𝑔 ( + 𝐵) 𝑣2 𝑑𝑆 (21)
𝑆𝐶 𝑔 𝑆1 𝑔 𝑆2 𝑔
:: 12 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑆2
𝑛̂2
𝑣⃗2
𝑆1
𝑆𝑚
𝑛̂1
SECTION 2
𝑣⃗1
SECTION 1
𝑢 𝑢1 𝑢2
∫ 𝜌𝑔 ( + 𝐵) 𝑣⃗ ∙ 𝑛̂𝑑𝑆 = −𝜌𝑔 ( + 𝐵1 ) 𝑣1 𝑆1 + 𝜌𝑔 ( + 𝐵2 ) 𝑣2 𝑆2 (23)
𝑆𝐶 𝑔 𝑔 𝑔
Thus, for a steady flow through a stream tube with one entrance and one
exit (as in Fig. 9), the general equation of the energy is written as:
𝑢2 − 𝑢1 𝑑𝑄̂ 𝑑𝑊𝐸 𝑑𝑊𝜏
𝜌𝑔𝑄 ( + 𝐵2 − 𝐵1 ) = − − (25)
𝑔 𝑑𝑡 𝑑𝑡 𝑑𝑡
:: 13 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
For adiabatic flows (no heat exchange) of ideal fluids without work done
by the fluid, we obtain that the Bernoulli remains constant. :
𝐵2 = 𝐵1 = 𝐶𝑜𝑛𝑠𝑡. (27)
𝑝 𝑣2 𝑝 𝑣2
∫ 𝜌𝑔 (𝑧 + + ) 𝑣𝑑𝑆 = 𝜌𝑔 (𝑧 + ) ∫ 𝑣𝑑𝑆 + 𝜌𝑔 ∫ 𝑣𝑑𝑆 (30)
𝑆 𝜌𝑔 2𝑔 𝜌𝑔 𝑆 𝑆 2𝑔
𝑝𝐺 𝑣2
∫ 𝜌𝑔𝐵𝑣𝑑𝑆 = 𝜌𝑔 (𝑧𝐺 + ) ∫ 𝑣𝑑𝑆 + 𝜌𝑔 ∫ 𝑣𝑑𝑆 (31)
𝑆 𝜌𝑔 𝑆 𝑆 2𝑔
:: 14 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
We want to have:
Replacing Eq. 31 in the left side of Eq. 33 and Eq. 32 in the integral of the
right side:
𝑝𝐺 𝑣2 𝑝 𝑣̅ 2
𝜌𝑔 (𝑧𝐺 + ) ∫ 𝑣𝑑𝑆 + 𝜌𝑔 ∫ 𝑣𝑑𝑆 = ∫ 𝜌𝑔 (𝑧 + + 𝛼 ) 𝑣̅ 𝑑𝑆 (34)
𝜌𝑔 𝑆 𝑆 2𝑔 𝑆 𝜌𝑔 2𝑔
𝑝𝐺 𝑣2
𝜌𝑔 (𝑧𝐺 + ) ∫ 𝑣𝑑𝑆 + 𝜌𝑔 ∫ 𝑣𝑑𝑆
𝜌𝑔 𝑆 𝑆 2𝑔
𝑝𝐺 𝑣̅ 2 (35)
= 𝜌𝑔 (𝑧𝐺 + ) ∫ 𝑣̅ 𝑑𝑆 + 𝜌𝑔 ∫ 𝛼 𝑣̅ 𝑑𝑆
𝜌𝑔 𝑆 𝑆 2𝑔
But
∫ 𝑣𝑑𝑆 = ∫ 𝑣̅ 𝑑𝑆 = 𝑄 (36)
𝑆 𝑆
From where we can obtain an expression for the coefficient 𝛼 that takes
into account that the velocity is not constant in any section of the stream tube:
∫𝑆 𝑣 3 𝑑𝑆 (38)
𝛼=
𝑣̅ 3 𝐴
The coefficient 𝛼 is called Coriolis coefficient. It is easy to see that in order
to compute 𝛼 from Eq. 38, the velocity distribution has to be known, which is not
always possible, and many times it is determined from experiments or field
measurements. For turbulent flow in rectilinear pipes ~ 1.03 − 1.1 , depending
on 𝑅𝑒. As 𝛼 is close to 1, the value 𝛼 = 1 is commonly used. For a laminar flow
in a pipe, 𝛼 = 2.
:: 15 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
with
𝑝 𝑣̅ 2
∗
𝐵 =𝑧+ +𝛼 (32)
𝜌𝑔 2𝑔
However, it is customary that the bar ̅ over the velocity and the asterisk
∗
of the Bernoulli are not written, and Eq. 32 becomes:
𝑝 𝑣2
𝐵=𝑧+ +𝛼 (40)
𝜌𝑔 2𝑔
A word of caution is necessary here: Eq. 41 is easily confused with Eq. 19,
although they are different. Eq. 41 can be applied to a real fluid with non-
uniform velocity profile, whereas Eq. 19 is restricted to uniform velocity profile,
something that cannot be attained by flows of real fluids with solid boundaries.
Following the common usage, we will write 𝐵, although we will be working
with𝐵 ∗ . Similarly, most of the time we will not write the Coriolis coefficient
because the flows will be turbulent.
If there is a hydraulic machine like a pump or a turbine, 𝑑𝑊𝐸 ⁄𝑑𝑡
corresponds to the power of the machine, 𝑃. We use to write the power divided
by (𝜌𝑔𝑄). For an ideal flow without heat exchange:
1 𝑑𝑊𝐸 𝑃
(𝐵2 − 𝐵1 ) = − =− (42)
𝜌𝑔𝑄 𝑑𝑡 𝜌𝑔𝑄
:: 16 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
1 𝑑𝑊𝜏
(𝐵2 − 𝐵1 ) = − (43)
𝜌𝑔𝑄 𝑑𝑡
The dissipated energy per unit weight of the fluid is denoted by, which
is defined as:
1 𝑑𝑊𝜏
≡ (44)
𝜌𝑔𝑄 𝑑𝑡
𝐵2 = 𝐵1 − (45)
The energy loss is usually divided in two terms: one associated to friction
and other to singularities in the flow line. Let’s restrict our attention to the
friction loss. The energy loss per unit length is denoted by 𝐽, defined as:
𝑑𝐵
𝐽=− (46)
𝑑𝑥
Thus, the friction loss in a pipe of length 𝐿 is given by = 𝐽𝐿.Computation
of 𝐽 is generally performed using the Darcy-Weisbach equation, which requires
to know the friction factor 𝑓. For a cylindrical pipe of diameter 𝐷:
1 𝑣2
𝐽=𝑓 (47)
𝐷 2𝑔
For simplicity, let’s apply Eq. 48 to the same control volume used in the
derivation of the energy general equation (Fig. 9). It can be shown that the term
:: 17 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
that indicates the variation of momentum in Eq. 48, when apply to a fluid,
becomes:
𝑑 𝜕
(𝑚𝑣⃗) = ∫ 𝜌𝑣⃗ 𝑑𝑉 + ∫ 𝜌𝑣⃗ 𝑣⃗ ∙ 𝑛̂𝑑𝑆 (49)
𝑑𝑡 𝜕𝑡 𝑉𝐶 𝑆𝐶
The integral over the surfaces of the control volume is split in three terms:
Considering that there is not flow through the mantle and that 𝑣⃗1 ∙ 𝑛̂1 < 0,
𝑣⃗2 ∙ 𝑛̂2 > 0, constant properties at each transverse section of the stream tube, and
𝑣1 𝑆1 = 𝑣2 𝑆2 = 𝑄, Eq. 48 can be written as:
𝜕
∫ 𝜌𝑣⃗ 𝑑𝑉 + 𝜌𝑄(𝑣⃗2 − 𝑣⃗1 ) = 𝐹⃗ (51)
𝜕𝑡 𝑉𝐶
As it happened with the energy general equation, Eq. 52 is valid only for
uniform velocity profiles. In order to take into account the non-uniformity of the
velocity profile imposed by boundaries in flow of real fluids, the cross section
mean velocity is used in conjunction with a coefficient 𝛽 and Eq. 53 becomes:
∫𝑆 𝑣 2 𝑑𝑆 (53)
𝛽=
𝑣̅ 2 𝐴
For most turbulent flows 𝛽 ~ 1.01 − 1.04, and the value 𝛽 = 1 is
considered. For a laminar flow in a pipe, 𝛽 = 4⁄3.
The uniform, steady, two-dimensional flow with free surface of a real
fluid: As an application of the three fundamental principles of the physics
applied to hydraulics, we will analyse the flow in an infinitely wide inclined
channel in which the permanent and uniform flow with a free surface exists. The
problem is sketched in Fig. 10. The bottom of the channel coincides with the 𝑥
axis which is inclined an angle 𝜃 with respect to the horizontal line. The flow
:: 18 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
depth is 𝐻 and the cross sectional mean velocity is 𝑈. There is not heat transfer
between the fluid sustem and the environment and the flow is turbulent.
The first thing that needs to be done when working with the equations
derived with the integral approach, is to choose the control volume. In this case
𝑉𝐶 corresponds to the volume defined by the segmented line. The application of
the continuity equation, energy equation and momentum theorem must be done
in this volume.
𝑝𝑎𝑡𝑚 𝑔⃗
𝑉𝐶
𝜃
Fig. 10.- Steady, uniform, 2-D free surface
flow
Continuity equation: As the flow is 2-D, we work with both the discharge and the
area of the flow section per unit width, i.e., 𝑞 = 𝑄 ⁄𝑏 and 𝑎 = 𝐴⁄𝑏. Calling (1) to
the entrance section and (2) to the exit section, and considering steady flow, the
application of Eq. 1 to the control volume reduces to 𝑞1 = 𝑞2 . Uniformity of the
flow means that the flow depth in both sections is the same, from where 𝑎1 =
𝑎2 = 𝐻. Thus, from Eq. 2: 𝑣1 = 𝑣2 = 𝑈.
Energy equation: The control volume corresponds to a steamtube, as shown in
Fig. 11. As the flow is turbulent we assume 𝛼 = 1. Application of Eq. 45:
𝐵2 = 𝐵1 − (45)
(1)
(2)
𝑧1
𝑧2
:: 19 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
As the flow is uniform, the streamlines are parallels, and the sum 𝑧 +
𝑝⁄(𝜌𝑔) remains constant in any cross section. That means that the sum can be
evaluated at any 𝑧 of the section. For simplicity, we choose to evaluate it at the
free surface, because at that location the pressure is known (𝑝𝑎𝑡𝑚 ). Thus, Eq. 45
becomes:
𝑝𝑎𝑡𝑚 𝑈 2 𝑝𝑎𝑡𝑚 𝑈 2
𝑧2 + + = 𝑧1 + + − (54)
𝜌𝑔 2𝑔 𝜌𝑔 2𝑔
Thus:
= 𝑧1 − 𝑧2 (55)
𝐽 = sin 𝜃 (56)
The result above obtained indicates that for steady, uniform open-channel
flows the gradient of the energy line is equal to the slope of the channel. Or,
equivalent, the energy line is parallel to the bottom of the channel (and parallel
to the free surface).
We can relate the flow velocity with the slope of the channel using Eq. 47.
It is easy to show that for this flow, 𝑅𝐻 = 𝐻. Thus, we have:
1 𝑈2
sin 𝜃 = 𝑓 (57)
4𝐻 2𝑔
The problem is not solved yet because we have not said anything about
the friction factor. We will get some relationships for it later.
Momentum theorem: In order to apply Newton’s second law to the control volume
of Fig. 10, we have to recognize that Eq. 52 is a vectorial equation, and it should
be applied to each component of the coordinate system. As we already have
defined the control volume, the first step in our analysis is to identify the forces
and momentum fluxes acting in 𝑉𝐶 . To do this, we will use Fig. 12.
:: 20 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
(1)
(2)
𝑔⃗
𝐹𝑝1
𝑈 𝐹𝑝2
𝑈
𝐹𝜏𝑜
𝑊
:: 21 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑊 = 𝜌𝑔𝐻∆𝑥 (62)
𝐹𝜏0 = 𝜏0 ∆𝑥 (63)
In the above equation, 𝜏0 is the shear stress acting on the bottom of the
channel. Eqs. 61, 62 and 63 give:
𝜏0 = 𝜌𝑔𝐻𝐽 (65)
Although Eq. 65 was obtained for steady and uniform flow, the same result
is obtained for steady and gradually varied flow.
Combining Eqs. 57 and 64:
1
𝜏0 = 𝜌𝑓𝑈 2 (66)
8
From where:
𝑢∗ 𝑓
=√ (68)
𝑈 8
:: 22 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑝𝑎𝑡𝑚 𝑔⃗
𝑉𝐶 𝜃
Fig. 13.- Control volume chosen to determine the variation of the shear
stress
The forces acting in 𝐶𝑉 are depicted in Fig. 14. Essentially, they are the
same than those shown in Fig. 12, but for a control volume that does not reach
the bottom of the channel.
(1)
(2)
𝐹𝑝𝑦1 𝑔⃗
𝑈 𝐹𝑝𝑦2
𝐹𝜏𝑦 𝑈
𝑊𝑦
𝐹𝑝𝑦 is the force due to the fluid pressure acting on the surfaces comprised
between 𝐻 and 𝑦, 𝑊𝑦 is the weight of the fluid contained in the control volume,
and 𝐹𝜏𝑦 the force due to friction at a distance 𝑦 from the bottom. With the same
arguments that we obtained Eq. 61, we get:
:: 23 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
The weight of fluid and force due to shear stress are given by:
𝐹𝜏𝑦 = 𝜏𝑦 ∆𝑥 (71)
Eq. 72 indicates that the shear stress varies linearly with depth, from zero
on the free surface (𝑦 = 𝐻) to 𝜏𝑦 = 𝜌𝑔𝐻 sin 𝜃 on the bottom (𝑦 = 0), as sketched in
Fig. 15. Dividing Eq. 72 by Eq.64:
𝜏𝑦 𝑦
= (1 − ) (73)
𝜏0 𝐻
𝜏0 𝜏𝑦
Fig. 15.- Shear stress distribution
:: 24 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
CONTINUITY EQUATION
Let’s analyse the flow of mass through an element of volume 𝑑𝑉 = 𝑑𝑥𝑑𝑦𝑑𝑧
immersed in the flow, as shown in Fig. 16. The velocity field is 𝑣⃗ = (𝑢, 𝑣, 𝑤).
Through this imaginary volume, the flow (represented by the streamlines in the
figure) passes transporting mass of fluid. Conservation of mass indicates that
the variation of mass per unit time inside of 𝑑𝑉 is equal to the mass rate that
enters to the volume, less that the mass rate that exit it. This statement can be
𝑑𝑧
𝑦
𝑑𝑦
𝑥 𝑑𝑥
written as:
𝜕𝑚
= 𝐺𝑖 − 𝐺𝑜 (74)
𝜕𝑡
where 𝑚 is the mass of fluid contained in 𝑑𝑉, given by
:: 25 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑑𝑚 = 𝜌𝑑𝑉 (75)
𝐺𝑖 is the mass rate that enters to 𝑑𝑉, and 𝐺𝑜 is the mass rate that exits the
volume. The fluid can enters (and exits) through any of the six surfaces that
define the volume. For the flow in the 𝑥 direction, we have a surface located at 𝑥
with normal (−𝑖̂), and other located at 𝑥 + 𝑑𝑥, with normal (+𝑖̂).The same
happens for the other coordinates. Thus, we can decompose the mass that enters
to 𝑑𝑉 in three terms, and write:
𝐺𝑖 = 𝐺𝑥 |𝑥 + 𝐺𝑦 |𝑦 + 𝐺𝑧 |𝑧 (76)
where the symbol |𝑥 stands for “evaluated at 𝑥”, and so on for the other
directions. Similarly, the mass that exits the volume can be written as:
𝜕
(𝜌𝑑𝑉) = 𝐺𝑥 |𝑥 + 𝐺𝑦 | + 𝐺𝑧 |𝑧 − (𝐺𝑥 |𝑥+𝑑𝑥 + 𝐺𝑦 | + 𝐺𝑧 |𝑧+𝑑𝑧 ) (78)
𝜕𝑡 𝑦 𝑦+𝑑𝑦
𝜕𝐺𝑥
𝐺𝑥 |𝑥 − 𝐺𝑥 |𝑥+𝑑𝑥 = − | 𝑑𝑥 (80)
𝜕𝑥 𝑥
The mass rate is defined as the discharge times density. The discharge in
the 𝑥 direction is 𝑢𝑑𝐴𝑥 , where 𝑑𝐴𝑥 is the element of area of the surface with
normal 𝑖̂, 𝑑𝐴𝑥 = 𝑑𝑦𝑑𝑧. Thus, Eq. 80 becomes:
𝜕𝐺𝑥 𝜕
− | 𝑑𝑥 = − (𝜌𝑢)| 𝑑𝑥𝑑𝑦𝑑𝑧 (81)
𝜕𝑥 𝑥 𝜕𝑥 𝑥
:: 26 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝜕𝐺𝑦 𝜕 𝜕𝐺𝑧 𝜕
− | 𝑑𝑦 = − (𝜌𝑣)| 𝑑𝑥𝑑𝑦𝑑𝑧 − | 𝑑𝑧 = − (𝜌𝑤)| 𝑑𝑥𝑑𝑦𝑑𝑧 (82)
𝜕𝑦 𝑦 𝜕𝑦 𝑦
𝜕𝑧 𝑧 𝜕𝑧 𝑧
𝜕𝜌 𝜕 𝜕 𝜕
𝑑𝑥𝑑𝑦𝑑𝑧 = − ( (𝜌𝑢) + (𝜌𝑣) + (𝜌𝑤)) 𝑑𝑥𝑑𝑦𝑑𝑧 (83)
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧
𝜕𝜌 𝜕 𝜕 𝜕
+ (𝜌𝑢) + (𝜌𝑣) + (𝜌𝑤) = 0 (84)
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧
𝜕𝜌
+ ∇ ∙ (𝜌𝑣⃗) = 0 (85)
𝜕𝑡
For an incompressible fluid, Eq. 85 is greatly reduced to:
∇ ∙ 𝑣⃗ = 0 (86)
Or, equivalently:
𝜕𝑢 𝜕𝑣 𝜕𝑤
+ + =0 (87)
𝜕𝑥 𝜕𝑦 𝜕𝑧
MOMENTUM EQUATION
Before applying the Newton’s second law to the elementary volume of Fig.
16, it is important to remember that two kind of forces can be identified in the
element of fluid: Forces that depend on the amount of matter and that their
application point is the centre of gravity of the element of fluid (for example,
weight). They are called “body forces”, and notated as 𝐹⃗𝐵 . The second kind of
forces act on the surfaces that define the volume of fluid (like pressure and shear
stresses), and “surface forces”, and written as 𝐹⃗𝑆 . Thus, Newton’s second law can
be written as:
:: 27 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑑
(𝑚𝑣⃗) = 𝐹⃗𝑆 + 𝐹⃗𝐵 (88)
𝑑𝑡
As the total mass of the system remains constant, Eq. 88 evolves into:
𝑑𝑣⃗
𝜌𝑑𝑉 = 𝐹⃗𝑆 + 𝐹⃗𝐵 (89)
𝑑𝑡
The most common body force is that due to gravity. In this case, 𝑓⃗𝐵 = 𝑔⃗,
and Eq. 92 becomes:
The analysis of surface forces requires to evaluate the forces acting on the
six surfaces that define de element of fluid volume, as shown in Fig. 17. There
are three surface forces acting on each surface. The surface stresses are
represented as 𝜏𝑖𝑗 , where 𝑖 indicates the direction of the vector normal to the
surface and 𝑗 the direction of the force (where 𝑖 and 𝑗 can take the values 𝑖̂, 𝑗̂ or
𝑘̂ ). Thus, the net force in the direction 𝑗 is the result of the stresses of the
surfaces located at 𝑥, with normal (−𝑖̂); at 𝑥 + 𝑑𝑥, with normal (+𝑖̂); at 𝑦, with
normal (−𝑗̂); at 𝑦 + 𝑑𝑦, with normal (+𝑗̂); and at 𝑧, with normal (−𝑘̂); at 𝑧 + 𝑑𝑧,
with normal (+𝑘̂). Obviously, the direction 𝑗 can be 𝑖̂, 𝑗̂ or 𝑘̂.
The net surface force has components in the three directions:
As it was said, six surfaces contribute to each component of Eq. 94. For
the component in the 𝑥 direction we have:
:: 28 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝐹𝑆𝑥 = 𝐹𝑆𝑥𝑥 |𝑥 + 𝐹𝑆𝑥𝑥 |𝑥+𝑑𝑥 + 𝐹𝑆𝑦𝑥 |𝑦 + 𝐹𝑆𝑦𝑥 |𝑦+𝑑𝑦 + 𝐹𝑆𝑧𝑥 |𝑧 + 𝐹𝑆𝑧𝑥 |𝑧+𝑑𝑧 (95)
In term of the stresses, any surface force can be written as 𝐹𝑆𝑖𝑗 = 𝜏𝑖𝑗 𝑑𝑆⃗𝑗 ,
where 𝑑𝑆⃗𝑗 is the surface on which the stress is acting, with a vector normal to
the surface in the 𝑗 direction. Thus, Eq. 95 is given by:
𝜏𝑧𝑧
𝜏𝑧𝑦
𝜏𝑧𝑥 𝜏𝑥𝑧
𝜏𝑦𝑧 𝜏𝑥𝑦
𝑧 𝜏𝑥𝑥
𝑑𝑧
𝜏𝑦𝑦 𝜏𝑦𝑥
𝑦
𝑑𝑦
𝑥 𝑑𝑥
𝐹𝑆𝑥 = 𝜏𝑥𝑥 |𝑥 (−𝑖̂)𝑑𝑆𝑥 + 𝜏𝑥𝑥 |𝑥+𝑑𝑥 (𝑖̂)𝑑𝑆𝑥 + 𝜏𝑦𝑥 |𝑦 (−𝑗̂)𝑑𝑆𝑦 + 𝜏𝑦𝑥 |𝑦+𝑑𝑦 (𝑗̂)𝑑𝑦
(96)
+ 𝜏𝑧𝑥 |𝑧 (−𝑘̂)𝑑𝑆𝑧 + 𝜏𝑧𝑥 |𝑧+𝑑𝑧 (𝑘̂)𝑑𝑆𝑧
Expanding in Taylor’s series and neglecting the terms of second order and
higher, we have:
:: 29 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝜕𝜏𝑥𝑥
𝜏𝑥𝑥 |𝑥+𝑑𝑥 = 𝜏𝑥𝑥 |𝑥 + | 𝑑𝑥 + ⋯
𝜕𝑥 𝑥
𝜕𝜏𝑦𝑥
𝜏𝑦𝑥 |𝑦+𝑑𝑦 = 𝜏𝑦𝑥 |𝑦 + | 𝑑𝑦 + ⋯ (98)
𝜕𝑦 𝑦
𝜕𝜏𝑧𝑥
𝜏𝑧𝑥 |𝑧+𝑑𝑧 = 𝜏𝑧𝑥 |𝑧 + | 𝑑𝑧 + ⋯
𝜕𝑧 𝑧
Following the same procedure for the components along 𝑦 and 𝑧 of 𝐹⃗𝑆 , we
obtain:
𝜕𝜏𝑥𝑦 𝜕𝜏𝑦𝑦 𝜕𝜏𝑧𝑦
𝐹𝑆𝑦 = ( + + ) 𝑑𝑥𝑑𝑦𝑑𝑧 (100)
𝜕𝑥 𝜕𝑦 𝜕𝑧
Replacing Eqs. 91, 93, 99, 100 and 101 in Eq. 94, Newton’s second law is
written as:
We can recognize that 𝜏𝑥𝑥 , 𝜏𝑦𝑥 , 𝜏𝑧𝑥 , 𝜏𝑥𝑦 , … are the elements of matrix array.
Actually, they are a tensor, and 𝜏𝑖𝑗 corresponds to the stress tensor:
:: 30 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
As Eq. 104 is vectorial, we can write three equations, one for each
component.
Component 𝑥:
𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝜏𝑥𝑥 𝜕𝜏𝑦𝑥 𝜕𝜏𝑧𝑥
𝜌( +𝑢 +𝑣 + 𝑤 ) = 𝜌𝑔𝑥 + + +
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑥 𝜕𝑦 𝜕𝑧 (105)
Component 𝑦:
𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝜏𝑥𝑦 𝜕𝜏𝑦𝑦 𝜕𝜏𝑧𝑦
𝜌( +𝑢 +𝑣 + 𝑤 ) = 𝜌𝑔𝑦 + + +
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑥 𝜕𝑦 𝜕𝑧 (106)
Component 𝑧:
𝜕𝑤 𝜕𝑤 𝜕𝑤 𝜕𝑤 𝜕𝜏𝑥𝑧 𝜕𝜏𝑦𝑧 𝜕𝜏𝑧𝑧
𝜌( +𝑢 +𝑣 +𝑤 ) = 𝜌𝑔𝑧 + + +
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑥 𝜕𝑦 𝜕𝑧 (107)
:: 31 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
Note that Eq. 113 gives an interesting result: the mechanical pressure, 𝑝̅ ,
is not equal to the thermodynamic pressure, 𝑝. An interesting issue is raised
with the second coefficient of viscosity. In his work, Stokes assumed that
(2𝜇 ⁄3 + ) = 0, meaning that < 0. However, some measurements indicate that
> 0. Nevertheless, the value of should not bother us because for
incompressible fluids ∇ ∙ 𝑣⃗ = 0 (Eq. 86), and in this case 𝑝̅ = 𝑝. Also, for
uncompressible fluids, Eq. 110 is reduced to:
𝜕𝑢𝑖 𝜕𝑢𝑗
𝜏𝑖𝑗 = −𝑝𝛿𝑖𝑗 + 𝜇 ( + ) (114)
𝜕𝑥𝑗 𝜕𝑥𝑖
Thus, the nine components of the tensor 𝜏𝑖𝑗 (Eq. 103) are:
:: 32 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝜕𝑢
𝜏𝑥𝑥 = −𝑝 + 𝜇2 (115)
𝜕𝑥
𝜕𝑣
𝜏𝑦𝑦 = −𝑝 + 𝜇2 (116)
𝜕𝑦
𝜕𝑤
𝜏𝑧𝑧 = −𝑝 + 𝜇2 (117)
𝜕𝑧
𝜕𝑢 𝜕𝑣
𝜏𝑥𝑦 = 𝜏𝑦𝑥 = 𝜇 ( + ) (118)
𝜕𝑦 𝜕𝑥
𝜕𝑢 𝜕𝑤
𝜏𝑥𝑧 = 𝜏𝑧𝑥 = 𝜇 ( + ) (119)
𝜕𝑧 𝜕𝑥
𝜕𝑣 𝜕𝑤
𝜏𝑦𝑧 = 𝜏𝑧𝑦 = 𝜇 ( + ) (120)
𝜕𝑧 𝜕𝑦
The terms containing 𝜇 are the viscous stresses. Introducing Eqs. 115 to 120 in
Eqs.105 to 107 and using Eq. 87, we obtain the momentum equations for an
incompressible Newtonian fluid with constant viscosity in the gravitational field:
Component 𝑥:
𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑝 𝜕 2𝑢 𝜕 2𝑢 𝜕 2𝑢
𝜌( +𝑢 +𝑣 +𝑤 )=− + 𝜇 ( 2 + 2 + 2 ) + 𝜌𝑔𝑥 (121)
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑥 𝜕𝑥 𝜕𝑦 𝜕𝑧
Component 𝑦:
𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝑝 𝜕 2𝑣 𝜕 2𝑣 𝜕 2𝑣
𝜌( +𝑢 +𝑣 +𝑤 )=− + 𝜇 ( 2 + 2 + 2 ) + 𝜌𝑔𝑦 (122)
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑦 𝜕𝑥 𝜕𝑦 𝜕𝑧
Component 𝑧:
𝜕𝑤 𝜕𝑤 𝜕𝑤 𝜕𝑤 𝜕𝑝 𝜕 2𝑤 𝜕 2𝑤 𝜕 2𝑤
𝜌( +𝑢 +𝑣 +𝑤 )=− +𝜇( 2 + + ) + 𝜌𝑔𝑧 (123)
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑧 𝜕𝑥 𝜕𝑦 2 𝜕𝑧 2
Eqs. 121, 122 and 123 can be written in a more compact form using
vectorial notation:
𝜕𝑣⃗
𝜌 ( + (𝑣⃗ ∙ ∇)𝑣⃗) = −∇𝑝 + 𝜇∇2 𝑣⃗ + 𝜌𝑔⃗ (124)
𝜕𝑡
Eqs 121 to 123 (or Eq. 124) are called the Navier-Stokes equations. The
steps given in this note to obtain the equations are based in the work of George
:: 33 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
:: 34 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
equations is done numerically, and it has open a complete field of research y fluid
mechanics, the computational fluid dynamics (CFD).
Fig. 18xx.- Tattoo in the arm of one student of the Mining Engineering
Department of the University of Chile.
The steady, uniform 2-D laminar flow with free surface over an inclined
plane. To fix ideas, we will solve now the problem that we already analysed
using the integral approach. A fluid of density 𝜌 and dynamic viscosity 𝜇 flows
over a plane inclined an angle 𝜃 with a flow depth 𝐻 due to the action of gravity.
The flow is steady, uniform and laminar. The problem is to determine the
velocity and pressure distributions.
We choose the coordinate system indicated in Fig. 18. As the flow is 2-D,
we can omit the 𝑧 component of the momentum equation and drop 𝑤 and 𝑧
𝑝𝑎𝑡𝑚 𝑔⃗
𝜃
Fig. 18.- Steady, uniform, 2-D free surface laminar flow over an inclined
plane
derivatives. Thus the equation of continuity and Navier-Stokes are reduced to:
:: 35 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
Continuity equation:
𝜕𝑢 𝜕𝑣
+ =0 (125)
𝜕𝑥 𝜕𝑦
As the flow is uniform, the streamlines are parallel to the 𝑥 axis (the free
surface is parallel to the bottom). Thus, there is not component of the velocity in
the 𝑦 direction. Therefore:
and
𝜕𝑢
=0 (127)
𝜕𝑥
Navier-Stokes equation in the 𝑥 direction:
𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑝 𝜕 2𝑢 𝜕 2𝑢
𝜌( +𝑢 +𝑣 )=− + 𝜇 ( 2 + 2 ) + 𝜌𝑔𝑥 (128)
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑥 𝜕𝑥 𝜕𝑦
Eqs. 128 and 129 are greatly simplified with the conditions of the problem
and the result obtained from the continuity equation (Eqs. 126 and 127)
The condition of steady state means that the partial derivatives with
respect to 𝑡 are zero. This condition and Eqs. 126 and 127 leads to:
Navier-Stokes equation in the 𝑥 direction:
𝜕𝑝 𝜕 2𝑢
0=− + 𝜌𝑔𝑥 + 𝜇 2 (130)
𝜕𝑥 𝜕𝑦
The last equation indicates that the pressure distribution varies linearly
with 𝑦, i.e., we have a hydrostatic pressure distribution. Integrating Eq. 131:
:: 36 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
From Eq.127 we know that the third term of Eq. 133 does not depends on
𝑥. It means that we can take that term as a constant 𝐾 if we integrate Eq. 133
with respect to 𝑥:
𝜕𝐶1
= 𝜌𝑔𝑥 + 𝐾 (134)
𝜕𝑥
Thus:
𝑝 = 𝜌𝑔𝑦 (𝑦 − 𝐻) (139)
:: 37 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑝 = 𝜌𝑔 cos 𝜃 (𝐻 − 𝑦) (141)
Known the pressure, we can go back to Eq. 130. From Eq. 141, we have
𝜕𝑝⁄𝜕𝑥 = 0, thus:
𝜕 2𝑢
𝜇 2 = −𝜌𝑔𝑥 (142)
𝜕𝑦
𝜕 2𝑢 𝑔
= − sin 𝜃 (143)
𝜕𝑦 2
𝑔𝑦2
𝑢 = − sin 𝜃 + 𝐴𝑦 + 𝐵 (144)
2
The second boundary condition is on the free surface. As there is not shear
stress applied on it, the condition is:
That is to say:
𝜏𝑦𝑥 | =𝜇
𝜕𝑢
| =0 (147)
𝑦=𝐻 𝜕𝑦 𝑦=𝐻
:: 38 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
From Eq. 149 we see that the velocity distribution is parabolic, with a
maximum value at the free surface equal to 𝑢 = 𝑔 sin 𝜃 𝐻 2 ⁄(2). We can also
compute the average velocity, 𝑢̅ = 𝑈:
1 𝐻
𝑈 = ∫ 𝑢 𝑑𝑦
𝐻 0
(150)
1𝑔
𝑈= sin 𝜃 𝐻 2
3
We can write Eq. 150 in dimensionless form multiplying both sides by 𝑈⁄(𝑔𝐻):
𝑈 2 1 𝑈𝐻
= sin 𝜃 (151)
𝑔𝐻 3
We recognize in the left hand side of Eq. 151 a Reynolds number based on the
flow depth:
𝑈𝐻
𝑅𝑒𝐻 =
(152)
In the left hand side there is a dimensionless number that appears when we are
work when gravity forces are important. It is the square of Froude number, which is
defined as:
𝑈
𝐹𝑟 =
√𝑔𝐻 (153)
The equation that relates the average velocity of the flow with its depth is called
in hydraulics “resistance law”, which in this case can be written as:
1
𝐹𝑟 2 = 𝑅𝑒𝐻 sin 𝜃
3 (154)
As we know the velocity distribution, we can also compute the coefficients 𝛼 and
𝛽 introduced in the integral approach (Eqs. 38 and 53), resulting 𝛼 = 54⁄35 ≈ 1.543
and 𝛽 = 18⁄15 = 1.2.
We can also compute the shear stress distribution and its value at the bottom,
that we called 𝜏0 :
:: 39 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝜕𝑢
𝜏𝑦𝑥 = 𝜇
𝜕𝑦
(155)
𝜏𝑦𝑥 = 𝜌𝑔 sin 𝜃(𝐻 − 𝑦)
(156)
𝜏0 = 𝜌𝑔𝐻 sin 𝜃
Note that the relationships given by Eqs. 155 and 156 are the same than
those obtained with the application of the momentum theorem (Eqs. 72 and 64).
Something that we cannot obtain from the integral approach is the resistance
relationship given by Eq. 154. The equivalent equation in the integral approach
is the Darcy-Weisbach equation (Eq. 47 that evolved into Eq. 57). However, this
equation requires to know the friction factor 𝑓, that should be determined from
other way (theoretically, numerically or experimentally). Using Eqs. 57 and 154
we can obtain the friction factor for a steady, uniform, 2-D laminar flow in a
channel:
𝑈2 𝐹𝑟 2
sin 𝜃 = 𝑓 =3 (157)
8𝑔𝐻 𝑅𝑒𝐻
From where:
24
𝑓=
𝑅𝑒𝐻 (158)
Do not mistake the expression given by Eq. 158 with the expression for a
cylindrical pipe, 𝑓 = 64⁄𝑅𝑒 (and shown graphically in Moody’s diagram).
Replacing in 𝑅𝑒 the diameter 𝐷 by 4𝑅𝐻 we do not obtain Eq. 158 but 𝑓 = 16⁄𝑅𝑒𝐻 ,
which is wrong. The factor 24 has been confirmed experimentally (Chow, 1988).
:: 40 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
Turbulent flows are unsteady flows, in the sense that at any point of the
flow domain, the flow properties (𝑣⃗, 𝑝) are fluctuating in time. This is shown in
Fig.2.1, where the three components of the velocity measured with an acoustic
Dopper velocimeter (ADV) are presented. It is easily seen the fluctuating
characteristic of the data. It is rather difficult to use the data as it is presented
in the figure. Osborne Reynolds in 1895 proposed that in the turbulent flow
velocity and pressure could be decomposed in two terms, called by him mean-
mean-motion and relative-mean-motion (Reynolds, 1895). The peer reviewers of
the paper were G.G. Stokes and H. Lamb. In the first response that they sent to
the editor (Lord Rayleigh), Stokes acknowledged that he did not understand the
work, and Lamb indicated that much of the paper was obscure (Jackson and
Launder, 2007). Currently, we call those terms mean (or average) and
fluctuation. But, mean or average of what?
2.5
(cm/s)
1.5
v(t)
1
0.5
0
w(t)
-0.5
0 20 40 60 80 100 120 140 160
t (seg)
Fig. 2.1.- Record of the three components of the velocity measured at
one location in a turbulent flow in an open channel
:: 41 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
1 𝑡0 +𝑇
𝑢̅(𝑥⃗) = ∫ 𝑢(𝑥⃗, 𝑡) 𝑑𝑡 (2.1)
𝑇 𝑡0
The two components presented in Eq. 2.2 are shown in Fig. 2.2. It is easy
to show that:
𝑢̅ = 𝑢̅ 𝑢̅′ = 0 (2.3)
𝑢̅
𝑢′
Obviously, the same decomposition is valid for the other components of the
velocity and pressure:
:: 42 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑣 = 𝑣̅ + 𝑣′ 𝑤=𝑤
̅ + 𝑤′ 𝑝 = 𝑝̅ + 𝑝′ (2.4)
In the analysis that follows we will consider that the experimental data
obtained in a turbulent flow behave statistically as an ergodic process strictly
stationary (i.e., not only the averages are the same but also any other statistical
property of the flow).
It is easy to show the following properties for any variable 𝑏 that is a
function of time
̅̅̅̅
𝜕𝑏 𝜕𝑏̅
= (2.4)
𝜕𝑥𝑖 𝜕𝑥𝑖
̅̅̅̅̅̅̅̅̅̅
∫ 𝑏 𝑑𝑥𝑖 = ∫ 𝑏̅ 𝑑𝑥𝑖 (2.5)
:: 43 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
u
REALIZATION 1
𝑢1 (𝑡0 )
𝑡0 t
u REALIZATION 2
𝑢2 (𝑡0 )
𝑡0 t
. ⬚ .
. ⬚ .
. ⬚ .
u
REALIZATION 𝑁
𝑢𝑁 (𝑡0 )
𝑡0 t
ENSEMBLE
u AVERAGE
𝑁
1
〈𝑢(𝑡0 )〉 = ∑ 𝑢𝑖 (𝑡0 )
𝑁
𝑖=1
𝑡0 t
̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
𝜕𝑢̅ 𝜕𝑣̅ 𝜕𝑤 ̅ 𝜕𝑢′ 𝜕𝑣′ 𝜕𝑤′
+ + + + + =0 (2.8)
𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑥 𝜕𝑦 𝜕𝑧
:: 44 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝜕𝑢̅ 𝜕𝑣̅ 𝜕𝑤
̅
+ + =0 (2.10)
𝜕𝑥 𝜕𝑦 𝜕𝑧
Eqs. 2.10 and 2.11 indicate that the averaged velocities and their
fluctuations satisfy continuity.
Before repeating the same analysis with the Navier-Stokes equations, we
will modify slightly the equations. We will work first with the 𝑥 component of
the momentum equation. Let’s multiply the continuity equation by 𝑢:
𝜕𝑢 𝜕𝑣 𝜕𝑤
𝑢 +𝑢 +𝑢 =0 (2.12)
𝜕𝑥 𝜕𝑦 𝜕𝑧
𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑢2
𝑢 +𝑢 = 2𝑢 =
𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑥
𝜕𝑢 𝜕𝑣 𝜕𝑢𝑣
𝑣 +𝑢 = (2.14)
𝜕𝑦 𝜕𝑦 𝜕𝑦
𝜕𝑢 𝜕𝑤 𝜕𝑢𝑤
𝑤 +𝑢 =
𝜕𝑧 𝜕𝑧 𝜕𝑧
:: 45 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
Now, we will replace the Reynolds decomposed variables in Eq. 2.13: and
take the average. The resulting equation is:
̅̅̅̅̅̅̅̅̅̅̅
𝜕(𝑢̅ + 𝑢′) 𝜕(𝑢 ̅̅̅̅̅̅̅̅̅̅̅̅
̅ + 𝑢′)2 𝜕(𝑢 ̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
̅ + 𝑢′)(𝑣̅ + 𝑣′) 𝜕(𝑢 ̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
̅ + 𝑢′)(𝑤 ̅ + 𝑤′)
𝜌( + + + )
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧
̅̅̅̅̅̅̅̅̅̅
𝜕(𝑝̅ + 𝑝′) (2.14)
=− ̅̅̅̅̅̅̅̅̅̅̅
+ 𝜇∇2 (𝑢̅ + 𝑢′) + 𝜌𝑔𝑥
𝜕𝑥
The linear terms in Eq. 2.14 are easily decomposed into their average and
fluctuating parts:
̅̅̅̅̅̅̅̅̅̅̅
𝜕(𝑢 ̅ + 𝑢′) 𝜕𝑢̅̅ 𝜕𝑢′ ̅
= + =0 (2.15)
𝜕𝑡 𝜕𝑡 𝜕𝑡
̅̅̅̅̅̅̅̅̅̅
𝜕(𝑝̅ ̅ 𝜕𝑝̅
+ 𝑝′) 𝜕𝑝̅̅ 𝜕𝑝′
= + = (2.16)
𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑥
∇2 ̅̅̅̅̅̅̅̅̅̅̅ ̅ = ∇2 𝑢̅
(𝑢̅ + 𝑢′) = ∇2 𝑢̅ + ∇2 𝑢′ (2.17)
̅̅̅̅̅̅̅̅̅̅̅̅
𝜕(𝑢̅ + 𝑢′)2 𝜕(𝑢 ̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
̅ 2 + 2𝑢̅𝑢′ + 𝑢′2 ) 𝜕𝑢 ̅̅̅ ̅̅̅̅̅̅
̅ 2 𝜕2𝑢 ̅̅̅̅
̅ 𝑢′ 𝜕𝑢′̅ 2 𝜕𝑢̅2 𝜕𝑢′
̅̅̅̅2
= = + + = + (2.18)
𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑥
̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
𝜕(𝑢 ̅̅̅̅̅
̅ + 𝑢′)(𝑣̅ + 𝑣′) 𝜕𝑢̅𝑣̅ 𝜕𝑢′𝑣′
= + (2.19)
𝜕𝑦 𝜕𝑦 𝜕𝑦
̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
𝜕(𝑢̅ + 𝑢′)(𝑤 ̅ + 𝑤′) 𝜕𝑢̅𝑤 ̅̅̅̅̅̅
̅ 𝜕𝑢′𝑤′
= + (2.20)
𝜕𝑧 𝜕𝑧 𝜕𝑧
:: 46 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
Remember that Eq. 2.21 is not more that Newton’s second law applied to
an incompressible Newtonian fluid and expressed in terms of temporal mean
values. In order to interpret the meaning of Eq. 2.21, we should remember the
equation of Newton’s second law: 𝑑(𝑚𝑣⃗)⁄𝑑𝑡 = 𝐹⃗ . As the mass is preserved, we
can write 𝑚𝑑(𝑣⃗)⁄𝑑𝑡 = 𝐹⃗ . For simplicity, let’s consider the 𝑥 component:
:: 47 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑑𝑢
𝑚 = 𝐹𝑥 (2.26)
𝑑𝑡
Of course, Newton’s second law is not applied to fluids in the form of Eq.
2.26, but as Eq. 51 (integral approach) or Eq. 124 (differential approach), but it
is used in this explanation for the sake of clarity. The problem with Eq. 2.26 (and
the reason why in fluids is expressed in a different way is to define the mass 𝑚
in a flow. Anyway, we can divide Eq. 2.26 by a volume and work with forces per
unit volume, 𝐹𝑉𝑥 ≡ 𝐹𝑥 ⁄𝑉 :
𝑑𝑢
𝜌 = 𝐹𝑉𝑥 (2.27)
𝑑𝑡
We can identify 𝐹𝑉𝑥 in Eq. 124 as the resulting of the force due to pressure,
viscosity, and gravity. We would like to work with a temporal mean acceleration,
and write Eq. 2.27 as:
𝑑𝑢̅
𝜌 = 𝐹̅𝑥 (2.28)
𝑑𝑡
We can see that Eq. 2.30 is very similar to Eq. 2.28. The only problem is
the terms due to the fluctuations in the parenthesis. In order to have only the
variation of momentum in the left hand side of Eq. 2.30, we take the terms due
to the fluctuations towards the right hand side of the equation:
To pass the terms due to the fluctuations from one side of the equal
sign to the other side is much more than an algebraic step. It changes the
interpretation that we can give to the average of products of the fluctuations. As
:: 48 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
they are now in the right side of the equation, we can interpret them as forces
arising from the turbulence. Thus, the terms −𝜌𝑢 ̅̅̅̅
′2 , −𝜌𝑢 ′ 𝑣 ′ and −𝜌𝑢
̅̅̅̅̅̅ ′ 𝑤 ′ are
̅̅̅̅̅̅
called Reynold’s apparent stresses or, simply, Reynold’s stresses , or turbulent
stresses. Let’s work with the terms associated to the viscous stresses 𝜇∇2 𝑢̅ :
𝜕 2 𝑢̅ 𝜕 2 𝑢̅ 𝜕 2 𝑢̅
∇2 𝑢̅ = + + (2.32)
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2
𝜕
∇2 𝑢̅ = ∇2 𝑢̅ + ∇ ∙ 𝑣̅⃗ (2.33)
𝜕𝑥
𝜕 2 𝑢̅ 𝜕 2 𝑢̅ 𝜕 2 𝑢̅ 𝜕 2 𝑢̅ 𝜕 2 𝑣̅ 𝜕 2𝑤
̅
∇2 𝑢̅ = 2
+ 2
+ 2
+ + + (2.34)
𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑥𝜕𝑥 𝜕𝑥𝜕𝑦 𝜕𝑥𝜕𝑧
:: 49 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
̅̅̅̅2 ,
𝜏 𝑇𝑥𝑥 = −𝜌𝑢′ ̅̅̅̅̅ ,
𝜏 𝑇𝑦𝑥 = −𝜌𝑢′𝑣′ ̅̅̅̅̅̅
𝜏 𝑇𝑧𝑥 = −𝜌𝑢′𝑤′ (2.42)
We define the total stress as the sum of the viscous and the turbulent one:
𝑇𝑥𝑥 = 𝜏𝑉𝑥𝑥 + 𝜏 𝑇𝑥𝑥 , 𝑇𝑦𝑥 = 𝜏𝑉𝑦𝑥 + 𝜏 𝑇𝑦𝑥 , 𝑇𝑧𝑥 = 𝜏𝑉𝑧𝑥 + 𝜏 𝑇𝑧𝑥 (2.43)
In general:
𝜕𝑢̅𝑖 𝜕𝑢̅𝑗
𝜏𝑉𝑖𝑗 = 𝜇 ( + ) (2.45)
𝜕𝑥𝑗 𝜕𝑥𝑖
̅̅̅̅̅̅
𝜏 𝑇𝑖𝑗 = −𝜌𝑢 ′ ′
𝑖 𝑢𝑗 (2.46)
:: 50 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝜕𝑤
̅ 𝜕𝑤
̅ 𝜕𝑤
̅ 𝜕𝑝̅ 𝜕𝑇𝑥𝑧 𝜕𝑇𝑦𝑧 𝜕𝑇𝑧𝑧
𝜌 (𝑢̅ + 𝑣̅ +𝑤
̅ )=− + + + + 𝜌𝑔𝑧 (2.49)
𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑧 𝜕𝑥 𝜕𝑦 𝜕𝑧
Eqs. 2.47, 2.48 and 2.49 are the Reynolds equations for turbulent flows,
and they constitute an important advance in the study and analysis of the
turbulence. In his paper, Reynolds also derived and discussed the equation for
the turbulent kinetic energy.
Although we have had some progress in the analysis of turbulent flows,
we are not in conditions to solve any problem yet. We have a system formed by
4 partial differential equations (with their boundary conditions): Eqs. 2.10, 2.47,
̅, 𝑝̅ , ̅̅̅̅
2.48 and 2.49. But we have 10 unknowns: 𝑢̅, 𝑣̅ , 𝑤 𝑢′2 , ̅̅̅̅
𝑣′2 , ̅̅̅̅̅
𝑤′2 , ̅̅̅̅̅
𝑢′𝑣′ , ̅̅̅̅̅̅
𝑢′𝑤′ , ̅̅̅̅̅̅
𝑣′𝑤′. We
cannot solve any problem of turbulence using the equations derived with the
Reynolds decomposition approach if we do not know relations for −𝜌𝑢 𝑖 𝑢𝑗 . This is
̅̅̅̅̅̅
′ ′
:: 51 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
his mistake was to assume that the velocity fluctuations were not correlated, i.e.,
𝑢𝑖′ 𝑢𝑗′ = 0 (using our notation). In this way, he lost additional components to the
̅̅̅̅̅̅
stresses. However, it was known that the equations of Navier-Stokes provided
results in agreement with the experiments for conduits with small flow area, but
it failed for larger conduits. In the latter case, additional effects appear in the
flow, with the final effect that the viscosity seems to be larger (in our language
we would say that the Navier-Stokes were in agreement with measurements in
laminar flows, but other effects should be taking into account when dealing with
turbulent flows. Details are in the first 50 pages of Boussinesq’s book. For open
channel flows he gives:
𝜀 = 𝜌𝑔𝐴ℎ𝑢0 (2.50)
Eq. 2.50 corresponds to Boussinesq’s Eq. 13. In an open channel flow, ℎ is the
flow depth and 𝑢0 is the “velocity at the wall”. 𝐴 is a coefficient that depends on
the wall roughness varies little with ℎ and 𝑢0 . Note that in order that Eq. 2.50
be dimensionally homogeneous, the dimensions of 𝐴 must be T2L-1
(𝑡𝑖𝑚𝑒 𝑙𝑒𝑛𝑔𝑡ℎ).
2⁄
𝜕𝑢̅
𝜏 𝑇𝑥𝑥 = 𝜀2 (2.51)
𝜕𝑥
𝜕𝑣̅
𝜏 𝑇𝑦𝑦 = 𝜀2 (2.52)
𝜕𝑦
:: 52 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝜕𝑤
̅
𝜏 𝑇𝑧𝑧 = 𝜀2 (2.53)
𝜕𝑧
𝜕𝑢̅ 𝜕𝑣̅
𝜏 𝑇𝑥𝑦 = 𝜏 𝑇𝑦𝑥 = 𝜀 ( + ) (2.54)
𝜕𝑦 𝜕𝑥
𝜕𝑢̅ 𝜕𝑤
̅
𝜏 𝑇𝑥𝑧 = 𝜏 𝑇𝑧𝑥 = 𝜀 ( + ) (2.55)
𝜕𝑧 𝜕𝑥
𝜕𝑣̅ 𝜕𝑤
̅
𝜏 𝑇𝑦𝑧 = 𝜏 𝑇𝑧𝑦 = 𝜀 ( + ) (2.60)
𝜕𝑧 𝜕𝑦
𝜕𝑢̅
=0 (2.61)
𝜕𝑥
:: 53 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑇 = 𝐻𝑢∗ (1 − )
𝑦
= 0.4 ; =
𝐻
Axisymetric jet
𝑇 = 0.013𝑉0 𝑑0
𝜕𝑝̅
0=− + 𝜌𝑔𝑦 (2.63)
𝜕𝑦
:: 54 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑝̅ = 𝜌𝑔 cos 𝜃 (𝐻 − 𝑦) (2.64)
Reynolds equation along 𝑥 (Eq. 2.62) can be integrated once with respect
to 𝑦:
As there is not shear acting on the free surface, the boundary condition at
𝑦 = 𝐻 is 𝑇𝑦𝑥 = 0. Thus:
𝐶1 = 𝜌𝑔𝑥 𝐻 (2.66)
Considering that the molecular viscosity 𝜇 is much smaller than the eddy
viscosity 𝜀:
𝜕𝑢̅
𝜀 = 𝜌𝑔𝑥 (𝐻 − 𝑦) (2.69)
𝜕𝑦
To drop 𝜇 from the momentum equation limits its use only to the region
where turbulence dominates over viscosity. We will see later that even in highly
turbulent flows bounded by smooth walls, in a small region near the wall
molecular viscosity effects are important.
In order to integrate Eq. 2.69, we have to decide what value of 𝜀 will be
used. We can work with the expression given by Boussinesq (Eq. 2.50) or that
presented in Table 2.1.
Solution considering Boussinesq expression for 𝜀:
Eq. 2.50 is 𝜀 = 𝜌𝑔𝐴ℎ𝑢0 , and Eq. 2.69 becomes:
𝜕𝑢̅ 𝑔𝑥
= (𝐻 − 𝑦) (2.70)
𝜕𝑦 𝜌𝑔𝐴ℎ𝑢0
:: 55 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
Integrating:
𝑔𝑥 1
𝑢̅ = (𝐻𝑦 − 𝑦 2 ) + 𝐶2 (2.71)
𝑔𝐴ℎ𝑢0 2
Fig. 2.4.- Comparison of the velocity distribution for laminar flow and
turbulent flow according to the eddy viscosity model of Boussinesq
(1877) that includes a velocity 𝒖𝟎 at the wall.
:: 56 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
Integrating:
𝑔𝑥 𝐻
𝑢̅ = ln(𝑦) + 𝐶3 (2.75)
𝑢∗
:: 57 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑢̅(𝑦) 𝑢′(𝑦)
𝑦+𝑙
𝑢̅(𝑦 + 𝑙) B
𝑣′(𝑦)
𝑦
𝑢̅(𝑦) A
𝑢̅
Fig. 2.6.- Concept of mixing length
figure, the continuous curve represents the mean velocity profile and the circles
:: 58 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
a mass of fluid. A mass of fluid that initially was at A with velocicty 𝑢̅(𝑦), due to
the vertical fluctuation 𝑣′ is transported to the location B, without losing its
identity, i.e., preserving its momentum in the 𝑥 direction (per unit volume) 𝜌𝑢̅.
The distance of the displacement of the mass of fluid is 𝑙, named mixing length
by Prandtl. As the mass of fluid did not changed its properties during the
displacement along the 𝑦 axis, when it arrives to the location B, has the velocity
𝑢̅(𝑦), which is imposed at the level (𝑦 + 𝑙). It is easy to see that the difference
between this new velocity at (𝑦 + 𝑙 ) and the mean velocity at this level, 𝑢̅(𝑦 + 𝑙),
corresponds to the velocity fluctuation 𝑢′:
Thus:
𝜕𝑢̅
𝑢′ = −𝑙 (2.79)
𝜕𝑦
Note that, for the mean velocity profile of Fig. 2.6, 𝜕𝑢̅⁄𝜕𝑦 > 0, and 𝑢′ < 0 .
The fluctuating velocity in the 𝑦 direction was positive, 𝑣 ′ > 0. This means that
𝑢′ 𝑣 ′ < 0 (2.80)
From the las two arguments (Eqs. 2.80 and 2.81), we can write:
𝜕𝑢̅
𝑣 ′ ~𝑙 (2.82)
𝜕𝑦
̅̅̅̅̅
Thus, the turbulent shear stress 𝜏 𝑇𝑥𝑦 = −𝜌𝑢 ′ 𝑣′ (Eq. 2.42) can be written
as:
:: 59 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝜕𝑢̅ 2
𝜏 𝑇𝑥𝑦 2
= 𝜌𝑙 ( ) (2.83)
𝜕𝑦
From Eq. 2.83 we obtain that the eddy viscosity for this flow is given by:
𝜕𝑢̅
𝜀 = 𝜌𝑙 2 (2.84)
𝜕𝑦
We can try to solve the problem of the permanent uniform flow over an
inclined plane using Prandtl’s expression for the turbulent shear stress (Eq.
2.83). We had obtained (Eq. 2.67):
𝜕𝑢̅ 𝜕𝑢̅ 2
𝜇 2
+ 𝜌𝑙 ( ) = 𝜌𝑔𝑥 (𝐻 − 𝑦) (2.85)
𝜕𝑦 𝜕𝑦
𝜕𝑢̅ 2
𝑙2 ( ) = 𝑔𝑥 (𝐻 − 𝑦) (2.86)
𝜕𝑦
:: 60 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑑𝑢̅ 4
( )
𝑑𝑦
2 2 = 𝑔𝑥 (𝐻 − 𝑦) (2.88)
𝑑 2 𝑢̅
( 2)
𝑑𝑦
𝑢∗ 𝑦 𝑦
𝑢̅ = 𝑈𝑚𝑎𝑥 + [ln (1 − √1 − ) + √1 − ] (2.90)
𝐻 𝐻
Note that Eq. 2.90 fails at 𝑦 = 0. This is correct because it is valid only in
the region where the turbulent stresses dominate over the viscous ones.
Further, von Karman assumes that 𝑙 should “quietly diminish to zero and
the velocity distribution near to the wall becomes” (Eq. 25 in von Karman’s
paper):
𝜕𝑢̅ 𝜕𝑢̅ 2
𝜏𝑦𝑥 − 𝜇 = 𝜌(𝑦)2 ( ) (2.91)
𝜕𝑦 𝜕𝑦
:: 61 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑙 = 𝑦 (2.92)
OUTER TURBULENT
REGION REGION
FLOW
𝑦 INNER
REGION
BUFFER LAYER
VISCOUS SUBLAYER
The applicability of Eq. 2.92 is near the wall, but in the turbulent region.
Using 𝑙 = 𝑦 in Eq. 2.86, we have:
𝜕𝑢̅ 2
(𝑦) ( ) = 𝑔𝑥 (𝐻 − 𝑦)
2 (2.93)
𝜕𝑦
At this point if the analysis, we should mention that according to
Schlichting (1979, p. 587) Eq. 2.92 was Prandtl’s idea that was presented in a
couple of papers in 1925 and 1926 (As I do not read German, I cannot verify
Schlichting’s statement, but both papers do not contain an explicit equation
similar to Eq. 2.92. It is possible that a linearly decreasing mixing length near
the wall be mentioned in the text. Neither in von Karman's article there is an
:: 62 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
explicit expression for the mixing length, but it is inferred from his Eq. 25, as
indicated before).
Limiting the application of Eq. 2.93 to a region very near the wall, 𝑦 ≪ 𝐻,
and recalling that 𝜏0 = 𝜌𝑔𝐻 sin 𝜃, Eq. 2.93 is simplified to:
𝑑𝑢̅ 𝜏0
= √ (2.94)
𝑑𝑦 𝜌
Eq. 2.95 is named the logarithmic law for the velocity distribution or,
simply, logarithmic law. The constants and 𝐶 needs to be determined from
experimental data. We should not be surprised to see that we got the same result
as that obtained with the eddy viscosity. The value of 𝜀 used to obtain the
logarithmic profile given by Eq. 2.76 was obtained using the solution resulting
from the mixing length (Eq. 2.95).
Although there is a discussion about the universality of von Karman’s
constant, it is usually taken as = 0.4. A dependency on the suspended
sediments concentration exists. Determination of the constant 𝐶 deserves a
separated analysis, because it depends on the characteristics of the wall.
In addition to the mixing length given by Eq. 2.92, other have been
proposed. For example, for the turbulent flow in a pipe of radius 𝑅, Nikuradse
suggested (Schlichting, 1979):
𝑙 𝑦 2 𝑦 4
= 0.14 − 0,08 (1 − ) − 0,06 (1 − ) (2.96)
𝑅 𝑅 𝑅
Buffer layer (region that assembles the viscous dominated layer near the
wall with the turbulent region), van Driest proposed (Schlichting, 1979):
1 𝑦𝑢∗
𝑙 = 𝑦 [1 − exp (− )] , 𝐴 = 26 (2.97)
𝐴
:: 63 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑢∗
𝑢̅ = (ln(𝑦) − ln(𝑦0 )) (3.1)
The distance 𝑦0 is of the order of magnitude of 𝛿𝑉 . Using dimensional
analysis, we can get a dimensionless parameter involving 𝑦0 . The relevant
variables near the bottom is the fluid properties (viscosity 𝜇 and density 𝜌), the
shear stress acting on the wall (𝜏0 ). Thus, we can expect a functional relationship
of the form 𝑦0 = 𝑓(𝜇, 𝜌, 𝜏0 ). The number of variables is 𝑛 = 4 and the number of
fundamental dimensions is 𝑟 = 3 (F,L,T). Applying the Buckingham theorem,
we have only one dimensionless parameter (𝑛 − 𝑟 = 1), given by:
⁄
𝜏01 2 𝜌1⁄2 𝑦0
= (3.2)
𝜇
:: 64 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑢∗ 𝑦0
= (3.3)
Thus we have a functional relationship () = 0, that means that must
be a constant. Calling 𝛽 such a constant:
𝑢∗ 𝑦0
=𝛽 (3.4)
Replacing 𝑦0 from Eq. 3.4 into Eq. 3.1 and dividing by the shear velocity:
𝑢̅ 1 𝑦𝑢∗
= (ln ( ) − ln(𝛽)) (3.5)
𝑢∗
It is customary to define the inner (or wall) variables:
𝑢̅ 𝑦𝑢∗
𝑢̅+ ≡ , 𝑦+ ≡ (3.6)
𝑢∗
Eq. 3.5 is re-written as:
1
𝑢̅+ = (ln(𝑦 + ) − ln(𝛽)) (3.7)
The logarithmic profile written in dimensionless form as Eq. 3.7 will help
us to define the viscous sublayer and buffer region thicknesses. However, before
doing that, it will be useful to know the velocity distribution in the viscous
sublayer.
Velocity distribution in the viscous sublayer
The characteristic of the viscous sublayer is that the viscous stresses are
much larger than the turbulent ones, thus we can write 𝑇𝑦𝑥 = 𝜏𝑉𝑦𝑥 . In addition,
as this is a thin layer, we can assume that the shear stress is not so different
than its value on the bottom, 𝜏0 . Thus, we can write:
𝑑𝑢̅
𝜇 = 𝜏0 (3.8)
𝑑𝑦
𝜇 𝑢̅ = 𝜏0 𝑦 (3.9)
:: 65 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝜇 𝜏0
𝑢̅ = 𝑦 (3.10)
𝜌 𝜌
𝑢̅+ = 𝑦 + (3.11)
Thus, in the viscous sublayer the velocity varies linearly with the distance
to the wall.
The viscous sublayer thickness
At this point, we have done a lot to determine the viscous sublayer
thickness, but we need to rely on experimental data. The utility of having the
velocity distribution in terms of dimensionless variables is that it is possible to
compare data obtained under different flow conditions. Some experimental data
is presented in Fig. 3.2, which was taken from White (1991, Fig. 6.11). In
addition to the experimental data, the linear velocity profile form the viscous
sublayer (Eq. 3.11), and the logarithmic profile form the turbulent region
(Eq.3.5) are also plotted.
Obviously, the thickness of the viscous sublayer is defined by the distance
from the wall where the experimental data departs from the linear profile (that
appears as a curve in a semilogarithmic plot). This happens at 𝑦 + ≈ 5. Thus, the
dimensionless viscous sublayer thickness, 𝛿𝑉+ = 𝛿𝑉 𝑢∗ ⁄ is taken as:
𝛿𝑉+ = 5 (3.12)
The experimental data join the line defined by the logarithmic distribution
around 𝑦 + ≈ 70. That means that for 𝑦 + > 70 turbulence dominates over viscous
effects. Finally the buffer layer, or region where the viscous effects are as
important as the turbulent ones is comprised in 5 < 𝑦 + < 70. Often, the buffer
region is neglected and 𝛿𝑉+ ~ 11 is taken. Below that height the motion is
considered laminar and above, fully turbulent.
For the flow in a smooth conduit, the three regions mentioned before are
present. We are still talking about smooth walls but we have not given a precise
definition regarding when a wall can be considered smooth. Following Prandtl
(Schlichting, 1979) and von Karman (1930), a wall is hydraulically smooth when
the roughness is completely contained within the viscous sublayer. A wall is
hydraulically rough when the roughness size is large enough to preclude the
:: 66 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
̅+
𝒖
LINEAR DISTRIBUTION
LOGARITHMIC PROFILE
𝒚+
Fig. 3.2.- Experimental data and velocity distributions for the viscous
sublayer (Eq. 3.11) and turbulent region (Eq. 3.7).
:: 67 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝛿𝑉
a) SMOOTH WALL
𝛿𝑉 𝛿𝑉
𝛿𝑉
𝜀
b) ROUGH WALL
Fig. 3.3.- Definition of hydraulically smooth and rough surfaces.
a) On a smooth wall, the viscous sublayer cover completely the
roughness of the wall. The flow in the turbulent region is not
influenced by the size of the roughness and the flow resistance is
due to the viscous shear acting in the viscous sublayer.
b) The viscous sublayer is completely destroyed by the roughness
elements and the flow resistance is originated by the drag exerted
by them.
𝑘𝑆
Fig. 3.4.- Nikuradse’s roughness. Sand grains were glued to the wall of the
pipes.
Based on the limits of the different regions indicated before, the
hydrodynamic type of wall is defined depending on which of them the roughness
is contained. It is expressed in terms of the roughness size made dimensionless
with the inner variables:
𝑘𝑆 𝑢∗
𝑘𝑆+ = (3.13)
Due to the arbitrariness in the definition of the region boundaries, the
values assigned to 𝑘𝑆+ to define the type of wall vary slightly among different
authors, as shown in Table 3.1,although those assigned by Prandtl are the most
used in practice.
:: 68 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
PRANDTL WHITE
HYDRAULIC TYPE OF WALL
(Schlichting, 1979) (White, 1991)
𝑢̅ 1 𝑦
= ln ( ) (3.14)
𝑢∗ 𝑦0
In this case, we already found that 𝑦0 ~ ⁄𝑢∗ (Eq. 3.4) and Eq. 3.7 is valid,
which usually is written as:
1
𝑢̅+ = ln(𝑦 + ) + 𝐵 (3.15)
and 𝐵 are determined from experimental data, such as those presented
in Fig. 3.2, resulting = 0.4 and 𝐵 = 5.5 (Nikuradse, 1933).
The flow resistance mechanism is due to the drag generated by the grains
large enough to preclude the existence of a viscous sublayer. In this case, it is
relevant the size of the roughness, so 𝑦0 ~ 𝑘𝑆 and Eq. 3.14 becomes:
𝑢̅ 1 𝑦
= ln ( ) + 𝐶 (3.16)
𝑢∗ 𝑘𝑆
:: 69 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
I this case, the flow resistance is the result of both viscous and drag effects.
Thus, a functional relationship 𝑦0 = 𝑓(𝜇, 𝜌, 𝜏0 , 𝑘𝑆 ) should exist. From the
Buckingam theorem we obtain:
𝑦0 𝑢∗ 𝑘𝑆
1 = , 2 = (3.17)
𝑘𝑆
The dimensionless relation between both dimensionless parameters is
written as:
𝑦0 +
( ,𝑘 ) = 0 (3.18)
𝑘𝑆 𝑆
:: 70 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑦0
= 1 (𝑘𝑆+ ) (3.19)
𝑘𝑆
𝑦0 = 𝑘𝑆 1 (𝑘𝑆+ ) (3.20)
𝑢̅ 1 𝑦
= ln ( ) + 𝐴(𝑘𝑆+ ) (3.21)
𝑢∗ 𝑘𝑆
𝐴(𝑘𝑆+ )
:: 71 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
expressions for the velocity profile in the buffer layer. The difference between
them is the presence of a term raised to the fourth power. However, the equation
without this term adjust very well to the experimental data, as shown by White
(1991, Fig. 6.11). The equation is:
−𝐵
(𝑢+ )2 (𝑢+ )3
[𝑒 𝑢
+
+
𝑦 =𝑢 +𝑒 +
− 1 − 𝑢 −
+
− ] (3.22)
2 6
Note that Eq. 3.22 is not explicit in 𝑢+ , with = 0.4 and 𝐵 = 5.5. The range
of validity of this equation extends further than the buffer layer, spanning 0 ≤
𝑦 + < 300.
The different velocity profiles arising in the turbulent flow are
summarized in Table 3.2.
HYDRAULICALLY ROUGH 1 𝑦
𝑢+ = ln ( ) + 8.5
𝑘𝑆+ > 70 𝑘𝑆
:: 72 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑢∗ 𝑓 (68)
=√
𝑈 8
The velocity distributions have the form 𝑢̅⁄𝑢∗ as a function of the distance
from the wall. It looks evident that, integrating the velocity distribution across
the flow section will allows to obtain 𝑈⁄𝑢∗ , hence the friction factor 𝑓. The
analysis that follows will consider cylindrical pipes because in this case exists an
extensive set of data that permits to confirm the validity of the expressions
obtained. The results can be generalized to other geometries.
1
𝑢̅+ = ln(𝑦 + ) + 𝐵 (3.15)
The cross-sectional mean velocity is computed from
1
𝑈= ∫ 𝑢̅ 𝑑𝐴 (3.23)
𝐴 𝐴
For a pipe with inner diameter 𝐷, 𝐴 = 𝜋𝐷2 ⁄4. The distance from the wall
(𝑦) is related to the radial distance from the axis of the pipe (𝑟) trough 𝑦 + 𝑟 =
𝐷⁄2. The element of surface is 𝑑𝐴 = 2𝜋𝑟𝑑𝑟. In order to simplify the computations,
we will neglect the viscous sublayer and buffer regions, i.e., we will assume that
Eq. 3.15 is valid in all the flow domain. Thus, from Eq.s 3.15 and 3.23:
𝐷
4 2 𝑢∗ 𝐷 𝑢∗ (3.24)
𝑈= ∫ ( ln (( − 𝑟) ) + 𝑢∗ 𝐵) 2𝜋𝑟𝑑𝑟
𝜋𝐷 𝑜
2 2
𝐷
8 𝑢∗ 2 𝐷 𝑢∗ (3.25)
𝑈 = 2 ∫ (ln ( − 𝑟) + ln ( ) + 𝐵) 𝑟𝑑𝑟
𝐷 𝑜 2
:: 73 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝐷 𝐷
8 𝑢∗ 2 𝐷 𝑢∗ 2
𝑈 = 2 [∫ ln ( − 𝑟) 𝑟𝑑𝑟 + (ln ( ) + 𝐵) ∫ 𝑟𝑑𝑟] (3.26)
𝐷 𝑜 2 𝑜
𝐷
8 𝑢∗ 2 𝐷 𝑢∗ 1 𝐷 2
𝑈 = 2 [∫ ln ( − 𝑟) 𝑟𝑑𝑟 + (ln ( ) + 𝐵) ( ) ] (3.27)
𝐷 𝑜 2 2 2
𝑢∗ 1 𝐷𝑢∗ 3
𝑈= [(ln ( ) − ) + 𝐵] (3.28)
2 2
𝑢∗ 𝑓 1
=√ = (3.29)
𝑈 8 1 (ln (1 𝐷𝑢∗ ) − 3) + 𝐵
2 2
Generally, the flow mean velocity 𝑈 = 𝑄/𝐴 is known, and not the shear
velocity, and it is convenient to change the argument of the logarithm in Eq. 3.29
multiplying and diving it by 𝑈:
𝐷𝑢∗ 𝐷𝑈 𝑢∗ 𝑓 (3.30)
= = 𝑅𝑒√
𝑈 8
Replacing Eq. 3.30 in Eq. 3.29 and working with its reciprocal:
1 1 3
= [ln(𝑅𝑒√𝑓) − ln(√32) − + 𝐵] (3.31)
√𝑓 √8 2
:: 74 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
1
= 2.035log(𝑅𝑒√𝑓) − 0.913 (3.33)
√𝑓
Eq. 3.33 was obtained by Prandtl in 1935 (White, 1991). As the buffer
region and viscous sublayer were neglected in its deduction, the numerical
constants that appear in it were re-computed in order to fit the experimental
data taken by Nikuradse (1933), resulting:
1
= 2log(𝑅𝑒√𝑓) − 0.8 (3.34)
√𝑓
This is Prandtl’s universal law of friction for smooth pipes (Schlichting,
1979) and it is not have limitations regarding the Reynolds number (within the
turbulent regime). Computation of 𝑓 from Eq. 3.34 must be done graphically or
numerically, usually by iteration. A simpler equation is the empirical
relationship presented by Blasius in 1913, but limited to 3 × 103 ≤ 𝑅𝑒 ≤ 105
(Blasius, 1913, p. 12)
0.3164
𝑓= (3.35)
𝑅𝑒 1⁄4
Friction factor for turbulent flow with hydraulically rough walls
The friction factor for turbulent flows with rough walls is obtained from
the velocity distribution given by Eq. 3.16.
𝑢̅ 1 𝑦
= ln ( ) + 𝐶 (3.16)
𝑢∗ 𝑘𝑆
Assuming eq. 3.16 valid in all the cross-section, computing the average
velocity 𝑈 , rearranging the final result to form √𝑓 = √8 𝑢∗ ⁄𝑈, and using = 0.4
and 𝐶 = 8.5, the following expression is obtained:
1 𝐷
= 0.942 ln (
2𝑘𝑆
) + 1.68 (3.36)
√𝑓
In terms of the logarithm in base 10:
1 𝐷
= 2.2 log ( ) + 1.68 (3.36)
√𝑓 2𝑘𝑆
:: 75 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑈𝐷 𝑢∗ 1
2log (
√8
𝑈
)− = 0.8 (3.40)
√𝑓𝑆
𝑢∗ 𝐷 1
2log ( √8) − = 0.8 (3.41)
√𝑓𝑆
𝑢∗ 𝑘𝑆 𝐷 1
2log ( √8) − = 0.8 (3.42)
𝑘𝑆 √𝑓𝑆
𝐷 1
2log ( ) + 2log(𝑘𝑆+ ) + 2log(√8) − = 0.8 (3.43)
𝑘𝑆 √𝑓𝑆
:: 76 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝐷 1
2log (3.7 ) − 2log(3.7) + 2log(𝑘𝑆+ ) + 2log(√8) − = 0.8 (3.44)
𝑘𝑆 √𝑓𝑆
1 1
− 2log(3.7) + 2log(𝑘𝑆+ ) + 2log(√8) − = 0.8 (3.45)
√𝑓𝑅 √𝑓𝑆
1 1
− = 0.8 + 2log(3.7) − 2log(√8) − 2log(𝑘𝑆+ ) (3.46)
√𝑓𝑅 √𝑓𝑆
1 1
− = 1.033 − 2log(𝑘𝑆+ ) (3.47)
√𝑓𝑅 √𝑓𝑆
Eq. 3.47 is the deviation (in terms of 1⁄√𝑓 ) of the friction factor for smooth
walls from the friction factor for rough walls. Now, we will defining the deviation
of the friction factor for any type of wall from that associated to the rough surface
as:
𝐷 1
𝑀 = 2log (3.7
𝑘𝑆
)− (3.48)
√𝑓
As Eq. 3.48 is valid for any type of wall, 𝑀 should be function of 𝑘𝑆+ . Using
the experimental data presented by Nikuradse (1933) is possible to know 𝑀(𝑘𝑆+ ).
In effect, with the data is possible to generate the Table. 3.3. The first four
columns contains Nikuradse’s data and the fifth is computed using Eq. 3.48.
:: 77 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
The relation between 𝑀 and 𝑘𝑆+ is presented in Fig. 3.7, taken from the
article by Colebrook and White (1937) and corresponds to line labelled
“Nikuradse sanded”. The straight line labelled “smooth law” is Eq. 3.47.
Obviously, for rough walls 𝑀 = 0 (“rough law” in the figure).
𝑘𝑆+
Fig. 3.7.- Deviation of 𝟏⁄√𝒇 from the rough wall. The function 𝑴
covers all type of walls.
The objective of the article by Colebrook and White (1937) was “to
determine how the nature of the roughness influenced the transition”. To
accomplish that goal, sand of different sizes and following some particular
arrangements were glued to the pipe. But they also added information of
commercial pipes reported by other authors (Freeman, quoted by Mills, 1923;
and Heywood, 1924). The commercial pipes generate the curve labelled as
“Galvanised and new wrought iron” in Fig. 3.7. The different behaviour among
the measurements carried out in commercial pipes and those obtained by
Nikuradse is result of the non-uniformity of the roughness size in the commercial
pipes. In this case, protuberances larger than the average disturb and destroy
the viscous sublayer before that the rest of the protuberances. This is a gradual
process that makes an earlier and gradual separation of the 𝑀 curve from the
smooth case when compared with pipes of uniform roughness.
As it was indicated before, it is customary to denote by 𝜀 the equivalent
roughness that arise in pipes with non-uniform size of protuberances. A
dimensionless equivalent roughness is defines as:
𝜀𝑢∗
𝜀+ ≡ (3.49)
:: 78 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
3.29
𝑀 = 2 log (1 + ) (3.50)
𝜀+
It is easy to verify that Eq. 3.50 does not have limitations regarding the
type of wall, covering all of them. For smooth walls, 𝜀 + is very small, and
3.29⁄𝜀 + ≫ 1, and 𝑀 ≈ 2 log(3.29⁄𝜀 + ) = 2 log(3.29) − 2 log(𝜀 + ), that is to say:
Eq. 3.51 is in agreement with Eq. 3.47, the deviation for smooth surfaces.
In the same way, for hydraulically rough walls, 𝜀 + is large, resulting 3.29⁄𝜀 + ≪
1, and 𝑀 ≈ 2log(1) = 0, i.e., there is not deviation from the friction factor for
rough surface.
Thus, it is possible to determine an expression for the friction factor for
turbulent flows, valid for all type of walls. In effect, replacing Eq. 3.50 into eq.
3.48:
3.29 𝐷 1
2 log (1 +
𝜀+
) = 2 log (3.7
𝜀
) − (3.52)
√𝑓
But
1 3.29√8 𝐷 𝐷
− = 2 log (1 + ) − 2 log (3.7 ) (3.54)
√𝑓 𝑅𝑒√𝑓 𝜀 𝜀
1 3.29√8 𝐷 𝜀
− = 2 log [(1 + ) ] (3.55)
√𝑓 𝑅𝑒√𝑓 𝜀 3.7𝐷
1 𝜀 3.29√8 1
− = 2 log ( + ) (3.56)
√𝑓 3.7𝐷 3.7 𝑅𝑒√𝑓
Finally, the friction factor for turbulent flows is pipes is given by:
:: 79 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
1 𝜀 2.51
= −2 log ( + ) (3.57)
√𝑓 3.7𝐷 𝑅𝑒√𝑓
:: 80 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
Barr:
1 𝜀 5.15
= −2 log [ + 0.892 ] (3.59)
√𝑓 3.7𝐷 𝑅𝑒
Eck:
1 𝜀 15
= −2 log [ + ] (3.60)
√𝑓 3.715𝐷 𝑅𝑒
:: 81 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
As the above equations (an others similar) are approximations to Eq. 3.57,
the friction factor computed with them will have an error.
Finally in Table 3.4, typical values of the roughness size for some
materials are presented.
Table 3.4.- TYPICAL VALUES OF THE ROUGHNESS SIZE
MATERIAL 𝜀 (mm)
:: 82 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
:: 83 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
1 𝑘𝑆 𝐾3
= −𝐾1 log ( + ) (3.61)
√𝑓 𝐾2 𝑅 4𝑅𝑒𝑅𝐻 √𝑓
In Eq. 3.61, 𝑅𝑒𝑅𝐻 is the Reynolds number based on the hydraulic radius, i.e.,
𝑅𝑒𝑅𝐻 = 𝑈𝑅𝐻 ⁄ . 𝐾1 , 𝐾2 and 𝐾3 are coefficients fitted by different authors. Yen
(2002) presents the summary shown in Table 3.5.
CHANNEL
REFERENCE 𝐾1 𝐾2 𝐾3 REMARKS
GEOMETRY
Full circular pipe Colebrook (1939) 2.0 14.83 2.52
:: 84 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
When Eq. 3.61 is applied to natural channels (with fixed boundaries), it is also
necessary to define the roughness, 𝑘𝑆 . The bed of natural channels, when it is
formed by granular non-cohesive sediments, usually contains a wide range of
particle sizes. The matter of what is the appropriate size to represent 𝑘𝑆 depends
on the researcher, and usually is expressed in terms of the diameter 𝑑𝑋 of the
size distribution (𝑑𝑋 is the size of the particle under which it is found the 𝑋% of
the sediment):
:: 85 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
𝑘𝑆 = 𝛼𝑆 𝑑𝑋 (3.62)
Yen (2002) gives a list of the diameter 𝑑𝑋 used by different authors and
the corresponding value of the coefficient 𝛼𝑆 , which is reproduced in Table 3.6.
:: 86 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
REFERENCES
BELUCO, A. and E.B. CAMANO SCHETTINI (2016) “An Improved Expression
for a Classical Type of Explicit Approximation of the Colebrook White Equation
with Only One Internal Iteration”, International Journal of Hydraulic
Engineering, Vol. 5, No. 1, pp. 19-23.
BLASIUS, H. (1913) “Das Aehnlichkeitsgesetz bei Reibungsvorgangen in
Flussigkeiten”, Mitteilungen Über Forschungsarbeiten auf dem Gebiete des
Ingenieurwesens, Heft 131, pp. 1-40.
BOUSSINESQ, J. (1877) Essai sur la théorie des eaux courantes, Mémoires
présentés par divers savants à l’Académie des Sciences. Imprimerie Nationale,
Paris.
BRADSHAW, P. (1972) “The understanding and prediction of turbulent flow”,
Aeronautical Journal, Vol. 76, No. 739, pp. 403-418.
COLEBROOK, C.F. (1939) “Turbulent Flow in Pipes, with particular reference
to the Transition Region between the Smooth and Rough Pipe Laws”, J. Inst.
Civil Eng. (London), Vol. 11, No. 4, pp. 133–156.
COLEBROOK, C.F. and C.M. WHITE (1937) “Experiments with Fluid Friction
in Roughened Pipes”, Proc. R. Soc. Lond. A, Vol. 161, pp. 367-381.
CHOW, V.T. (1988) Open-channel Hydraulics, McGraw-Hill.
HINZE, J.O. (1975) Turbulence, 2nd edition, McGraw‐Hill, New York.
JACKSON, D., and B. LAUNDER (2007) “Osborne Reynolds and the Publication
of His Papers on Turbulent Flow” Annu. Rev. Fluid Mech. Vol. 39, pp.19–35.
MOODY, L.F. (1944). "Friction factors for pipe flow." Trans. ASME, Vol. 66, pp.
671-678
MOSKVITCH, K. (2014) “Fiendish million-dollar proof eludes mathematicians.
Proposed solution to Navier–Stokes equations, used to model fluids, is wrong”,
Nature News (05 August 2014) doi:10.1038/nature.2014.15659
NAVIER, C.L.M.H. (1823) "Memoire sur les lois du mouvement des fluides,"
Mem. Acad. R. Sci. Paris, Vol. 6, pp. 389-416.
NIKURADSE, J. (1933) "Stromungsgesetze in rauhen Rohren." VDI-
Forschungsheft 361. Beilage zu "Forschung auf dem Gebiete des
Ingenieurwesens" Ausgabe B Band 4, July/August 1933. English translation:
“Laws of flow in rough pipes”, NACA TM 1292, Washington, November 1950.
:: 87 ::
Università degli Studi della Campania Luigi Vanvitelli
Dipartimento di Ingegneria Civile Design Edilizia e Ambientale
:: 88 ::