Fluid Mech
Fluid Mech
Olivier Cleynen
A one-semester course for students in
the Chemical and Energy Engineering program
at the University Otto von Guericke of Magdeburg
2015-2020
3
5.2 Shear forces on walls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
5.3 Shear fields in fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
5.4 Resistance to shear: viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
5.5 Special case: shear in simple laminar flows . . . . . . . . . . . . . . . . . . . 101
5.6 Solved problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
5.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
6 Prediction of fluid flows 111
6.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
6.2 Organizing calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
6.3 Equations for all flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
6.4 Equations for incompressible flow . . . . . . . . . . . . . . . . . . . . . . . . 123
6.5 CFD: the Navier-Stokes equations in practice . . . . . . . . . . . . . . . . . . 127
6.6 Solved problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
6.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
7 Pipe flows 135
7.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
7.2 Frictionless flow in pipes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
7.3 Parameters to quantify losses in pipes . . . . . . . . . . . . . . . . . . . . . . 137
7.4 Laminar flow in pipes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
7.5 Turbulent flow in pipes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
7.6 Engineer’s guide to pipe flows . . . . . . . . . . . . . . . . . . . . . . . . . . 146
7.7 Solved problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
7.8 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
8 Engineering models 161
8.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
8.2 Comparing influences: the weighted momentum balance . . . . . . . . . . . 161
8.3 Making models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
8.4 Comparing results: coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . 169
8.5 Solved problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
8.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
9 Dealing with turbulence 179
9.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
9.2 Recognizing turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
9.3 The effects of turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
9.4 Quantifying turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
9.5 Computing turbulent flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
9.6 Commented bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
9.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
10 Flow near walls 199
10.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
10.2 The concept of boundary layer . . . . . . . . . . . . . . . . . . . . . . . . . . 199
10.3 Laminar boundary layers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
10.4 Boundary layer transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
10.5 Turbulent boundary layers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
10.6 Flow separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
10.7 Solved problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
10.8 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
11 Large- and small-scale flows 221
11.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
11.2 Flow at large scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
11.3 Plotting velocity with functions . . . . . . . . . . . . . . . . . . . . . . . . . 223
11.4 Flow at very small scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
11.5 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
Appendix 245
A1 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
A2 Vector operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
A3 Field operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
A4 Derivations of the Bernoulli equation . . . . . . . . . . . . . . . . . . . . . . 253
A5 Flow parameters as force ratios . . . . . . . . . . . . . . . . . . . . . . . . . . 256
A6 Details of the winter 2020-2021 final examination (updated February 2021) . 259
A7 Example of previous examinations . . . . . . . . . . . . . . . . . . . . . . . . 261
A8 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292
About this course (syllabus)
Fluid dynamics for engineers by Olivier Cleynen, PhD
University of Magdeburg, 2015-2020
https://fluidmech.ninja/
Welcome! From 2015 to 2020, these notes were the core of the Fluid Dynamics course of
the Chemical and Energy Engineering program at the University of Magdeburg, Germany.
Objectives
Starting with little or no experience with fluid mechanics, after taking this course:
• you should have a good understanding of what can, and cannot, be calculated with
fluid mechanics in engineering: how we approach problems depending on how
much information is available.
• you should be able to solve several real-world engineering fluid mechanics prob-
lems with confidence: calculating forces within fluids and on objects, predicting
flow in pipes, near walls, at small and large scales.
My objective is to enable you to get there with the minimum amount of your time and
energy (but not minimum power!).
If all goes well, at the end of the semester, you should be well-prepared to begin a course
in Computational Fluid Dynamics, where the knowledge and skills you acquire here can
be used to solve applied problems in great detail.
Contact
My email is olivier cleynen.fr. I am available to answer queries about re-using the
course materials. Unfortunately, I cannot provide assistance with the content itself (e.g.
how to solve the end-of-chapter problems) or answer any questions about the course.
If you would like such assistance, please contact my colleagues at the fluid dynamics
laboratory of the University of Magdeburg.
I hope you have a great semester! Fluid mechanics is one of the most exciting disciplines
out there. Now, let us begin!
Olivier Cleynen
April 2020 (updated October 2022)
8
Fluid Dynamics
Chapter 1 – Basic flow quantities
last edited April 18, 2021
by Olivier Cleynen — https://fluidmech.ninja/
These notes are based on textbooks by White [22], Çengel & al.[25], Munson & al.[29], and de Nevers [17].
There are about 2 ⋅ 1022 molecules in the air within an “empty” 1-liter bottle
continuously differentiated.
at ambient temperature and pressure. Even when the air within the bottle is
completely still, these molecules are constantly colliding with each other and
Figure 1.1: Measurement of the average value of a property (here, velocity 𝑉 ; but it
could be pressure, or temperature) inside a given volume. As the volume shrinks
towards zero, the fluid can no longer be treated as a continuum; and property
measurements will oscillate wildly.
Figure CC-by-sa Olivier Cleynen
10
A direct benefit of the continuum abstraction is that the mathematical com-
plexity of our problems is greatly simplified. Finding the solution for the
equation (eq. 4/15 p. 80) instead of a system of 2 ⋅ 1022 equations with 2 ⋅ 1022
bottle of “still air” mentioned above, for example, requires only a single
Analytical fluid mechanics which is the main focus of these lectures and
which consists in predicting fluid flows mathematically. As we shall
SMBC #2010-08-29: cooperation
see, it is only able to provide (exact) solutions for very simple flows. between theoretical and experi-
In fluid mechanics, theory nevertheless allows us to understand the mental scientists
by Zach Weinersmith
mechanisms of complex fluid phenomena, describe scale effects, and https://www.smbc-
comics.com/comic/2010-08-29
predict forces associated with given fluid flows;
11
1.3 Important concepts in mechanics
Mechanics in general deals with the study of forces and motion of bodies. A
few concepts relevant for us are recalled here.
The velocity 𝑉⃗ of the object is the rate of change in time of its position:
d𝑥⃗
𝑉⃗ ≡
d𝑡
(1/1)
𝑉 ≡ ||𝑉⃗ || (1/2)
⎛ 𝑢𝑟 ⎞
𝑉 = ⎜ 𝑢𝜃
⃗ ⎟
⎜ ⎟
⎝ 𝑢𝑧 ⎠
The acceleration 𝑎⃗ of the object is the rate of change in time of its velocity:
d𝑉⃗
𝑎⃗ ≡
d𝑡
(1/3)
12
Figure 1.2: A body (black dot) is following a trajectory, plotted in blue. The accel-
eration vectors are plotted in red for two cases: on top, when the magnitude of
its velocity remains constant, and on the bottom, when this magnitude changes
continuously.
Figure CC-0 Olivier Cleynen
The three main types of forces relevant to fluid mechanics are those due
to pressure, those due to shear, and those due to gravity (the force due to
torsion (or “twisting”) effort named moment.w Like a force, a moment has a
vector 𝑀⃗ is expressed as the cross product of the arm 𝑟⃗ and the force 𝐹⃗:
direction as well as a magnitude, and so is best expressed with a vector. This
𝑀⃗ ≡ 𝑟⃗ ∧ 𝐹⃗ (1/4)
See Appendix A2.2 p. 248 for a short briefing about the cross product of
vectors.
1.3.3 Energy
Energy, measured in joules (J), is in most general terms the ability of a body
to set other bodies in motion. It can be accumulated or spent by bodies in a
large number of different ways. The most relevant forms of energy in fluid
mechanics are:
13
Workw noted 𝑊 , which is energy spent on displacing an object over a
distance 𝑙 with a force 𝐹 :
𝑊 ≡ 𝐹⃗ ⋅ 𝑙⃗ (1/6)
where 𝑊 is the work (J);
𝐹⃗ is the force (vector with magnitude in N);
and 𝑙⃗ is the movement distance (vector with magnitude in m).
See Appendix A2.1 p. 247 for a short briefing about the dot product of
vectors.
Internal energyw noted 𝐼 stored as heat within the body itself. As long
as no phase changes occurs, the internal energy 𝐼 of fluids is roughly
proportional to their absolute temperature 𝑇 .
1.4.1 Density
The density 𝜌 (Greek letter rho) is the amount of mass per unit volume:
𝑚
𝜌≡
(1/7)
conditions, air has a density of approximately 𝜌air = 1,2 kg m−3 ; that of water
Two orders of magnitude that are useful to remember: at ambient atmospheric
1.4.2 Phase
Fluids can be broadly classified into phases,w which are loosely-defined sets
of physical behaviors. Typically one distinguishes liquids which are fluids
with large densities on which surface tension effects play an important role,
from gases or vapors which have low densities and no surface tension effects.
Phase changes are often brutal (but under specific conditions can be blurred
14 or smeared-out); they usually involve large energy transfers. The presence of
multiple phases in a flow is an added layer of complexity in the description
of fluid phenomena.
1.4.3 Temperature
Temperaturew is a scalar property measured in Kelvins (an absolute scale). It
represents a body’s potential for receiving or providing heat and is defined,
Although we can “feel” temperature in daily life, it must be noted that the
human body is a very poor thermometer in practice. This is because we
constantly produce heat, and we infer temperature by the power our body
loses or gains as heat. This power not only depends on our own body
temperature (hot water “feels” hotter when we are cold), but also on the
heat capacity of the fluid (cold water “feels” colder than air at the same
temperature) and on the amount of convection taking place (ambient air
“feels” colder on a windy day).
is extremely high (𝑐air ≈ 1 kJ kg−1 K, 𝑐water ≈ 4 kJ kg−1 K). Unless very high
In spite of these impressions, the fact is that the heat capacity of fluids
velocities are attained, the temperature changes associated with fluid flow
are much too small to be measurable in practice.
where 𝑅specific depends on the state and nature of the gas (J K−1 kg−1 );
and 𝑝 is the pressure (Pa, see §1.5.2 further down).
Note that 𝑅specific here is a specific gas constant (whose value depends on the
gas); chemists often instead use a universal definition of 𝑅 in J mol−1 K−1 .
molecules within the fluid, and it is called the speed of sound,w noted 𝑐.
by the movement of an object). This speed is equal to the average speed of
In fluid dynamics, we often quantify how fast the fluid is flowing relative to
𝑉
[Ma] ≡
𝑐
(1/10)
Since both 𝑉 and 𝑐 can be functions of space in a given flow, [Ma] may not be
uniform (e.g. the Mach number around an aircraft in flight is different at the
nose and above its wings). Nevertheless, a single value is typically chosen to
identify “the” representative Mach number of any given flow.
at [Ma] ≤ 0,3, their density 𝜌 stays constant. Density variations in practice can
It is observed that providing no heat or work transfer occurs, when fluids flow
be safely neglected below [Ma] = 0,6. When the density is uniform, the flow
is said to be incompressible. Above these Mach numbers, it is observed that
when subjected to pressure variations, fluids exert work upon themselves,
which translates into measurable density and temperature changes: these are
called compressibility effects, and we will not study them in this course.
In most flows, the density of liquids is almost invariant – so that water flows
are generally entirely incompressible.
√
𝑐 = 𝛾 𝑅𝑇 (1/11)
1.4.6 Viscosity
We have said above that a fluid element can deform continuously under
pressure and shear efforts: it will never “snap” or break apart. However this
deformation is not “for free”: it will require force and energy inputs which
are not reversible (they are not reversed if the motion is reversed). Resistance
to straining in a fluid is measured with a property named viscosity.w Video: how to make sense of vis-
cosity
In informal terms, viscosity is the “stickiness” of fluids: for example, honey by Olivier Cleynen (CC-by)
https://youtu.be/5YUFt-V_kdk
and sugar syrups are more viscous than water.
More formally, viscosity is quantified as follows. Imagine a small brick-
in fig. 1.3. The continuous straining of the brick requires a force 𝐹 per unit
shaped element of fluid, which is deformed (strained) horizontally, as shown
area 𝐴, called shear stress 𝜏 (see also §1.5.3 further down). We expect that
and with the speed Δ𝑣 at which the brick is strained. Conversely, we expect
the shear increases with both the “stickiness” of the fluid —the viscosity—
that the shear will decrease when the element height Δ𝑦 is increased.
We define the viscosity 𝜇 as the ratio between the required shear stress,
𝜏 = 𝐹 /𝐴, and the rate at which the brick is strained, Δ𝑣/Δ𝑦:
𝜏
𝜇≡
( Δ𝑦 )
Δ𝑣
(1/12)
and Δ𝑦 is the height difference between top and bottom planes in fig. 1.3 (m).
distance), and so this turns out as Pascal − seconds in si units. We will come
The dimension of viscosity is in (force per area) divided by (velocity per
difference Δ𝑣 between the top and bottom surface. The horizontal force 𝐹 required
Figure 1.3: A brick-shaped element of fluid is strained, by applying a velocity
on the top edge, divided by the area 𝐴, is the shear stress 𝜏 (see also §1.5.3 p. 18
further down). The higher the required shear stress for a given strain rate, the more
viscous the fluid is.
Figure CC-0 Pm.schroeder/Olivier Cleynen
17
1.5 Forces on fluids
Fluids are subjected to, and subject their surroundings and themselves to
forces. Identifying and quantifying those forces allows us to determine how
they will flow. Three types of forces are relevant in fluid dynamics: gravity,
pressure and shear.
1.5.1 Gravity
1.5.2 Pressure
The concept of pressurew can be approached with the following conceptual
the pressure 𝑝 will be the ratio of the perpendicular force 𝐹⟂ to the surface
experiment: if a flat solid surface is placed in a fluid at zero relative velocity,
area 𝐴:
𝐹⟂
𝑝≡
𝐴
(1/14)
where 𝑝 is the pressure (N m−2 or Pascals, 1 Pa ≡ 1 N m−2 );
𝐹⟂ is the component of force perpendicular to the surface (N);
and 𝐴 is the surface area (m2 ).
1.5.3 Shear
In the same thought experiment as above, the shearw , noted 𝜏 (greek letter
“tau”), expresses the efforts of a force parallel to a surface of interest:
𝐹∥
𝜏≡
𝐴
(1/15)
Mass floww is noted 𝑚̇ and represents the amount of mass flowing through
a chosen surface per unit time. When the velocity across the surface Video: how to deal with the ⟂
symbol when calculating mass
is uniform, it can be quantified as: flow
𝑚̇ = 𝜌𝑉⟂ 𝐴
by Olivier Cleynen (CC-by)
https://youtu.be/H5l–WlCZQ8
(1/16)
where 𝑚̇ is the mass flow (kg s−1 );
𝜌 is the fluid density (kg m−3 );
𝐴 is the area of the considered surface (m2 );
and 𝑉⟂ is the component of velocity perpendicular to the surface (m s−1 ).
𝑚̇ = 𝜌𝑉𝐴⟂ (1/17)
where 𝑉 is the flow speed (m s−1 );
and 𝐴⟂ is the area of a surface perpendicular to the flow velocity (m2 ).
Volume floww is noted ̇ and represents the volume of the fluid flowing
through a chosen surface per unit time. Much like mass flow, when the
velocity is uniform, it is quantified as:
𝑚̇
̇ = 𝑉⟂ 𝐴 = 𝑉𝐴⟂ =
𝜌
(1/18)
at a pressure 𝑝 is:
𝑚̇
(1/20)
= 𝑝
𝜌
(1/21)
If a fluid passes across a volume, the net power 𝑃̇ pressure, net required to
both enter and leave the volume may be expressed as
𝑚̇
𝑃̇ pressure, net = Δ𝑝
𝜌
(1/22)
19
Power to increase temperature is also a time rate of energy transfer. If
As we will see in chapter 2 (Analysis of existing flows with one dimension), fluid
flow involves many forms of energy changes. We will learn to combine and
compare them progressively.
1. Mass balance:
The total amount of matter at hand in a given phenomenon must
remain constant (since in fluid mechanics, we do not usually consider
nuclear reactions). This statement can be expressed as:
𝑚system = cst
d𝑚system
=0
d𝑡
(1/24)
law,w which states that the net force 𝐹⃗net ≡ Σ𝐹⃗ applying to any given
The momentum balance equation is a formulation of Newton’s second
d
𝐹⃗net = 𝑚𝑉⃗ )
d𝑡 (
(1/25)
plied about an axis. It states that the net moment 𝑀⃗ net, X ≡ Σ𝑀⃗X applied
where rotation about an axis, or moments (“twisting” efforts) are ap-
d
𝑀⃗ net, X = 𝑟⃗ ∧ 𝑚𝑉⃗ )
d𝑡 (
(1/26)
4. Balance of energy:
20 This equation, also known as the “first principle of thermodynamics”,w
states that the total amount of energy within an isolated system must
remain constant:
d𝐸isolated system
=0
d𝑡
(1/27)
In special cases, further equations are used to describe other phenomena af-
fecting the fluid flow (e.g. chemical reactions, or interaction between phases).
In most cases however, the four equations above are the only important equa-
tions written in fluid mechanics. We usually apply those balance statements
to our problem in either one of two ways:
• We may have information about a flow which already exists, and want
to calculate how fluid properties change as it flows through the area
of interest, and what the related forces are. In that case, we write
the equations in an integral form: we will do this in chapters 2 and 3
(Analysis of existing flows).
• We may instead wish to predict how the fluid is going to flow through
our zone of interest. In order to do this, we need to calculate flow
properties in an extensive manner, aiming to obtain vector fields for
the velocity and pressure everywhere, at all times. To this effect, we
write the equations in a differential form: we will do this in chapter 6
(Prediction of fluid flows).
Time dependence Flows which do not vary with time are called steady.
Steadiness is dependent on the chosen point of view: for example,
the air flow around an aircraft in cruise flight is seen to be steady
from within the aircraft, but is obviously highly unsteady from the
point of view of the air particles directly in the path of the airliner.
Steadiness is also dependent on the selection of a suitable time frame:
time-variations in a flow may be negligible if the time window used in
the study is short enough.
tive 𝜕/𝜕𝑡 or the total time derivative D/D𝑡 —concepts we will study
Mathematically, flows are steady when either the partial time deriva-
number (eq. 1/10 p. 16). Below 0,3 the flow is always incompressible.
To find out whether a gas flow is compressible, compute the Mach
of 104 or more, the flow is very likely to become turbulent over the
length or width of an obstacle in the flow). If the result is on the order
length 𝐿. By contrast, with [Re] of the order of 102 or less, the flow is
very likely to remain laminar over this length.
This crude quantification, of course, deserves more explanation —
we will be coming back to the Reynolds number in chapters 7 and
following.
23
1.10 Solved problems
(1/30)
What is the force resulting from fluid pressure on each side of the plate?
incorrect! The correct result is 1,2124 MN. The method, numbers, equations etc.
Note: Unfortunately Olivier made an error in this video: the end computation is
are all ok — only the final result is affected. Many thanks to the students who
double-checked and reported the problem!
24
A sailboat travels at velocity 𝑉⃗
Power and moment resulting from a force
25
26
Problem sheet 1: Basic flow quantities
last edited September 3, 2020
by Olivier Cleynen — https://fluidmech.ninja/
The atmosphere has 𝑝atm. = 1 bar; 𝜌atm. = 1,225 kg m−3 ; 𝑇atm. = 11,3 °C; 𝜇atm. = 1,5 ⋅ 10−5 Pa s
Except otherwise indicated, assume that:
Air behaves as a perfect gas: 𝑅air =287 J kg−1 K−1 ; 𝛾air =1,4; 𝑐𝑝 air =1 005 J kg−1 K−1 ; 𝑐𝑣 air =718 J kg−1 K−1
Liquid water is incompressible: 𝜌water = 1 000 kg m−3 , 𝑐𝑝 water = 4 180 J kg−1 K−1
Figure 1.6: Pressure distribution on a flat panel that is part of the wall of a swimming pool
Figure CC-0 Olivier Cleynen
1.3.1. What is the pressure force (i.e. the force resulting from the pressure) exerted on
the left side of the plate?
[Hint: we will explore the required expression in chapter 5 (Effects of shear) as eq. 5/3
p. 92]
The cannon is shot 8,4 km away from Newton. What is the air temperature if the
smoke produced by a cannon blast and the hearing of the detonation.
Figure 1.8: Top view of a truck traveling at velocity 𝑉⃗ and subject to a aerodynamic force 𝐹⃗
Figure CC-0 Olivier Cleynen
28
1.6.1. What is the power given by the wind to the truck?
The force 𝐹⃗ is applying at a distance 0,8 m behind the center of gravity of the truck.
cylindrical outlet pipe, whose outlet is slanted at an angle 𝜃 = 25° to improve the good
The engine exhaust gases of a student’s hot-rod car are flowing quasi-steadily in a
Figure 1.9: Exhaust gas pipe of a car. The outlet cross-section is at an angle 𝜃 relative to the axis
of the pipe.
Figure CC-0 Olivier Cleynen
Photo cropped, mirrored and edited from an original CC-by-sa by kazandrew2
The outlet velocity is measured at 15 m s−1 , and the exhaust gas density is 1,1 kg m−3 . The
slanted outlet section area 𝐴 is 420 cm2 .
pressure loss of 21 Pa – we will learn to quantify this in chapter 7 (Pipe flows). In these
Because of the shear within the exhaust gases, the flow through the pipe induces a
conditions, the specific heat capacity of the exhaust gases is 𝑐𝑝gases = 1 100 J kg−1 K−1 .
1.7.3. What is the power required to carry the exhaust gases through the pipe?
1.7.4. What is the gas temperature increase due to the shear in the flow? 29
1.8 Acceleration of a particle
particle of width 0,1 mm. The particle is accelerating at a rate of 2,5 m s−2 .
Inside a complex, turbulent water flow, we are studying the trajectory of a cubic fluid
30
Answers
1.2 If you adopt [Ma] = 0,6 as an upper limit, you will obtain 𝑉max = 709 km h−1 (eqs. 1/10
& 1/11 p. 16). Note that propellers, fan blades etc. will meet compressiblity effects
far sooner.
1.3 1) 𝐹left = 400 kN (eq. 1/14 p. 18); 2) 𝐹right = 480 kN (eq. 4/3 p. 74).
1.6 1) 𝑊̇ = 𝐹⃗aero ⋅ 𝑉⃗truck = 65,3 kW. See Appendix A2.1 p. 247 for a short briefing about
⎛ 0 ⎞
the dot product of vectors;
1.8 1) 𝐹net = 2,5 ⋅ 10−9 N (eq 1/25 p. 20), such are the orders of magnitude involved in
cfd calculations!
2) Only three kinds: forces due to pressure, shear, and gravity.
1.9 1) yes, 2) yes if [Re] is high enough, 3) yes (in very specific cases such as high
pressure changes combined with high heat transfer or high irreversibility, therefore
generally no), 4) open the cap of a water bottle and turn it upside down: you have
an isothermal, unsteady, incompressible flow. An example of compressible flow
could be the expansion in a jet engine nozzle.
31
32
Fluid Dynamics
Chapter 2 – Analysis of existing flows
with one dimension
last edited May 22, 2020
by Olivier Cleynen — https://fluidmech.ninja/
2.1 Motivation 33
2.2 One-dimensional flow problems 33
2.3 Balance of mass 35
2.3.1 Mass balance equation 35
2.3.2 Problems with the mass balance equation 36
2.4 Balance of momentum 37
2.5 Balance of energy 39
2.6 The Bernoulli equation 41
2.6.1 Theory 41
2.6.2 Reality 42
2.7 Solved problems 44
2.8 Problems 45
These notes are based on textbooks by White [22], Çengel & al.[25], Munson & al.[29], and de Nevers [17].
2.1 Motivation
In this chapter, we learn to analyze fluid flows for which a lot of information
is already available. We want, when confronted to a simple flow (for example,
flow entering and leaving a machine), to be able to answer three questions:
2. At inlet and outlet, the fluid properties are uniform, so that they can be
evaluated in bulk (e.g. the inlet has only one velocity, one temperature
etc.);
Providing that those conditions are met, we can answer the question: what is
the net effect of the fluid flow through the considered volume?
In order to write useful equations, we need to begin with rigorous definitions,
with the help of figure 2.1:
The three equations that we write in this chapter state that basic physical
laws apply to the system. They are balance equations (see §1.7 p. 20). Each
time, we will express what is happening to the system, as a function of
the fluid properties at the inlet and outlet of the control volume. This
will allow us to answer three questions:
• What is the mass flow entering and leaving the control volume?
• What is the force required to move the flow through the control vol-
ume?
• What energy transfer is required to move the flow through the control
volume?
Figure 2.1: A control volume within a flow. The system is the amount of mass
included within the control volume at a given time. Because mass enters and leaves
the control volume, the system is being moved and deformed (bottom).
Figure CC-0 Olivier Cleynen
34
2.3 Balance of mass
=0= +
the rate of change of the rate of change the net mass flow
the fluid’s mass of mass inside at the borders
as it transits the considered volume of the considered volume
= +
the sum of the sum of
time rate of creation
incoming mass flows outgoing mass flows
or destruction of mass
(negative terms) (positive terms)
0 = 𝑚̇ in 1 + 𝑚̇ in 2 + 𝑚̇ out 1 + 𝑚̇ out 2
̇ in 1 − |𝑚|
0 = − |𝑚| ̇ in 2 + |𝑚|
̇ out 1 + |𝑚|
̇ out 2
We can substitute 𝑚̇ = 𝜌𝑉⟂ 𝐴 (eq. 1/16 p. 19) into the equations above, obtain-
ing:
Looking again at an example case where there were two inlets and two
outlets, this equation 2/3 would become:
0 = 𝜌in 1 𝑉⟂ in 1 𝐴in 1 + 𝜌in 2 𝑉⟂ in 2 𝐴in 2 + 𝜌out 1 𝑉⟂ out 1 𝐴out 1 + 𝜌out 2 𝑉⟂ out 2 𝐴out 2
0 = (𝜌𝑉⟂ 𝐴)in 1 + (𝜌𝑉⟂ 𝐴)in 2 + (𝜌𝑉⟂ 𝐴)out 1 + (𝜌𝑉⟂ 𝐴)out 2
0 = − (𝜌|𝑉⟂ |𝐴)in 1 − (𝜌|𝑉⟂ |𝐴)in 2 + (𝜌|𝑉⟂ |𝐴)out 1 + (𝜌|𝑉⟂ |𝐴)out 2
For steady flow through a fixed considered volume with two inlets and two outlets.
In a simple case where there is only one inlet and one outlet, this last equation
can be rewritten as
mis-use this equation is to draw the conclusion that “if 𝐴 decreases, then 𝑉
The equation 2/4 above is interesting, but also treacherous. The best way to
must increase”. This is only true some of the time, and here are two reasons
why:
1. The density 𝜌 may change between inlet and outlet. In low-speed flows
Video: Dangers associated with
without heat transfer, 𝜌 does not vary significantly (see §1.8 p. 1.8).
the mass balance equation
by Olivier Cleynen (CC-by)
https://youtu.be/M1Dw6qU-FOU
creases in velocity are not “for free”: they require force be applied and
energy be spent. The mass balance equation cannot account for those
phenomena.
36
2.4 Balance of momentum
What force is applied to the fluid for it to travel through the control volume?
We answer this question by writing a momentum balance equation. It
compares the rate of change of the system’s momentum (which by definition
is the net force applying to it, see eq. 1/25 p. 20), to the flow of momentum
through the borders of the control volume:
d(𝑚𝑉⃗sys ) d
= 𝐹⃗net = 𝑚𝑉⃗ ) + ⃗
(𝑚̇ 𝑉 )net
d𝑡 d𝑡 ( CV
(2/5)
= 𝐹⃗net = +
the rate of change of the rate of change the net flow of momentum
the fluid’s momentum of momentum through the boundaries
as it transits within the considered volume of the considered volume
Since we are here interested only in steady flows, and we have clearly-
identified inlets and outlets, this becomes:
= +
the vector sum the sum of incoming the sum of outgoing
of forces momentum flows momentum flows
on the fluid (with negative 𝑚̇ terms) (with positive 𝑚̇ terms)
The same convention as above is applied for the sign of the mass flow 𝑚. ̇
For example, in a case where there were one inlet and one outlet, we would
write:
𝐹⃗net = ⃗
(𝑚̇ 𝑉 )in + (𝑚̇ 𝑉⃗ )
𝐹⃗net = − (|𝑚| ̇ 𝑉⃗ )
̇ 𝑉⃗ ) + (|𝑚|
out
in out
For steady flow through a fixed considered volume with one inlet and one outlet.
As before, we can substitute 𝑚̇ = 𝜌𝑉⟂ 𝐴 (eq. 1/16 p. 19) into the equations
above, obtaining:
In the example case where there is one inlet and one outlet, we would write:
𝐹⃗net = ⃗
(𝜌𝑉⟂ 𝐴𝑉 )in + (𝜌𝑉⟂ 𝐴𝑉⃗ )
in out
volume is traversed by a steady flow with mass flow 𝑚,̇ with one inlet (point 1)
To make clear a few things, let us focus on the simple case where a considered
⃗
and one outlet (point 2). The net force 𝐹net applying on the fluid is
𝐹⃗net = |𝑚|
̇ (𝑉⃗2 − 𝑉⃗1 ) (2/8)
37
Three remarks can be made about this equation. First, we need to be aware
⎧
⎪ 𝐹 = |𝑚|̇ (𝑉2𝑥 − 𝑉1𝑥 )
⎪ net𝑥
⎨ 𝐹net𝑦 = |𝑚|̇ (𝑉2𝑦 − 𝑉1𝑦 )
⎪
⎪
Video: The net force thing in the
𝐹
⎩ net𝑧 = | ̇
𝑚| (𝑉2𝑧 − 𝑉1𝑧 )
momentum balance equation (2/9)
by Olivier Cleynen (CC-by)
https://youtu.be/AOcGNeY9ad0
Second, there are two reasons why we could calculate a non-zero net force
on the fluid in equation 2/8, as illustrated in fig. 2.2.
1. Even if 𝑉⃗2 is aligned and in the same direction as 𝑉⃗1 , they can be of
different magnitude. A force is required to accelerate or decelerate
the fluid (more precisely, an acceleration or deceleration of the fluid is
equivalent to a force);
2. Even if 𝑉⃗2 has the same magnitude as 𝑉⃗1 , they can have different
directions. A force is required to change the direction in which a flow
is flowing (or more precisely, a change of direction is equivalent to a
force).
̇ (𝑉2𝑥 − 𝑉1𝑥 )
𝐹net𝑥 = |𝑚| (2/10)
The final remark is that the equation does not describe a cause-effect re-
lationship. The net force does not cause the change in velocity any more
and simultaneous. Similarly, we have no way to know what 𝐹⃗net is made of.
than the change in velocity causes the net force: they are both equivalent
The exact mechanism which adds up to a net force (pressure change, shear
applied through a static wall, the movement of a turbine, etc.) is “hidden” in
the control volume, and unknown to us. In order to find out what happens
in the control volume, we need a different type of analysis, which we will
approach in chapter 6 (Prediction of fluid flows).
Figure 2.2: Two reasons can explain why a net force 𝐹⃗net appears in eq. 2/8. On
the left, 𝑉⃗2 and 𝑉⃗1 are aligned, but have different lengths. On the right 𝑉⃗2 and 𝑉⃗1
have the same length, but different directions. We will look at the second case in
chapter 3 (Analysis of existing flows with three dimensions).
Figure CC-0 Olivier Cleynen
38
2.5 Balance of energy
What power is applied to the fluid for it to travel through the control volume?
We answer this question by writing an energy balance equation. It compares
the rate of change of the system’s energy, to the flow of energy through the
For this we prefer to express the energy 𝐸 as the specific energy 𝑒 (in J kg−1 )
borders of the control volume.
= = +
the rate of change of the sum of the rate of change the net flow of energy
the fluid’s energy powers as of energy within through the boundaries
as it transits heat and work the considered volume of the considered volume
• the net power transferred as work with moving solid surfaces 𝑊̇ surfaces, net
(for example, a moving piston, turbine blade, or rotating shaft, positive
inwards);
• and the net power transferred to and from the fluid by the fluid itself,
in order to enter and leave the considered volume. This power is called
Turning now to the specific energy 𝑒, we break it down into three components
(see also §1.3.3 p. 13):
• the specific internal energy 𝑖, which represents the energy per unit
ics, this is often noted 𝑢, but in fluid mechanics we reserve this symbol
mass contained as stored heat within the fluid itself. In thermodynam-
𝑒𝑝 ≡ 𝑔𝑧 (2/15)
𝑒 ≡ 𝑖 + 𝑒𝑘 + 𝑒𝑝 (2/16)
Now, we focus on steady flows (for which energy in the control volume does
not change with time), and we can come back to eq. 2/11 to rewrite it as:
𝑝 1
𝑄̇ net + 𝑊̇ shaft, net = Σ 𝑚̇ 𝑖 + + 𝑉 2 + 𝑔𝑧
[ ( 𝜌 2 )]in
𝑝 1
+Σ 𝑚̇ 𝑖 + + 𝑉 2 + 𝑔𝑧
[ ( 𝜌 2 )]out
(2/18)
As usual, let us focus on a case where there is only one inlet and one outlet.
We obtain:
𝑝 1 𝑝 1
𝑄̇ net + 𝑊̇ shaft, net = − |𝑚|̇ 𝑖 + + 𝑉 2 + 𝑔𝑧 + |𝑚|̇ 𝑖 + + 𝑉 2 + 𝑔𝑧
[ ( 𝜌 2 )]in [ ( 𝜌 2 )]out
𝑝 1
𝑄̇ net + 𝑊̇ shaft, net = |𝑚|
̇ Δ𝑖 + Δ + Δ ( 𝑉 2 ) + Δ(𝑔𝑧)
[ 𝜌 2 ]
(2/19)
This equation is very useful in principle, but not so much in practice, for two
reasons:
1. It contains a lot of terms. There are five fluid properties at inlet and
outlet which affect energy, and it is difficult to predict which one will
40
Advice from an expert
Again, remember the title of the chapter. To cal-
culate the value of any one property in equa-
tion 2/18, you need to input the value of the eleven
other ones. It is tempting to take shortcuts while
doing so (“oh, the pressure is probably the same”),
with disastrous consequences. There is no solu-
tion to this. If you are attempting to predict fluid
flow, and are missing information, better stop without a result than
take hazardous attempts at using equation 2/18.
practice this only works when very high powers are involved, such
as in a compressor or in a rocket engine nozzle. For ordinary flow
(say, air flow around a car, or water flow in a pipe), the temperature
changes are much too small to be measured. See exercise 2.3 p. 46
for an example of this.
2.6.1 Theory
The Bernoulli equation is the energy equation applied to specific cases.
To derive the Bernoulli equation, we will start from equation 2/18 and add
five constraints:
d(𝑚𝑒)CV
1. Steady flow.
d𝑡
(We had already implemented this restriction, when we set from
eq. 2/11 to zero in order to obtain eq. 2/18.)
41
Thus, 𝜌 stays constant;
2. Incompressible flow.
5. One-dimensional flow.
Thus, our considered volume has only one inlet (labeled 1) and one
outlet (labeled 2): all fluid particles move together with the same transit
time, and the overall trajectory is already known.
𝑝 1
0+0= 𝑚̇ 𝑖cst. + + 𝑉 2 + 𝑔𝑧
[ ( 𝜌cst. 2 )]1
𝑝 1
+ 𝑚̇ 𝑖cst. + + 𝑉 2 + 𝑔𝑧
[ ( 𝜌cst. 2 )]2
𝑝 1 𝑝 1
0 = − 𝑖cst. + + 𝑉 2 + 𝑔𝑧 + 𝑖cst. + + 𝑉 2 + 𝑔𝑧
( 𝜌cst. 2 )1 ( 𝜌cst. 2 )2
𝑝 1 𝑝 1
0=− + 𝑉 2 + 𝑔𝑧 + + 𝑉 2 + 𝑔𝑧
( 𝜌cst. 2 )1 ( 𝜌cst. 2 )2
2.6.2 Reality
Let us insist on the incredibly frustrating restrictions brought by the five
conditions above:
1. Steady flow.
Video: The Bernoulli equation This constrains us to continuous flows with no transition effects, which
will kill you
by Olivier Cleynen (CC-by)
is a reasonable limit;
https://youtu.be/4RBxVepjcQc
2. Incompressible flow.
4. No friction.
This is a tragic restriction! We cannot use this equation to describe a
42 turbulent or viscous flow, e.g. near a wall or in a wake.
5. One-dimensional flow.
This equation is only valid if we know precisely the trajectory of the
fluid whose properties are being calculated.
There are indeed cases where the pressure losses due to the imperfection
of the flow are well-understood, and can be easily quantified. This is true
of flow in pipes, for example (we study those in chapter 7). In those cases,
eq. 2/21 is extremely useful.
Nevertheless, this approach is also easily misused. In a fluid flow where
several of the restrictions above do not hold —and many such flows can be
found in everyday life as well as engineering applications— equation 2/21
will betray its users. Convince yourself that any wrong equation can be
2 + 3 = −18 + Δ𝑝loss .
made correct by adding an unknown “bucket” term at the end: for example
43
2.7 Solved problems
Flow in a nozzle
What is the mass flow? What is the outlet velocity? And what is the
pressure change across the pipe?
mass flow entering the pipe is 2 kg s−1 , and it enters the pipe with a
Water is flowing through a straight pipe with constant diameter. The
What is the outlet velocity? What is the net force on the fluid as it
transits? What is the power dissipated as friction?
44
Problem sheet 2: Analysis of existing flows
with one dimension
last edited April 15, 2021
by Olivier Cleynen — https://fluidmech.ninja/
The atmosphere has 𝑝atm. = 1 bar; 𝜌atm. = 1,225 kg m−3 ; 𝑇atm. = 11,3 °C; 𝜇atm. = 1,5 ⋅ 10−5 Pa s
Except otherwise indicated, assume that:
Air behaves as a perfect gas: 𝑅air =287 J kg−1 K−1 ; 𝛾air =1,4; 𝑐𝑝 air =1 005 J kg−1 K−1 ; 𝑐𝑣 air =718 J kg−1 K−1
Liquid water is incompressible: 𝜌water = 1 000 kg m−3 , 𝑐𝑝 water = 4 180 J kg−1 K−1
𝑝 1
𝑄̇ net + 𝑊̇ shaft, net = Σ 𝑚̇ 𝑖 + + 𝑉 2 + 𝑔𝑧
[ ( 𝜌 2 )]in
𝑝 1
+Σ 𝑚̇ 𝑖 + + 𝑉 2 + 𝑔𝑧
[ ( 𝜌 2 )]out
(2/18)
is 8 cm, the water arrives with a uniform velocity of 1,5 m s−1 . The diameter increases
Water flows from left to right in a pipe, as shown in fig. 2.3. On the left, the diameter
gently until it reaches 16 cm; the expansion is smooth, so that losses (specifically, energy
losses due to wall friction and flow separation) are negligible.
2.2.1. What are the mass and volume flows at inlet and outlet?
2.2.2. What is the average velocity of the water at the right end of the expansion?
45
Figure 2.3: A simple pipe expansion, with water flowing from left to right.
CC-0 Olivier Cleynen
2.2.3. What is the pressure change in the water across the expansion?
The water is drained from the pipe, and instead, air with density 1,225 kg m−3 is flowed
in the pipe, incoming with a uniform velocity of 1,5 m s−1 .
of the pipe length (fig. 2.4). Water arrives the pipe with a uniform velocity of 1,5 m s−1
Water flows in a long pipe which has constant diameter; a valve is installed in the middle
using the dimensionless loss coefficient 𝐾valve (we later will later encounter it as eq. 7/6
The pipe itself and the valve, together, induce a pressure loss which can be quantified
p. 137). With this tool, the pressure loss Δ𝑝valve is related to the mean incoming speed
𝑉incoming as:
|Δ𝑝valve |
𝐾valve ≡ = 2,6
1
𝜌𝑉incoming
2
(2/22)
2
2.3.4. If the heat losses of the pipe and valve are negligible, what is the temperature
increase of the water?
Figure 2.4: A simple, straight pipe, featuring a partially-open valve in the center
CC-0 Olivier Cleynen
46
2.4 Combustor from a jet engine
A jet engine is equipped with several combustors (sometimes also called combustion
chambers). We are interested in fluid flow through one such combustor, shown in fig. 2.5.
Air from the compressor enters the combustor, is mixed with fuel, and combustion occurs,
which greatly increases the temperature and specific volume of the mix, before it is run
through the turbine.
The conditions at inlet are as follows:
At the outlet, the hot gases have pressure 24,5 bar and temperature 1 550 °C, and exit
with a speed of 50 m s−1 .
2.4.4. What is the net force exerted on the gas as it travels through the combustor?
Figure 2.5: A combustor in a sectioned jet engine (here, a Turboméca Adour). Air enters from the
left, out of the compressor (whose blades are painted blue). It leaves the combustor on the right
side, into the turbine. In the combustor, high-temperature, steady combustion takes place.
Photo CC-by-sa Olivier Cleynen
47
2.5 Water jet on a truck
A water nozzle shoots water towards the back of a small stationary van. It has a 3 cm2
cross-sectional area, and the water speed at the nozzle outlet is 𝑉jet = 20 m s−1 . As the
horizontal water jet hits the back of the van, it is split in two symmetrical vertical
(𝑉2 = 𝑉3 = 20 m s−1 ).
flows (fig. 2.6). The two opposite vertical jets have same mass flow and same velocity
2.5.1. What is the net force exerted on the water by the truck?
2.5.2. What is the net force exerted on the truck by the water?
𝑉truck = 15 m s−1 .
Now, the truck moves longitudinally in the same direction as the water jet, with a speed
(This is a crude conceptual setup, which allows us to approach conceptually the case
where water acts on the blades of a turbine.)
2.5.3. What is the new force exerted by the water on the truck?
2.5.5. How would the power be modified if the volume flow was kept constant, but
the diameter of the nozzle was reduced? (briefly justify your answer, e.g. in 30
words or less)
Figure 2.6: A water jet flowing out of a nozzle (left), and impacting the vertical back surface of a
small electric truck, on the right.
Figure CC-0 Olivier Cleynen
48
2.6 High-speed gas flow
Scientists build a very high-speed wind tunnel. For this, they build a large compressed air
tank. Air escapes from the tank into a pipe which decreasing cross-section, as shown in
fig. 2.7. The pipe diameter reaches a minimum (at the tunnel throat), and then it expands
again, before discharging into the atmosphere.
Figure 2.7: A converging-diverging nozzle. Air flows from the left tank to the right outlet, with a
contraction in the middle.
Figure CC-0 Olivier Cleynen
For simplicity, we assume that heat losses through the tunnel walls are negligible, and
In the tank (point 1), the air is stationary, with pressure 7,8 bar and temperature 246,6 °C.
that the fluid has uniformly-distributed velocity in cross-sections of the pipe.
At the throat (point 2), the pressure and temperature have dropped to 4,2 bar and 160 °C.
The velocity has reached 417,2 m s−1 . The throat cross-section is 0,01 m2 .
the air has seen its pressure and temperature drop to 1,38 bar and 43 °C.
Downstream of the throat, the pressure keeps dropping. By the time it reaches a point 3,
(if you need to convince yourself that 𝐴3 > 𝐴1 , you may also calculate the
2.6.3. What is the fluid velocity at point 3?
cross-section area)
2.6.4. What is the net force exerted on the fluid between the points 2 and 3?
2.6.5. What is the kinetic energy per unit mass of the air at point 3?
Once it has passed point 3, the air undergoes complex loss-inducing evolutions (including
discharges into the atmosphere (point 4) with pressure 1 bar and temperature 165 °C.
going through a shock wave, where its properties change very suddenly), before it
2.2.1 At both inlet and outlet, 𝑚̇ = 7,53 kg s−1 and ̇ = 7,53 L s−1
2.2.2 𝑉2 = 0,375 m s−1
2.2.3 Δ𝑝1→2 = +1 054 Pa
2.2.4 Δ𝑝3→4 = +4 218 Pa
2.2.5 Δ𝑝5→6 = +1,29 Pa
2.3 p. 46
̇ = +215,37 W
2.3.3 𝑚Δ𝑖
2.3.4 With eq. 2/18, Δ𝑇 = +0,7 mK (very small!)
2.4 p. 47
2.5 p. 48
2.6 p. 49
3.1 Motivation 51
3.2 The Reynolds transport theorem 51
3.2.1 Control volume 51
3.2.2 Rate of change of an additive property 52
3.3 Balance of mass 54
3.4 Balance of momentum 56
3.5 Balance of angular momentum 57
3.6 Balance of energy 59
3.7 Limits of integral analysis 59
3.8 Solved problems 60
3.9 Problems 63
These notes are based on textbooks by White [22], Çengel & al.[25], Munson & al.[29], and de Nevers [17].
3.1 Motivation
In this chapter, we use the same tools that we developed in chapter 2, but we
improve them so we can apply them to more complex cases. Specifically, we
would like to answer the following questions:
1. What are the mass flows and forces involved when a flow has non-
uniform velocity?
2. What are the forces and moments involved when a flow changes direc-
tion?
we will write equations that work inside any generic velocity field 𝑉⃗ =
are no longer limited to one-inlet, one-oulet steady-flow situations. Instead,
through the control volume at the exact time when we write the equation,
and its properties (volume, pressure, velocity etc.) may change in the process.
All along the chapter, we are focusing on the question: based on measured
fluid properties at some chosen area in space and time (the properties at the
control surface), how can we quantify what is happening to the system (the
mass inside the control volume)?
term additive property, we mean that the total amount of property is divided
if the fluid is divided. For instance, this works for properties such as mass,
with compressed air): we measure this with the term d𝐵CV / d𝑡.
accumulation (for example, mass may be increasing in an air tank fed
We can now link these three terms with the simple equation:
d𝐵sys d𝐵CV
= + 𝐵̇ net
d𝑡 d𝑡
(3/1)
d𝑡
(3/3)
The second term of eq. 3/1, 𝐵̇ net , can be evaluated by quantifying, for each
area element d𝐴 of the control volume’s surface, the surface flow rate 𝜌𝑏𝑉⟂
of property 𝐵 that flows through it, as shown in fig. 3.2. The integral over
the entire control volume surface CS of this term is:
control surface with area d𝐴. The 𝑛⃗ vector defines the orientation of d𝐴 surface,
Figure 3.2: Part of the system may be flowing through an arbitrary piece of the
d𝐵sys d
= 𝜌𝑏 d + ∬ 𝜌𝑏 (𝑉⃗rel ⋅ 𝑛)
⃗ d𝐴
d𝑡 d𝑡 ∭CV
(3/5)
CS
variables, it will prove extremely useful, allowing us to quantify the net effect
of the flow of a system through a volume for which border properties are
known.
In the following sections we are going to use this equation to write out four
key balance equations (see §1.7 p. 20):
• balance of mass;
• balance of energy.
d𝐵/ d𝑡 becomes d𝑚sys / d𝑡, which by definition is zero (see eq. 1/24 p. 20). Also,
𝑏 ≡ 𝐵/𝑚 = 𝑚/𝑚 = 1 and now the Reynolds transport theorem becomes:
d𝑚sys d
=0= 𝜌 d + ∬ 𝜌 (𝑉⃗rel ⋅ 𝑛)
⃗ d𝐴
d𝑡 d𝑡 ∭CV CS
(3/6)
=0= +
the rate of change the rate of change the net mass flow
of the fluid’s mass of mass inside at the borders
as it transits considered volume of the considered volume
54
This equation 3/6 is often called continuity equation. It allows us to compare
the incoming and outgoing mass flows through the borders of the control
volume.
0= 𝜌 d + ∑ {𝜌𝑉⟂ 𝐴} + ∑ {𝜌𝑉⟂ 𝐴}
tinuing to breathe, a technique
d𝑡 ∭CV
called circular breathing.w Can
(3/7) you identify the different terms
d
of eq. 3/8 as they apply to the
d𝑡 ∭CV
saxophonist’s mouth?
by David Hernando Vitores (CC-by-sa)
https://frama.link/vyH-cxCL
d
0= ̇ − ∑ {|𝑚|}
𝑚CV + ∑ {|𝑚|} ̇
out in
d𝑡
(3/8)
out in
inlet and outlet. Here eq. 3/6 translates as 0 = d𝑡d ∭CV 𝜌 d + 𝜌3 |𝑉⟂3 |𝐴3 + 𝜌2 |𝑉⟂2 |𝐴2 −
Figure 3.3: A control volume for which the system’s properties are uniform at each
𝜌1 |𝑉⟂1 |𝐴1 .
As before, in equation 3/7, the term 𝜌𝑉⟂ 𝐴 at each inlet or outlet corresponds
Figure CC-0 Olivier Cleynen
to the local mass flow ±𝑚̇ (negative inwards, positive outwards) through the
boundary.
Second, there is an unsteady term d𝑚CV / d𝑡. It is not used to solve prob-
have a nice smooth uniform outlet!
lems in this course, but one day when you are confronted to a case where
your inlet and outlet mass flows are not equal, it will save your day!
55
3.4 Balance of momentum
What force is applied to the fluid for it to travel through the control volume?
We answer this question by writing out a mass balance equation in the
that d𝐵/ d𝑡 becomes d𝑚𝑉⃗sys )/ d𝑡, which by definition is the net force 𝐹⃗net
applying to the system (see eq. 1/25 p. 20). Also, 𝑏 ≡ 𝐵/𝑚 = 𝑚𝑉⃗ /𝑚 = 𝑉⃗ and
Abstruse Goose #338: Newton’s
laws of motion almost didn’t
happen
by an anonymous artist (CC-by-nc) now the Reynolds transport theorem becomes:
https://abstrusegoose.com/338
d(𝑚𝑉⃗sys ) d
= 𝐹⃗net = 𝜌 𝑉⃗ d + ∬ 𝜌 𝑉⃗ (𝑉⃗rel ⋅ 𝑛)
⃗ d𝐴
d𝑡 d𝑡 ∭CV CS
(3/9)
= +
the vector sum the rate of change the net flow of momentum
of forces of momentum through the boundaries
on the fluid within the considered volume of the considered volume
In that case equation 3/9 reduces to forms that we have already identified in
{ } { }
the previous chapter (see §2.4 p. 37):
d
𝐹⃗net = 𝜌 𝑉⃗ d + ∑ (𝜌|𝑉⟂ |𝐴)𝑉⃗ − ∑ (𝜌|𝑉⟂ |𝐴)𝑉⃗
d𝑡 ∭CV
Video: Making sense of the 3D
linear momentum balance equa-
{ } { }
(3/10)
d
tion
= ⃗
𝑚𝑉 ) + ∑ |𝑚| ⃗
̇ 𝑉 − ∑ |𝑚| ⃗
̇ 𝑉
by Olivier Cleynen (CC-by) out in
d𝑡 (
https://youtu.be/iDCpqoJJSI4
(3/11)
CV out in
are uniform at each inlet and outlet, eq. 3/9 translates as 𝐹⃗net = d𝑡d ∭CV 𝜌 𝑉⃗ d +
Figure 3.4: The same control volume as in fig. 3.3. Here, since the system’s properties
56
To make clear a few things, let us focus on the simple case where a con-
equation 3/9, the net force 𝐹⃗net applying on the fluid is:
sidered volume has only one inlet (point 1) and one outlet (point 2). From
d
𝐹⃗net = 𝜌 𝑉⃗ d + ∬ 𝜌2 |𝑉⟂2 |𝑉⃗2 d𝐴2 − ∬ 𝜌1 |𝑉⟂1 |𝑉⃗1 d𝐴1
d𝑡 ∭CV
Video: as a person walks, the de-
flection of the air passing around
their body can be used to sustain
(3/12) the flight of a paper airplane (a
• The first term, which we could informally write as d(𝑚𝑉⃗ )CV / d𝑡, could
ume surrounding the glider, and
the resulting net force?
by Y:sciencetoymaker (styl)
https://youtu.be/S6JKwzK37_8
within the control volume is changing, such as when the fluid in a tank
sloshes back and forth against the walls.
̇ 2 𝑉⃗2 − |𝑚|
|𝑚| ̇ 1 𝑉⃗1 , could also be non-zero. This happens when the flux
• The sum of the last two terms, which we could informally write as
As you can see, a lot of different things may be happening at once! We will
study (separately) the most relevant of those effects in the problem section of
this chapter.
What moment (“twisting effort”) is applied to the fluid for it to travel through
the control volume? We answer this question by writing an angular momen-
tum balance (see eq. 1/26 p. 20) in the template provided by the Reynolds
All positions are measured with a position vector 𝑟⃗X𝑚 . We now state that
Video: rocket landing gone
d𝐵/ d𝑡 becomes d⃗𝑟X𝑚 ∧ 𝑚𝑉⃗sys / d𝑡, which by definition is the net moment 𝑀⃗ net
thruster around the base of the
rocket as it (unsuccessfully) at-
applying to the system (see again eq. 1/26 p. 20). Also, 𝑏 ≡ 𝐵/𝑚 = 𝑟⃗X𝑚 ∧ 𝑉⃗ and
tempts to compensate for the
collapsed landing leg?
by Y:SciNews (styl)
https://youtu.be/4cvGGxTsQx0
57
d(⃗𝑟X𝑚 ∧ 𝑚𝑉⃗ )sys d
= 𝑀⃗ net,X = 𝑟⃗X𝑚 ∧ 𝜌 𝑉⃗ d + ∬ 𝑟⃗X𝑚 ∧ 𝜌 (𝑉⃗rel ⋅ 𝑛)
⃗ 𝑉⃗ d𝐴
d𝑡 d𝑡 ∭CV CS
(3/13)
= +
the vector sum the rate of change of the net flow of angular momentum
of moments the angular momentum through the boundaries
on the fluid within the considered volume of the considered volume
in which 𝑟⃗X𝑚 is a vector giving the position of any mass 𝑚 relative to point 𝑋 .
{ } { }
In that case equation 3/13 reduces to a more readable form:
⃗ d ⃗ ⃗ ⃗
𝑀net,X = 𝑟⃗X𝑚 ∧ 𝜌 𝑉 d + ∑ 𝑟⃗X𝑚 ∧ |𝑚| ̇ 𝑉 − ∑ 𝑟⃗X𝑚 ∧ |𝑚| ̇ 𝑉
d𝑡 ∭
Video: Making sense of the an-
gular momentum balance equa- CV out in
tion
by Olivier Cleynen (CC-by) (3/14)
https://youtu.be/VR8LGr6PuRY
Figure 3.5: A control volume for which the properties of the system are uniform
𝑀⃗ net,X = d𝑡d ∭CV 𝑟⃗X𝑚 ∧ 𝜌 𝑉⃗ d + 𝑟⃗2 ∧ |𝑚̇ 2 |𝑉⃗2 − 𝑟⃗1 ∧ |𝑚̇ 1 |𝑉⃗1 .
at each inlet or outlet. Here the moment about point X in the bottom right is
The same remarks we made for linear momentum earlier apply here: there
are many possible phenomena which may equate to a moment on the fluid.
We will explore this equation rather shyly in the problem section of the
chapter.
d𝐸sys d
= 𝑄̇ net + 𝑊̇ shaft, net = 𝜌 𝑒𝑓 d + ∬ 𝜌 𝑒𝑓 (𝑉⃗rel ⋅ 𝑛)
⃗ d𝐴
d𝑡 d𝑡 ∭CV CS
(3/15)
with the term 𝑒𝑓 carrying all of the energy terms relevant for us in fluid
𝑝 1 2
mechanics:
𝑒𝑓 ≡ 𝑖 + + 𝑉 + 𝑔𝑧
𝜌 2
(3/16)
It can be seen here that this equation only differs from equation 2/18 in the
last chapter in the two following ways:
• The energy flows through the inlet and outlet are expressed as integrals,
allowing us to account for non-uniform distributions.
Save for those two differences, the equation is not any different — and, it
must be admitted, not much more useful in practice. In this course, we will
not be using this equation to solve problems.
• First, we are confined to calculating the net effect of fluid flow. The net
force, for example, encompasses the integral effect of all forces —due to
pressure, shear, and gravity— applied on the fluid as it transits through
the control volume. Integral analysis gives us absolutely no way of
distinguishing between those sub-components. In order to do that (for
example, to calculate which part of a pump’s mechanical power is lost
to internal viscous effects), we would need to look within the control
volume. 59
direction. The value d𝐵sys / d𝑡 of any finite integral cannot be used to
• Second, all four of our equations in this chapter only work in one
find which function 𝜌𝑏𝑉⟂ d𝐴 was integrated over the control surface to
which will result in a net force of +12 N. Knowing the net value of an
obtain it. For example, there are an infinite number of velocity profiles
60
Outlet with a non-uniform velocity
What is the net force exerting on each blade? What is the power trans-
mitted to the blade?
61
See this solution worked out step by step on YouTube
https://youtu.be/BgUjpaBYeDc (CC-by Olivier Cleynen)
𝛿
(3/17)
The plate has length 𝐿1 = 50 cm (in 𝑥-direction, along the flow) and
width 𝐿2 = 80 cm (in 𝑧-direction, across the flow). The density of air is
1,225 kg m−3 .
The layer has thickness 𝛿 at outlet. What is the inlet height of a control
volume that has the mass flow of this layer?
62
Problem sheet 3: Analysis of existing flows
with three dimensions
last edited May 15, 2020
by Olivier Cleynen — https://fluidmech.ninja/
The atmosphere has 𝑝atm. = 1 bar; 𝜌atm. = 1,225 kg m−3 ; 𝑇atm. = 11,3 °C; 𝜇atm. = 1,5 ⋅ 10−5 Pa s
Except otherwise indicated, assume that:
Air behaves as a perfect gas: 𝑅air =287 J kg−1 K−1 ; 𝛾air =1,4; 𝑐𝑝 air =1 005 J kg−1 K−1 ; 𝑐𝑣 air =718 J kg−1 K−1
Liquid water is incompressible: 𝜌water = 1 000 kg m−3 , 𝑐𝑝 water = 4 180 J kg−1 K−1
ter enters and leaves the pipe with the same speed 𝑉1 = 𝑉2 = 1,5 m s−1 . The velocity
distribution at both inlet and outlet is uniform.
3.2.2. What is the force exerted by the pipe bend on the water?
3.2.3. Represent the force vector qualitatively (i.e. without numerical data).
3.2.4. What would be the new force if all of the speeds were doubled?
63
Figure 3.6: A pipe bend, through which water is flowing. We assume that the velocity distribution
at inlet and outlet is identical.
Figure CC-0 Olivier Cleynen
Figure 3.7: A mobile exhaust gas deflector, used to deflect hot jet engine exhaust gases upwards
during ground tests.
Figure CC-0 Olivier Cleynen
speed is 𝑉jet = 600 km h−1 , temperature 400 °C and the pressure is atmospheric. As the
The deflector is fed with a horizontal air jet with a quasi-uniform velocity profile; the
3.3.1. What is the force exerted on the ground by the deflection of the exhaust gases?
3.3.2. Describe qualitatively (i.e. without numerical data) a modification to the deflec-
tor that would reduce the horizontal component of force.
3.3.3. What would the force be if the deflector traveled rearwards (positive 𝑥-direction)
with a velocity of 10 m s−1 ?
can be neglected, and the water jet is deflected entirely with a 180° angle.
nozzle hits a blade which is mounted on a rotor (fig. 3.8). In the ideal case, viscous effects
The nozzle has a cross-section diameter of 5 cm and produces a water jet with a speed
𝑉jet = 15 m s−1 . The rotor diameter is 2 m and the blade height is negligibly small.
We now let the rotor rotate freely. Friction losses are negligible, and it accelerates until it
reaches maximum velocity.
3.4.7. What is the maximum power that can be transmitted to the generator?
3.4.8. How would the above result change if viscous effects were taken into account?
(briefly justify your answer, e.g. in 30 words or less)
3.5 Snow plow derived from Gerhart & Gross [7] Ex5.9
A road-based snow plow (fig. 3.9) is clearing up the snow on a flat surface. We wish to
The snow plow is advancing at 25 km h−1 ; its blade has a frontal-view width of 4 m.
quantify the power required for its operation.
The snow on the ground is 30 cm deep and has density 300 kg m−3 .
The snow is pushed along the blade and is rejected horizontally with a 30° angle to the
left of the plow. Its density has then risen to 450 kg m−3 . The cross-section area 𝐴outlet of
the outflowing snow in the 𝑥-𝑦 plane is 1,1 m2 .
3.5.1. What is the force exerted on the blade by the deflection of the snow?
(Indicate its magnitude and coordinates)
3.5.2. What is the power required for the operation of the snow plow?
3.5.3. If the plow velocity was increased by 10 %, what would be the increase in power?
65
Figure 3.9: Outline schematic of a blade snow plow.
Figure CC-0 Olivier Cleynen
𝑟 7
relationship:
𝑢2(𝑟) = 𝑈center (1 − )
1
𝑅
(3/18)
66
Figure 3.10: Velocity profiles at the inlet and outlet of a circular pipe.
Figure CC-0 Olivier Cleynen
Figure 3.11: A cylinder profile set up in a wind tunnel, with the air flowing from left to right.
Figure CC-0 Olivier Cleynen
Upstream of the cylinder, the air flow velocity is uniform (𝑢1 = 𝑈 = 30 m s−1 ).
Downstream of the cylinder, the speed is measured across a 2 m height interval. Hori-
zontal speed measurements are gathered and modeled with the following relationship:
𝑢2(𝑦) = 29 + 𝑦 2 (3/19)
The width of the cylinder (perpendicular to the flow) is 2 m. The Mach number is very
low, and the air density remains constant at 𝜌 = 1,23 kg m−3 ; pressure is uniform all along
the measurement field.
and the function 𝑢2(𝑦) above only modeled time-averaged values of the horizontal
3.7.2. How would this value change if the flow in the cylinder wake was turbulent,
In order to achieve this, measurements of the horizontal velocity 𝑢 are made around the
We wish to measure the drag applying on a thin plate positioned parallel to an air stream.
At the leading edge of the plate, the horizontal velocity of the air is uniform: 𝑢1 = 𝑈 =
Figure CC-0 Olivier Cleynen
10 m s−1 .
by the effect of shear. This layer, called boundary layer, has a thickness of 𝛿 = 1 cm. The
At the trailing edge of the plate, we observe that a thin layer of air has been slowed down
𝑦 7
𝑢2(𝑦) =𝑈( )
1
𝛿
(3/20)
The width of the plate (perpendicular to the flow) is 30 cm and it has negligible thickness.
The flow is incompressible (𝜌 = 1,23 kg m−3 ) and the pressure is uniform.
3.8.3. Under which form is the kinetic energy lost by the flow carried away? Can this
new form of energy be measured? (briefly justify your answer, e.g. in 30 words
or less)
68
3.9 Drag measurements in a wind tunnel
A group of students proceeds with speed measurements in a wind tunnel. The objective
is to measure the drag applying on a wing profile positioned across the tunnel test section
(fig. 3.13).
Figure 3.13: Wing profile positioned across a wind tunnel. The horizontal velocity distributions
upstream and downstream of the profile are also shown.
Figure CC-0 Olivier Cleynen
Upstream of the profile, the air flow velocity is uniform (𝑢1 = 𝑈 = 50 m s−1 ).
Downstream of the profile, horizontal velocity measurements are made every 5 cm across
the flow; the following results are obtained:
5 50
10 49
15 48
20 45
25 41
30 39
35 40
40 43
45 47
50 48
55 50
60 50
The width of the profile (perpendicular to the flow) is 50 cm. The airflow is incompressible
(𝜌 = 1,23 kg m−3 ) and the pressure is uniform across the measurement surface.
3.9.2. How would the above calculation change if vertical speed measurements were
also taken into account?
69
3.10 Moment on gas deflector non-examniable
deflector viewed from the side. The midpoint of the inlet is 2 m above and 5 m behind
We revisit the exhaust gas deflector of exercise 3.3 p. 64. Figure 3.14 below shows the
the wheel labeled “A”, while the center axis of the outlet passes 1,72 m away from it, as
represented in fig 3.14
Figure 3.14: Side view of the mobile exhaust gas deflector which was shown in fig. 3.7 p. 64
Figure CC-0 Olivier Cleynen
What is the moment generated by the gas flow about the axis of the wheel labeled “A”?
• inlet A has a cross-section area of 0,2 m2 . It contributes hot exhaust gases of density
0,8 kg m−3 and velocity 12 m s−1 , aligned with the (𝑥) axis of the tail;
with a fixed velocity of 45 m s−1 . The angle 𝜃 at which gases are rejected is controlled by
The mix of exhaust gases and atmospheric air is rejected at the tip of the tail (outlet C)
3.11.1. What is the rejection angle 𝜃 required so that the tail generates a moment
of +6 kN m around the main rotor (𝑦) axis?
70
Figure 3.15: Top-view of a helicopter using a rotor-less tail.
Figure derived from a figure CC-by-sa by Commons User:FOX 52
3.11.2. Propose and quantify a modification to the tail geometry or operating conditions
71
Answers
1) 𝑚̇ = 1,0603 kg s−1
2) 𝐹⃗net = (−0,5681; −1,2184) N
3.2
1) 𝐹⃗net = (−9,532; +1,479) kN : ||𝐹⃗net || = 9,646 kN (force on ground is opposite: 𝐹⃗fluid on pipe =
−𝐹⃗pipe on fluid );
3.3
1) 𝐹net = (−883,6; 0) N;
2) 𝑀net 𝑋 = |𝐹net |𝑅 = 883,6 N m;
3.4
3) 𝑊̇ rotor = 0 W;
4) 𝜔 = 143,2 rpm (𝐹net = 0 N);
5) 𝑊̇ rotor = 0 W again;
6) 𝑊̇ rotor, max = 1,963 kW @ 𝑉blade, optimal = 13 𝑉water jet .
1) 𝑚̇ = 2 500 kg s−1 ; 𝐹net 𝑥 = +10,07 kN, 𝐹net 𝑧 = −12,63 kN (force on blade is opposite);
2) 𝑊̇ = 𝐹⃗net ⋅ 𝑉⃗plow = 𝐹net𝑥 |𝑉1 | = 69,94 kW
3.5
3) 𝑊̇ 2 = 1,13 𝑊̇ (+33 %)
3.10 Re-use 𝑚̇ = 67,76 kg s−1 , 𝑉1 = 166,7 m s−1 , 𝑉2 = 33,94 m s−1 from ex. 3.3. With 𝑅2⟂𝑉2 =
1,717 m, plug in numbers in eq. 3/14 p. 58: 𝑀net = +18,64 kN m in 𝑧-direction.
3.11 1) Work eq. 3/13 down to scalar equation (in 𝑦-direction), solve for 𝜃: 𝜃 = 123,1°.
be solved at the same time. 𝑟C can be shortened, the flow in C can be split into
2) There are multiple solutions which allow both moment and force equations to
72
Fluid Dynamics
Chapter 4 – Effects of pressure
last edited September 3, 2020
by Olivier Cleynen — https://fluidmech.ninja/
4.1 Motivation 73
4.2 Pressure forces on walls 73
4.2.1 Magnitude of the pressure force 73
4.2.2 Position of the pressure force 74
4.3 Pressure fields in fluids 75
4.3.1 The direction of pressure 75
4.3.2 Pressure on an infinitesimal volume 76
4.4 Special case: pressure in static fluids 79
4.4.1 Forces in static fluids 79
4.4.2 Pressure and depth 80
4.4.3 Buoyancy 82
4.5 Solved problems 83
4.6 Problems 85
These notes are based on textbooks by White [22], Çengel & al.[25], Munson & al.[29], and de Nevers [17].
4.1 Motivation
In fluid mechanics, only three types of forces apply to fluid particles: forces
due to gravity, pressure, and shear. This chapter focuses on pressure (we will
address shear in chapter 5), and should allow us to answer two questions:
When the pressure 𝑝 exerted is uniform and the wall is flat, the resulting
What is the force with which a fluid pushes against a wall?
When the fluid pressure 𝑝 is not uniform (for example, as depicted on the
hours of heavy tonnage transit
through the Miraflores locks in
Panama. Can you quantify the
force applying on a single lock
right side of the wall in figure 4.1), the situation is more complex: the force door?
The above equations work only for a flat surface. When we consider a two-
or three-dimensional object immersed in a fluid with non-uniform pressure,
the integration must be carried out with vectors.
where 𝑟⃗X𝐹 is a vector expressing the position of each infinitesimal surface relative to
point X.
Much like eq. 4/4 above, this eq. 4/5 is easily implemented in a software
algorithm but not very approachable on paper. In this course however,
we want to study the simple case where the surface is flat, and where the
reference point X is in the same plane as the surface. Equation 4/5 is then a
74
on an arbitrary surface (left: perspective view; right: side view). The vector 𝑛⃗ is a
Figure 4.2: Moment generated about an arbitrary point X by the pressure exerted
Once both 𝐹pressure and 𝑀X pressure have been quantified, the distance 𝑅X𝐹 be-
tween point X and the application point of the net pressure force is easily
computed:
𝑀X pressure
𝑅X𝐹 =
𝐹pressure
(4/7)
𝐹⟂
𝑝≡
𝐴
(4/8)
Equation 4/9 may appear unsettling at first sight, because as 𝐴 tends to zero,
𝐹⟂ also tends to zero; nevertheless, in any continuous medium, the ratio of
these two terms tends to a single non-zero value: the local pressure.
This brings us to the second particularity of pressure in fluids: the pressure
on either side of the infinitesimal flat surface is the same regardless of its
Video: Two weird things about orientation. In other words, pressure has no direction: there is only one (scalar)
pressure in fluid dynamics
by Olivier Cleynen (CC-by)
https://youtu.be/0d8gfsKllmU
value for pressure at any one point in space.
Thus, in a fluid, pressure applies not merely on the solid surfaces of its
What are those three components? In the 𝑥-direction, the pressure on faces 1
will therefore have three components: one for each pair of opposing faces.
Figure 4.3: The pressure on each face of an infinitesimal volume may have a different
These changes are labeled d𝑝|𝑖 in each of the 𝑖 = 𝑥, 𝑦, 𝑧 directions.
value. The net effect of pressure will depend on how the pressure varies in space.
76
We express 𝑝4 − 𝑝1 as the derivative of pressure in the 𝑥-direction (𝜕𝑝/𝜕𝑥),
multiplied by the distance d𝑥 which separates points 1 and 4, obtaining:
𝜕𝑝
𝐹net, pressure,𝑥 = d𝑦 d𝑧 − d𝑥
[ 𝜕𝑥 ]
−𝜕𝑝
= d
𝜕𝑥
(4/10)
Now generalizing eq. 4/10 for the other two directions, we can write:
−𝜕𝑝
𝐹net, pressure,𝑥 = d
𝜕𝑥
−𝜕𝑝
𝐹net, pressure,𝑦 = d
𝜕𝑦
−𝜕𝑝
𝐹net, pressure,𝑧 = d
𝜕𝑧
𝐹⃗net, pressure = − d ∇
⃗𝑝 (4/12)
Finally, we obtain:
1 ⃗
⃗𝑝
𝐹net, pressure = −∇
d
(4/13)
This last equation reads “the pressure force per unit volume is the opposite
of the pressure gradient”. It shows us that in any fluid and any situation, the
force due to pressure points the opposite way of the pressure gradient. Thus,
if a particle of any kind is “dropped” into a fluid flow, we can quantify in
which direction, and with which magnitude, pressure (a scalar field) is going
to “push” it. This is given by equation 4/13, which quantifies this effect as a
vector field (see figures 4.4 & 4.5).
for each point in space, once for each time point. This field of vectors
77
would indicate the amount of force per unit volume with which pressure
is pushing the fluid.
Having a good grasp of the tools we are using here is important, because
things will soon get more complicated when shear is added to the equa-
tion (in chapter 5), ultimately leading to the all-powerful Navier-Stokes
equation in chapter 6.
Figure 4.4: Water flow in a two-dimensional water tank, visualized with a computa-
enters with a velocity of 10 m s−1 in a 10 m-high tunnel, and flows around a “bump” at
tional fluid dynamics (cfd) software package. The flow is from left to right: water
the bottom. On the top, the magnitude of velocity is represented (background color),
with white lines indicating flow direction. On the bottom, pressure is displayed. The
minimum pressure in the flow is zero Pascal.
values for pressure have been arbitrarily adjusted for visual purposes so that the
Fluid statics is the study of fluids at rest, i.e. those whose velocity field 𝑉⃗ is
{
everywhere null and constant:
𝑉⃗ = 0⃗
𝜕 𝑉⃗
= 0⃗
(4/14)
𝜕𝑡
We choose to study this type of problem now, because it makes for a concep-
tually and mathematically simple case with which we can practice calculating
fluid-induced forces.
What are the forces applying on an arbitrary particle in a static fluid?
• The force due to shear is zero. We will indeed see in chapter 5 (Effects of
shear) that shear efforts can be expressed as a function of viscosity and
velocity. All ordinary fluids are unable to exert shear when they are
static.
In a moving fluid, the sum of these forces would add up to the mass of the
particle times its acceleration. But in a static fluid, the velocity is zero and
never changes. We can thus write:
⃗ 𝑝 = 𝜌 𝑔⃗
∇ (4/15)
in a static fluid.
This is a very useful equation, which states that in a static fluid, the only
parameter affecting pressure is gravity. More precisely, the fluid density
times the gravity vector is equal to the change in space of the pressure.
We will see in chapter 6 (Prediction of fluid flows) that equation 4/15 is the specific
case for a much larger general and powerful equation, the Navier-Stokes
equation. But more on that later!
Very often in studies of static fluids, the 𝑧-axis is oriented vertically, positive
It is now easy to quantify pressure everywhere inside a static fluid.
The first consequence we draw from equation 4/16 is that in a static fluid
(e.g. in a glass of water, in a swimming pool, in a calm atmosphere), pressure
depends solely on height. Within a static fluid, at a certain altitude, we will
measure the same pressure regardless of the surroundings (fig. 4.6).
d𝑝
= 𝜌water 𝑔
( d𝑧 )water
d𝑝
= 1 000 × 9,81 = 9,81 ⋅ 103 Pa m−1 = 9,81 ⋅ 10−2 bar m−1
( d𝑧 )water
(4/17)
80
the environment. Here, as long as the fluid remains static, 𝑝A = 𝑝B = 𝑝C = 𝑝D .
Figure 4.6: Pressure at a given depth (or height) in a static fluid does not depend on
gradient as:
d𝑝 1𝑔
= 𝜌air 𝑔 = 𝑝
( d𝑧 )atm. 𝑇𝑅
(4/18)
d𝑝 1 9, 81
= 1 ⋅ 105 × ×
( d𝑧 )atm. ambient 288,15 287
= 11,86 Pa m−1 = 1,186 ⋅ 10−4 bar m−1 (4/19)
This rate (approximately 0,1 mbar/m) is almost a thousand times smaller than
that of water (fig. 4.7).
Since the rate of pressure change depends on pressure, it also varies with
altitude, and the calculation of pressure differences in the atmosphere is a
little more complicated than for water.
2
1 𝑔 2
∫ 𝑝 d𝑝 = 𝑅𝑇 ∫ d𝑧
by Randall Munroe (CC-by-nc)
https://xkcd.com/2153
𝑝2 𝑔
1 cst. 1
ln = Δ𝑧
𝑝1 𝑅𝑇cst.
𝑝2 𝑔Δ𝑧
= exp
𝑝1 [ 𝑅𝑇cst. ]
(4/20)
81
Figure 4.7: Variation of pressure as a function of altitude for water and air at the
surface of a water reservoir. The gradient of pressure with respect to altitude is
almost a thousand times larger in water than in air.
Figure CC-0 Olivier Cleynen
4.4.3 Buoyancy
Any solid body immersed within a fluid is subjected to pressure on its walls.
When the pressure is not uniform (for example because the fluid is subjected
to gravity, although this may not be the only cause), then the net force due
to fluid pressure on the body walls will be non-zero.
Video: playing around with an
air pump and a vacuum chamber
When the fluid is purely static, this net pressure force is called buoyancy.
by Y:Roobert33 (styl) Since in this case, the only cause for the pressure gradient is gravity, the
https://youtu.be/ViuQKqUQ1U8
net pressure force is oriented upwards. The buoyancy force is completely
independent from (and may or may not compensate) the object’s weight.
Since it comes from equation 4/15 that the variation of pressure within a
fluid is caused solely by the fluid’s weight, we can see that the force exerted
on an immersed body is equal to the weight of the fluid it replaces (that is to
82 say, the weight of the fluid that would occupy its own volume were it not
there). This relationship is sometimes named Archimedes’ principle. The
force which results from the static pressure gradient applies to all immersed
bodies: a submarine in an ocean, an object in a pressurized container, and
of course, the reader of this document as presently immersed in the earth’s
atmosphere.
Figure 4.8: Immersion in a static fluid results in forces that depend on the body’s vol-
ume. They can evidenced by the removal of the fluid (for example in a depressurized
semi-spherical vessel).
Figure CC-0 Olivier Cleynen
A lake has the dimensions shown in the figure above. What is the pressure
at the bottom of the lake?
The side wall of a water tank has the dimensions shown in the figure
above. What is the force exerted due to the pressure of the water on the
wall?
84
Problem sheet 4: Effects of pressure
last edited April 25, 2021
by Olivier Cleynen — https://fluidmech.ninja/
The atmosphere has 𝑝atm. = 1 bar; 𝜌atm. = 1,225 kg m−3 ; 𝑇atm. = 11,3 °C; 𝜇atm. = 1,5 ⋅ 10−5 Pa s
Except otherwise indicated, assume that:
Air behaves as a perfect gas: 𝑅air =287 J kg−1 K−1 ; 𝛾air =1,4; 𝑐𝑝 air =1 005 J kg−1 K−1 ; 𝑐𝑣 air =718 J kg−1 K−1
Liquid water is incompressible: 𝜌water = 1 000 kg m−3 , 𝑐𝑝 water = 4 180 J kg−1 K−1
pressure 𝑝atm. .
Figure 4.10: Working principle of a simple liquid tube manometer. The outlet is at atmospheric
Figure CC-0 Olivier Cleynen
window is 3 m high, 4 m wide, and is positioned 0,4 m above the bottom of the tank.
A water tank has a window on one of its straight walls, as shown in figure 4.11. The
Figure 4.11: An aquarium tank with a window installed on one of its walls
Figure CC-0 Olivier Cleynen
4.4.1. What is the magnitude of the force applying on the window due to the pressure
exerted by the water?
4.4.2. At which height does this force apply?
The hinge stands 1,5 m below the water surface. The window has a length of 0,9 m and
and measurements. The window is hinged on its bottom face.
a width of 2 m. The walls of the channel are inclined with an angle 𝜃 = 60° relative to
horizontal.
Figure 4.12: A door installed on the wall of a water channel. The water in the canal is perfectly
still.
Figure CC-0 Olivier Cleynen
4.5.1. Represent graphically the pressure of the water and atmosphere on each side of
the window.
4.5.2. What is the magnitude of the moment exerted by the pressure of the water
about the axis of the window hinge?
4.5.3. If the same door was positioned at the same depth, but the angle 𝜃 was decreased,
would the moment be modified? (briefly justify your answer, e.g. in 30 words or
86 less)
4.6 Pressure force on a cylinder
An idealized flow over a cylinder is depicted in figure 4.13. This is a very primitive
flow solution, obtained using a model (the potential flow model, which we mention in
chapter 11) which cannot account for viscous effects or flow separation. Nevertheless, it
provides a good first “ideal flow” situation to compute surface pressure forces in fluid
flows.
Figure 4.13: Idealized flow around a cylinder, as predicted by potential flow theory. The stream-
along the 𝑧 direction.
lines are represented only in a plane crossing the center of the cylinder, but they are identical all
The cylinder has diameter 10 cm, and it spans 50 cm across the flow (in the 𝑧 direction).
𝑉 = 0 m s−1 and the air pressure is everywhere 𝑝∞ = 1 bar. We would like to calculate the
We start by considering the case where there is no flow: the velocity is everywhere
forces applying represented in figure 4.14, both caused by the air pressure.
87
(a couple of hints to help with the algebra: ∫ sin 𝑥 d𝑥 = − cos 𝑥 + 𝑘 and ∫ sin3 𝑥 d𝑥 =
1
3
cos3 𝑥 − cos 𝑥 + 𝑘)
We now consider the case there there is fluid flow: air with density 𝜌atm. = 1,225 kg m−3
is coming in at 𝑉∞ = 50 km h−1 . In this case, the pressure 𝑝𝑠 on the surface of the cylinder
coordinate 𝜃 as:
is no longer uniform (see also problem 11.3 p. 236). It is expressed as a function of the
1
𝑝𝑠 = 𝑝∞ + 𝜌 (𝑉∞2 − 4𝑉∞2 sin2 𝜃 )
2
(4/21)
(a couple of hints to help with the algebra: ∫ cos 𝑥 d𝑥 = sin 𝑥 + 𝑘 and ∫ cos 𝑥 sin2 𝑥 d𝑥 =
1
3
sin3 𝑥 + 𝑘)
4.7.1. Sketch the distribution of pressure on each of the immersed walls of the barge
(left and right sides, rear , bottom and slanted front).
4.7.2. What is the magnitude of the force resulting from pressure efforts on each of
these walls?
and constant (𝑇 = 𝑇cst. ). In practice, this may not always be the case.
tion in the atmosphere was based on the hypothesis that the temperature was uniform
88
Khalifa tower (800 m above the ground). Inside the tower, the temperature is controlled
A successful fluid dynamics lecturer purchases an apartment at the top of the Burj
everywhere at 18,5 °C. Outside, the ground temperature is 30 °C and it decreases linearly
with altitude (gradient: −7 K km−1 ).
For the purpose of the exercise, we pretend the tower is entirely hermetic (meaning air is
prevented from flowing in or out of its windows).
4.8.2. What is the pressure difference between each side of the windows in the apart-
ment at the top of the tower?
89
Answers
4.2 𝑝A = 𝑝atm. + 0,039 bar ≈ 1,039 bar.
U-tube star?
2) 𝑀net = 7,79 kN m;
3) observe the equation used to calculate 𝑀net to answer this question. If needed,
4.5
2) 𝐹rear = 0,2453 MN, 𝐹side = 2,1714 MN, 𝐹bottom = 13,734 MN, 𝐹front = 0,3468 MN;
3) 𝐹buoyancy = 13,979 MN (1 425 t).
4.7
1) 𝑝𝑝21 = (1 + 𝑘𝑧
𝑇1 )
𝑔
2 𝑘𝑅
2) Δ𝑝 = +247,4 Pa inwards.
4.8
90
Fluid Dynamics
Chapter 5 – Effects of shear
last edited September 19, 2020
by Olivier Cleynen — https://fluidmech.ninja/
5.1 Motivation 91
5.2 Shear forces on walls 91
5.2.1 Magnitude of the shear force 91
5.2.2 Direction and position of the shear force 92
5.3 Shear fields in fluids 92
5.3.1 The direction of shear 93
5.3.2 Shear on an infinitesimal volume 93
5.4 Resistance to shear: viscosity 97
5.4.1 Viscosity 97
5.4.2 Kinematic viscosity 98
5.4.3 Turbulent viscosity 99
5.4.4 Non-Newtonian fluids 100
5.4.5 The no-slip condition 100
5.5 Special case: shear in simple laminar flows 101
5.6 Solved problems 103
5.7 Problems 105
These notes are based on textbooks by White [22], Çengel & al.[25], Munson & al.[29], and de Nevers [17].
5.1 Motivation
In fluid mechanics, only three types of forces apply to fluid particles: forces
due to gravity, pressure, and shear. This chapter focuses on shear, and should
allow us to answer two questions:
When the shear 𝜏 exerted is uniform and the wall is flat, the resulting force 𝐹
What is the force which which a fluid shears (i.e. “rubs”) against a wall?
When the shear 𝜏 exerted by the fluid is not uniform (for example, because
more friction is occurring on some parts of the surface than on others), the
situation is more complex: the force must be obtained by integration. The
91
surface is split in infinitesimal portions of area d𝑆, and the corresponding
forces are summed up as:
Much like equation 4/4 in the previous chapter, eq. 5/4 is not too hard to
implement as a software algorithm to obtain numerically, for example, the
force resulting from shear due to fluid flow around a body such as the body
of a car. Its computation by hand, however, is far too tedious for us to even
attempt.
The position of the shear force is obtained with two moment vector equations,
in a manner similar to that described in §4.2.2 p. 74 with pressure. This is
outside of the scope of this course.
𝐹∥
𝜏≡
𝐴
(5/5)
Like we did with pressure, to appreciate the concept of shear in fluid mechan-
ics, we need to go beyond this equation.
92
5.3.1 The direction of shear
Already from the definition in eq. 5/5 we can appreciate that “parallel to a
flat plate” can mean a multitude of different directions, and so that we need
more than one dimension to represent shear. Furthermore, much in the same
way as we did for pressure, we do away with the flat plate and accept that
shear is a field, i.e. it is an effort applying not only upon solid objects but also Video: cloud movements in a
time-lapse video on an interest-
upon and within fluids themselves. We replace eq. 5/5 with a more general ing day are evidence of a highly-
definition:
𝐹⃗∥
strained atmosphere: pilots and
meteorologists refer to this as
𝜏⃗ ≡ lim
wind shear.
𝐴→0 𝐴
by Y:StormsFishingNMore (styl)
(5/6) https://youtu.be/LjWeYPEmCk8
Contrary to pressure, shear is not a scalar, i.e. it can (and often does) take
⎛ 𝜏𝑥 ⎞
𝜏⃗(𝑥,𝑦,𝑧,𝑡) ≡ ⎜ 𝜏𝑦 ⎟
⎜ ⎟
⎝ 𝜏𝑧 ⎠(𝑥,𝑦,𝑧,𝑡)
(5/7)
faces 1 to 3 represented). The shear tensor 𝜏⃗𝑖𝑗 has six members of three components
Figure 5.1: Shear efforts on a cubic fluid particle (with only the efforts on the visible
each.
Figure CC-0 Olivier Cleynen
93
For example, 𝜏⃗𝑥𝑦 represents the shear in the 𝑦-direction on a surface perpen-
dicular to the 𝑥-direction. On this face, the shear vector would be:
In eq. 5/8, a surprising term appears: 𝜏𝑥𝑥 . It is the shear effort perpendicular
to the surface of interest. How is this possible? The answer is that the faces
of the infinitesimal cube studied here are not solid. They are permeable, and
the local velocity may include have a component through the face of the cube
Video: what is 𝜏𝑥𝑥 , perpendicu- (in fact, this must happen for any flow to occur at all). Therefore, there is no
lar shear? reason for the shear effort, which is three-dimensional, to be aligned along
by Olivier Cleynen (CC-by)
https://youtu.be/_3EYEySeG4g each flat surface. As the fluid travels across any face, it can be sheared and
strained in any arbitrary direction, and therefore, shear can and most often
does have a component (𝜏𝑖𝑖 ) perpendicular to an arbitrary surface inside a
fluid.
Now, the net shear effect on the cube will have eighteen components: one
tree-dimensional vector for each of the six faces. Each of those components
en entity —a tensor— containing six vectors 𝜏⃗1 , 𝜏⃗2 , 𝜏⃗3 . . . 𝜏⃗6 . By convention,
may take a different value. The net shear could perhaps be represented as
however, shear is notated using only three vector components: one for each
field 𝜏⃗𝑖𝑗 :
pair of faces. Shear efforts on a volume are thus represented with a tensor
In this last equation 5/10, each of the nine components of the tensor acts as the
container for two contributions: one for each of the two faces perpendicular
to the direction expressed in its first subscript.
So much for the shear effort on an element of fluid. What about the net force
due to shear on the fluid element? Not every element counts: part of the
shear will accelerate (change the velocity vector) the particle, while part of it
start with the 𝑥-direction, which consists of the sum of the component of
shear in the 𝑥-direction on each of the six cube faces:
94
Given that 𝑆3 = 𝑆6 = d𝑥 d𝑦, that 𝑆2 = 𝑆5 = d𝑥 d𝑧 and that 𝑆1 = 𝑆4 = d𝑧 d𝑦, this
is re-written as:
𝐹⃗shear 𝑥 = d𝑥 d𝑦 (⃗
𝜏𝑧𝑥 3 − 𝜏⃗𝑧𝑥 6 )
𝜏𝑦𝑥 2 − 𝜏⃗𝑦𝑥 5 )
+ d𝑥 d𝑧 (⃗
𝜏𝑥𝑥 1 − 𝜏⃗𝑥𝑥 4 )
+ d𝑧 d𝑦 (⃗ (5/12)
In the same way we did with pressure in chapter 4 (§4.4.2 p. 80), we express
each pair of values as a derivative with respect to space multiplied by an
infinitesimal distance:
𝜕 𝜏⃗𝑧𝑥 𝜕 𝜏⃗𝑦𝑥 𝜕 𝜏⃗𝑥𝑥
𝐹⃗shear 𝑥 = d𝑥 d𝑦 d𝑧 + d𝑥 d𝑧 d𝑦 + d𝑧 d𝑦 d𝑥
( 𝜕𝑧 ) ( 𝜕𝑦 ) ( 𝜕𝑥 )
𝜕 𝜏⃗𝑧𝑥 𝜕 𝜏⃗𝑦𝑥 𝜕 𝜏⃗𝑥𝑥
= d + +
( 𝜕𝑧 𝜕𝑦 𝜕𝑥 )
(5/13)
We can see with this equation 5/13 that shear in the 𝑥-direction has three
⃗⋅ :
written ∇
Now, we introduce the operator divergent (see also Appendix A3 p. 250),
𝜕 ⃗ 𝜕 ⃗ 𝜕 ⃗
Video: the divergent of shear
⃗⋅ ≡
∇ 𝑖⋅ + 𝑗 ⋅ + 𝑘⋅
by Olivier Cleynen (CC-by)
𝜕𝑥 𝜕𝑦 𝜕𝑧
https://youtu.be/PvcT_E55Lgc
(5/14)
𝜕𝐴𝑥 𝜕𝐴𝑦 𝜕𝐴𝑧
⃗ ⋅ 𝐴⃗ ≡
∇ + +
𝜕𝑥 𝜕𝑦 𝜕𝑧
(5/15)
⎛ 𝜕𝑥 + 𝜕𝑦 + 𝜕𝑧 ⎞ ⎛
𝜕𝐴 𝜕𝐴𝑦𝑥 𝜕𝐴
⃗ ⋅ 𝐴⃗𝑖𝑥
∇ ⎞
⎜ ⎟ ⎜ ⎟
𝑥𝑥 𝑧𝑥
force in the 𝑥-direction is equal to the particle volume times the divergent of
With this new tool, we can go back to equation 5/13 to see that the net shear
𝐹⃗shear 𝑥 = d ∇
⃗ ⋅ 𝜏⃗𝑖𝑥 (5/17)
So much for the 𝑥-direction. The 𝑦- and 𝑧-direction are taken care of in the
same fashion, so that we can gather up our puzzle pieces and express the
force per unit volume due to shear as the divergent of the shear tensor:
⎛ 𝐹shear 𝑥 ⎞ ⃗ ⋅ 𝜏⃗𝑖𝑥 |
⎛ |∇ ⎞
𝐹shear = ⎜ 𝐹shear 𝑦
⃗ ⎟ = d ⎜ |∇
⃗ ⋅ 𝜏⃗𝑖𝑦 | ⎟ = d ∇
⃗ ⋅ 𝜏⃗𝑖𝑗
⎜ ⎟ ⎜ ⎟
⎝ 𝐹shear 𝑧 ⎠ ⃗ ⋅ 𝜏⃗𝑖𝑧 |
⎝ |∇ ⎠
(5/18)
1 ⃗
⃗ ⋅ 𝜏⃗𝑖𝑗
𝐹net, shear = ∇
d
(5/19)
An example of a divergent of shear vector field is shown in figures 5.2 and 5.3.
This equation 5/19 is more than we really need to go through the problems
in this chapter, but we will come back to it when we will want to calculate 95
in figure 4.4 p. 78. Vectors with magnitude lower than 1 Pa m−1 are not represented.
Figure 5.2: The divergent of shear in the flow field of the computed flow described
The arrows indicate the local force per unit volume with which shear is acting on
the fluid.
Figure CC-by by Arjun Neyyathala
Figure 5.3: The magnitude of the divergent of shear in the flow field of the computed
figure 5.2 above). In the top image, the color scale is saturated at 1 ⋅ 105 Pa m−1 , while
flow described in figure 4.4 p. 78 (the magnitude of the vector field represented in
on the bottom image, it reaches a value 20 times higher (2 ⋅ 106 Pa m−1 ), showing the
very high local values attained very close to the wall.
Figures CC-by by Arjun Neyyathala
the dynamics of fluid particles in chapter 6 (Prediction of fluid flows), where the
divergent of shear be a building block of the glorious Navier-Stokes equations.
For now, it is enough to sum up our findings as follows:
96
• The net force due to shear on a volume of fluid, expressed using the
divergent of the shear tensor, has three components — it is a vector
field.
If you were to calculate the effect of shear, then you would calculate its
divergent, which is also a three-dimensional vector field. To record this
⃗ ⋅ 𝜏⃗𝑖𝑥 |, |∇
of |∇ ⃗ ⋅ 𝜏⃗𝑖𝑦 |, and |∇
⃗ ⋅ 𝜏⃗𝑖𝑧 |, all in Pa m−1 ), for each combination of 𝑥,𝑦,𝑧
information, you would, likewise, have to store three values (one for each
and 𝑡. Just like the negative of the gradient of pressure, the divergent
of shear represents a force per unit volume, showing in which direction
shear is “pushing” the fluid particles as they flow.
5.4.1 Viscosity
In chapter 1, we saw already that viscosity 𝜇 is a fluid property that quantifies
its resistance to shear (see §1.4.6 p. 17). More precisely, we quantified 𝜇 as
the ratio of shear stress to strain rate with equation 1/12, reproduced here:
𝜏
𝜇≡
( Δ𝑦 )
Δ𝑣
(5/20)
Now, we generalize this equation: we shrink down the “brick” of fluid from
(fig. 5.4). The strain rate is now 𝜕𝑉𝑗 /𝜕𝑖, which is the rate of change in the
chapter 1 down to an infinitesimal volume of fluid inside an arbitrary flow
If we turn this equation around, we find that we can express the local shear
by differentiating the local velocity with respect to distance:
𝜕𝑉𝑗
||⃗
𝜏𝑖𝑗 || = 𝜇
𝜕𝑖
(5/22)
97
Figure 5.4: Any velocity gradient 𝜕𝑉𝑦 /𝜕𝑥 = 𝜕𝑣/𝜕𝑥 in the flow results in a shear force
𝐹∥ in the direction 𝑦. The ratio between the shear and the velocity gradient is called
viscosity.
Figure CC-0 Olivier Cleynen
1.5×10−2
Viscosity 𝜇 of liquids in Pa s
10−2
5×10−3
Crude Oil
0
Water
Air CO2
Figure 5.5: The viscosity of four fluids (crude oil, water, air, and C02) as a function
of temperature, plotted on a linear scale. This makes clearly visible the difference
in the order of magnitudes of the viscosities of the four fluids, but air and CO2 are
indistinguishably close to the zero axis. A more useful version of this figure is shown
as figure 5.6.
Figure CC-by by Arjun Neyyathala & Olivier Cleynen
98
2×10−2 2.4×10−5
10−2 2.2×10−5
9×10−3
8×10−3
7×10−3 ⟵ Crude Oil
6×10−3 Air ⟶
5×10−3
2×10−5
4×10−3
3×10−3
Viscosity 𝜇 of liquids in Pa s
Viscosity 𝜇 of gases in Pa s
2×10−3 1.8×10−5
CO2 ⟶
10−3 1.6×10−5
9×10−4
8×10−4
7×10−4
6×10−4
5×10−4
1.4×10−5
4×10−4 ⟵ Water
3×10−4
2×10−4 1.2×10−5
10−4 10−5
−20 0 20 40 60 80 100 120
Temperature 𝑇 in degree Celsius (◦C)
Figure 5.6: The viscosity of four fluids (crude oil, water, air, and C02) as a function of
temperature. The scale for liquids is logarithmic and displayed on the left; the scale
for gases is linear and displayed on the right.
Figure CC-by by Arjun Neyyathala & Olivier Cleynen
viscosity 𝜇T (in Pa s just like viscosity). This is because the bulk effect of
In computational fluid dynamics (cfd) simulations, use is made of turbulent
slope of the curve (𝜇) varies with 𝜕𝑉𝑗 /𝜕𝑖 (so, those which do not feature straight lines
Figure 5.7: Various possible viscosity characteristics of fluids. Those for which the
on this diagram) are called non-Newtonian.
100 Figure CC-0 Olivier Cleynen
5.4.5 The no-slip condition
We observe that whenever we measure the velocity of a fluid flow along
a solid wall, the speed tends to zero as we approach the wall surface. In
other words, the fluid adheres to the surface regardless of the overall faraway
flow velocity. This phenomenon, called the no-slip condition, is extremely
important in fluid dynamics. One consequence of this is that fluid flows near
walls are dominated by viscous effects (internal friction) due to the large
strain rates there.
Figure 5.8: A simple flow. The bottom wall is stationary, while the top wall slides
from left to right. In between the walls, fluid is strained uniformly.
Figure CC-by-sa Commons User:Kulmalukko
101
{
A reasonable guess for the velocity distribution in steady laminar regime is:
𝑉𝑥 = 𝑉bottom wall + 𝑘𝑦
𝑉𝑦 = 0
{
can re-write this as:
𝑉𝑥 = 0 + top𝐻wall 𝑦
𝑉
𝑉𝑦 = 0
d 𝑉top wall
𝜏𝑦𝑥 = 𝜇 0+ 𝑦
d𝑦 ( 𝐻 )
𝜇 𝑉top wall
=
𝐻
(it is independent of 𝑦 and 𝑥). A few slightly more complex cases are waiting
Thus, we see here that the shear applied in the fluid is the same everywhere
for us in the problem sheet; but to handle more realistic shear distributions,
what is needed is a software able to compute the behavior of fluids. The basic
but formidable equations to be solved for this are the topic of the upcoming
chapter 6 (Prediction of fluid flows).
102
5.6 Solved problems
Cylindrical viscometer
drawing) is 2 mm.
and the spacing between the two cylinders (greatly exaggerated on the
103
See this solution worked out step by step on YouTube
https://youtu.be/jc9qIv-jzC4 (CC-by Olivier Cleynen)
(𝜇 = 2,27 ⋅ 10−2 Pa s). Many thanks to the students who double-checked and
forgotten in the last few lines. The final numerical result is nevertheless correct
104
Problem sheet 5: Effects of shear
last edited June 5, 2020
by Olivier Cleynen — https://fluidmech.ninja/
The atmosphere has 𝑝atm. = 1 bar; 𝜌atm. = 1,225 kg m−3 ; 𝑇atm. = 11,3 °C; 𝜇atm. = 1,5 ⋅ 10−5 Pa s
Except otherwise indicated, assume that:
Air behaves as a perfect gas: 𝑅air =287 J kg−1 K−1 ; 𝛾air =1,4; 𝑐𝑝 air =1 005 J kg−1 K−1 ; 𝑐𝑣 air =718 J kg−1 K−1
Liquid water is incompressible: 𝜌water = 1 000 kg m−3 , 𝑐𝑝 water = 4 180 J kg−1 K−1
2×10−2 2.4×10−5
10−2 2.2×10−5
9×10−3
8×10−3
7×10−3 ⟵ Crude Oil
6×10−3 Air ⟶
5×10−3
2×10−5
4×10−3
3×10−3
Viscosity 𝜇 of liquids in Pa s
Viscosity 𝜇 of gases in Pa s
2×10−3 1.8×10−5
CO2 ⟶
10−3 1.6×10−5
9×10−4
8×10−4
7×10−4
6×10−4
5×10−4
1.4×10−5
4×10−4 ⟵ Water
3×10 −4
2×10−4 1.2×10−5
10−4 10−5
−20 0 20 40 60 80 100 120
Temperature 𝑇 in degree Celsius ( C)
◦
Figure 5.9: The viscosity of four fluids (crude oil, water, air, and C02) as a function of temperature.
The scale for liquids is logarithmic and displayed on the left; the scale for gases is linear and
displayed on the right. This is a reproduction of figure 5.6 p. 99.
Figure CC-by by Arjun Neyyathala & Olivier Cleynen
105
5.1 Quiz
Once you are done with reading the content of this chapter, you can go take
the associated quiz at https://elearning.ovgu.de/course/view.php?id=7199
In the winter semester, quizzes are not graded.
5.2 Flow in between two plates Munson & al. [29] Ex1.5
flow is laminar (smooth and fully steady), with a univorm velocity profile 𝑢 = 𝑓 (𝑦) which
A fluid is forced to flow between two stationary plates (fig. 5.10). We observe that the
3 𝑦 2
𝑢 = 𝑉average 1 − ( )
2 [ 𝐻 ]
(5/24)
Figure 5.10: Velocity distribution for laminar flow in between two plates, also known as Couette
flow.
Figure CC-0 Olivier Cleynen
The fluid is water at 40 °C, the average velocity is 0,6 m s−1 and the two plates are 1,5 mm
apart.
5.2.1. What is the shear effort 𝜏𝑦𝑥 plate generated on the lower plate?
5.2.2. What is the shear effort 𝜏𝑦𝑥 in the middle plane of the flow?
106
5.3 Friction on a plate
A plate the size of an A4 sheet of paper (210 mm × 297 mm) is moved horizontally at
constant speed above a large flat surface (fig. 5.11). We assume that the velocity profile of
the fluid betweeen the plate and the flat surface is entirely uniform, smooth, and steady.
5.3.1. Express the force 𝐹𝑦𝑧 due to shear on the plate as a function of its velocity 𝑈plate ,
the gap height 𝐻 , and the properties of the fluid.
5.3.2. The plate speed is 𝑈plate = 1 m s−1 and the gap height is 𝐻 = 5 mm. What is the
shear force 𝐹𝑦𝑧 when the fluid is air at 40 °C, and when the fluid is crude oil at
the same temperature?
5.3.3. If a very long and thin plate with the same surface area was used instead of the
A4-shaped plate, would the shear force be different? (briefly justify your answer,
e.g. in 30 words or less)
The two cylinders are 75 cm tall. The inner cylinder diameter is 15 cm and the spacing
within the stationary outer cylinder.
is 1 mm.
When the inner cylinder is rotated at 300 rpm, a friction-generated moment of 0,8 N m is
measured.
5.4.1. If the flow in between the cylinders corresponds to the simplest possible flow
case (steady, uniform, fully-laminar), what is the viscosity of the fluid?
5.4.2. Would a non-Newtonian fluid induce a higher moment? (briefly justify your
answer, e.g. in 30 words or less)
[Note: in practice, when the inner cylinder is turned at high speed, the flow displays
mesmerizing patterns called Taylor—Couette vortices, the description of which is much more
complex!]
107
Figure 5.12: Sketch of a cylinder viscometer. The width of the gap has been greatly exaggerated
for clarity.
Figure CC-0 Olivier Cleynen
layer. The speed 𝑢(𝑦) can then be modeled with the relation:
that viscous effects dominate the mechanics of the flow. This zone is designated boundary
𝜋𝑦
𝑢 = 𝑈 sin ( )
2𝛿
(5/25)
The fluid is CO2 at 20 °C; measurements yield 𝑈 = 10,8 m s−1 and 𝛿 = 3 cm.
5.5.2. At which height 𝑦1 above the surface will the shear effort be half of this value?
5.5.3. What would be the wall shear if the CO2 was replaced with water at the same
temperature?
108
5.6 Clutch Çengel & al. [25] 2-74
Two aligned metal shafts are linked by a clutch, which is made of two disks very close
disk diameters are both 30 cm and the gap between them is 2 mm; they are submerged in
one to another, rotating in the same direction at similar (but not identical) speeds. The
Figure 5.14: Sketch of the two disks constituting the clutch. The gap width has been exaggerated
for clarity.
Figure CC-0 Olivier Cleynen
The power shaft rotates at 1 450 rpm, while the powered shaft rotates at 1 398 rpm. We
consider the simplest possible flow case (steady, laminar) in between the two disks.
5.6.2. How would the moment change if the radius of each disk was doubled?
5.6.4. Briefly (e.g. in 30 words or less) propose one reason why in practice the flow in
between the two disks may be different from the simplest-case flow used in this
exercise.
109
Answers
1) 𝜏𝑦𝑥 |𝑦=−𝐻 = 1,44 N m−2 ;
2) 𝜏𝑦𝑥 |𝑦=0 = 0 N m−2 .
5.2
5.3 1) 𝐹𝑦𝑧 = 𝐿1 𝐿2 𝜇 𝐻𝑈 = 2,38 ⋅ 10−4 N for air, and 6,55 ⋅ 10−2 N for oil.
1) 𝑀 = 𝜋2 𝜇𝜔 𝑅 4 = 7,145 ⋅ 10−3 N m;
3) 𝑊̇ 2 = 𝜔2 𝑀 = 1,05 W; 𝜂clutch = 𝑊
ℎ
̇ 1 = 96,4 % (adequate for this very low-power, low-
5.6
𝑊̇ 2
110
Fluid Dynamics
Chapter 6 – Prediction of fluid flows
last edited September 24, 2021
by Olivier Cleynen — https://fluidmech.ninja/
These notes are based on textbooks by White [22], Çengel & al.[25], Munson & al.[29], and de Nevers [17].
6.1 Motivation
In this chapter we assign ourselves the daunting task of predicting the
movement of fluids completely. We wish to express formally, and calculate,
the dynamics of fluids —the velocity field as a function of time— in any
arbitrary situation. For this, we develop a methodology named derivative
analysis.
Let us start with the unfortunate truth: not only are the methods developed
here are incredibly complex, but they are also very ineffective to solve fluid
flow problems with a pen and paper. Despite this, this chapter is extraordi-
narily important, for two reasons:
111
6.2 Organizing calculations
travel: 𝑉⃗particle = 𝑓 (𝑥0 , 𝑦0 , 𝑧0 , 𝑡0 , 𝑡). This is the process used in solid mechanics
function of time which depends on where and when the particle started its
when we wish to describe the movement of one object, for example, a satellite
in orbit.
This kind of description, however, is poorly suited to the description of fluid
flow, for three reasons:
• Firstly, in order to describe a given fluid flow (e.g. air flow around the
then becomes very difficult to study and describe a problem that is local
in space (e.g. the wake immediately behind the car mirror), because
this requires finding out where the particles of interest originated, and
accounting for the trajectories of each of them.
• Secondly, the concept of a “fluid particle” is not well-suited to the draw-
ing of trajectories. Indeed, not only can particles strain indefinitely,
but they can also diffuse into the surrounding particles, “blurring” and
blending themselves one into another.
• Finally, the velocity of a given particle is very strongly affected by the
properties (velocity, pressure) of the surrounding particles. We have to
resolve simultaneously the movement equations of all of the particles.
𝑓 (𝑥point , 𝑦point , 𝑧point , 𝑡), through which particles of many different origins may
be passing. This is termed a Eulerian flow description, as opposed to the
in our flow study zone, we will obtain a velocity field 𝑉⃗point that is a function
particle-based Lagrangian description.w Grouping all of the point velocities
of time.
speed 𝑢 = 𝑈canal (fig. 6.1). The temperature 𝑇water of the water is constant
Let us imagine a canal in which water is flowing at constant and uniform
112
stationary probe, reading 𝑇probe = 𝑇water on the instrument. Even though the
(in time), but not uniform (in space). We measure this temperature with a
temperature 𝑇water is constant, when reading the value measured at the probe,
{
temperature will be changing with time:
𝜕𝑇water |
𝜕𝑡 |particle
= 0
𝜕𝑇water |
𝜕𝑡 |probe
= − 𝜕𝑇𝜕𝑥
water
𝑢water
Figure 6.1: A one-dimensional water flow, for example in a canal. The water has
a non-uniform temperature, which, even if it is constant in time, translates in a
temperature rate change in time at the probe.
Figure CC-0 Olivier Cleynen
Let us now study the case where the water temperature, in addition to being
non-uniform, is also decreasing everywhere because the canal is cooling
{
𝜕𝑇water |
𝜕𝑡 |particle
≠ 0
Video: figuring out the total
𝜕𝑇water |
𝜕𝑡 |probe
= 𝜕𝑇𝜕𝑡water ||particle − 𝜕𝑇𝜕𝑥 𝑢water
time derivative
by Olivier Cleynen (CC-by)
https://youtu.be/T6POOdLK4ok water
Re-arranging this last equation, we obtain the time change of the particle’s
temperature, expressed from the reference frame of the probe:
𝜕𝑇water | 𝜕𝑇water |
𝜕𝑡 |particle
= 𝜕𝑡 |probe
+ 𝜕𝑇water
𝜕𝑥
𝑢water (6/1)
We must keep in mind that all those derivatives can themselves be functions
of time and space; in equation 6/1, it is their value at the position of the probe
and at the time of measurement which is taken into account.
D 𝜕 𝜕 𝜕 𝜕
Video: half-century-old, but
≡ +𝑢 +𝑣 +𝑤
timeless didactic exploration of
D𝑡 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧
the concept of Lagrangian and
Eulerian derivatives, with ac- (6/2)
D𝐴 𝜕𝐴 𝜕𝐴 𝜕𝐴 𝜕𝐴
companying notes by Lumley[4]
= +𝑢 +𝑣 +𝑤
by the National Committee for Fluid
D𝑡 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧
Mechanics Films (ncfmf, 1969[21]) (styl)
https://youtu.be/mdN8OOkx2ko
(6/3)
⃗
D𝐴 𝜕 𝐴 ⃗ 𝜕𝐴⃗ 𝜕𝐴 ⃗ 𝜕 𝐴⃗
= +𝑢 +𝑣 +𝑤
D𝑡 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧
(6/4)
In equations 6/3 and 6/4, it is possible simplify the notation of the last three
velocity vector 𝑉⃗ (by definition), and the components of the operator gradient
terms. For this, we have to recall two ingredients: the coordinates of the
𝑉⃗ ≡ 𝑖⃗ 𝑢 + 𝑗⃗ 𝑣 + 𝑘⃗ 𝑤
𝜕 ⃗ 𝜕 𝜕
⃗ ≡ 𝑖⃗
∇ +𝑗 + 𝑘⃗
𝜕𝑥 𝜕𝑦 𝜕𝑧
1 Unfortunately this term has many denominations across the literature, including
advective, convective, hydrodynamic, Lagrangian, particle, substantial, substantive, or
114 Stokes derivative. In this document, the term total derivative is used.
We can now define the advective operator,w written (𝑉⃗ ⋅ ∇
⃗ ) (see also Ap-
pendix A3.3 p. 251):
𝜕 𝜕 𝜕
𝑉⃗ ⋅ ∇
⃗≡𝑢 +𝑣 +𝑤
𝜕𝑥 𝜕𝑦 𝜕𝑧
(6/6)
We can now rewrite eqs. 6/2 and 6/3 in a more concise way:
D 𝜕
≡ + (𝑉⃗ ⋅ ∇
⃗)
D𝑡 𝜕𝑡
(6/7)
D𝐴 𝜕𝐴
≡ + (𝑉⃗ ⋅ ∇
⃗ )𝐴
D𝑡 𝜕𝑡
(6/8)
D𝐴⃗ 𝜕 𝐴⃗
≡ + (𝑉⃗ ⋅ ∇
⃗ )𝐴⃗
D𝑡 𝜕𝑡
(6/9)
jet engine nozzle, the air particles accelerate sharply (D/D𝑡 ≠ 0), however
may be steady in the reference frame of the laboratory. For example, in a
the flow is steady from the point of view of the jet engine (𝜕/𝜕𝑡 = 0).
The total time derivative is the tool that we were looking for. From now on,
we can study fluid flows from a stationary reference frame, instead of in the
the same, but all the time derivatives d/ d𝑡 are replaced with total derivatives
reference frame of a moving particle. When we do so, all properties remain
field D𝑉 / D𝑡:
D𝑉⃗ 𝜕 𝑉⃗
= + (𝑉⃗ ⋅ ∇
⃗ )𝑉⃗
D𝑡 𝜕𝑡
(6/10)
115
6.3 Equations for all flows
We can reproduce our analysis from chapter 3 (Analysis of existing flows with
three dimensions) by quantifying the mass flows passing through an infinitesimal
volume. In the present case, the control volume is stationary and the particle
(our system) is flowing through it. We start with eq. 3/6 p. 54:
d𝑚particle d
=0= 𝜌 d + ∬ 𝜌(𝑉⃗rel ⋅ 𝑛)
⃗ d𝐴
d𝑡 d𝑡 ∭CV
(6/11)
CS
The first of these two integrals can be rewritten using the Leibniz integral
rule:
d 𝜕𝜌
∭ 𝜌 d = ∭ d + ∬ 𝜌𝑉𝑆 d𝐴
d𝑡 CV 𝜕𝑡
𝜕𝜌
=∭ d
CV CS
CV 𝜕𝑡
XKCD #1524: changes in time
by Randall Munroe (CC-by-nc)
https://xkcd.com/1524 (6/12)
Figure 6.2: Conservation of mass within a fluid particle. In the 𝑥-direction, a mass
flow 𝑚̇ 1 = ∬ 𝜌1 𝑢1 d𝑧 d𝑦 is flowing in, and a mass flow 𝑚̇ 2 = ∬ 𝜌2 𝑢2 d𝑧 d𝑦 is flowing
out. These two flows may not be equal, since mass may also flow in the 𝑦- and
𝑧-directions.
Figure CC-0 Olivier Cleynen
1 Ina cfd calculation in which the grid is deforming, this term ∬CS 𝜌𝑉𝑆 d𝐴 will have to
116 be re-introduced in the continuity equation.
In the direction 𝑥, the mass flow 𝑚̇ net 𝑥 flowing through our control volume
can be expressed as:
𝜕𝑥
𝜕
=∭ (𝜌𝑢) d
CS
CV 𝜕𝑥
(6/13)
CV [ 𝜕𝑥 𝜕𝑦 𝜕𝑧 ]
=∭ ⃗ ⋅ (𝜌 𝑉⃗ ) d
∇ (6/14)
CV
(note that here we have again used the operator divergent, which we first
used in chapter 5 p. 95 — see also Appendix A3 p. 250)
Now, with these two equations 6/12 and 6/14, we can come back to equa-
tion 6/11, which becomes:
d𝑚particle d
=0= 𝜌 d + ∬ 𝜌(𝑉⃗rel ⋅ 𝑛)⃗ d𝐴
d𝑡 d𝑡 ∭CV
𝜕𝜌
0= ∭ d + ∭ ∇ ⃗ ⋅ (𝜌 𝑉⃗ ) d
CS
CV 𝜕𝑡 CV
(6/15)
Since we are only concerned with a very small volume d, we drop the
integrals, obtaining:
𝜕𝜌
⃗ ⋅ (𝜌 𝑉⃗ ) = 0
+∇
𝜕𝑡
(6/16)
• The first term on the left, 𝜕𝜌/𝜕𝑡, is the local time-change of density. For
example, when a gas contracts as it cools down, its density increases,
and the term becomes positive. In an incompressible flow, it is always
zero.
117
6.3.2 Balance of linear momentum
What is the force field applying to the fluid everywhere in space and time,
and how does that affect its velocity field? To answer this question, we write
out a momentum balance equation.
fluid particle are of only three kinds, namely weight, pressure, and shear: 1
d𝑉⃗ ⃗
𝑚particle = 𝐹weight + 𝐹⃗net, pressure + 𝐹⃗net, shear
d𝑡
(6/17)
the time-change of velocity from the reference frame of the cube, making
good use of the total time derivative tool we developed earlier p. 115:
D𝑉⃗ ⃗
𝑚particle = 𝐹weight + 𝐹⃗net, pressure + 𝐹⃗net, shear
D𝑡
𝑚particle D𝑉⃗ 1 ⃗ 1 ⃗ 1 ⃗
= 𝐹weight + 𝐹net, pressure + 𝐹net, shear
d D𝑡 d d d
D𝑉⃗ 1 ⃗ 1 ⃗ 1 ⃗
𝜌 = 𝐹weight + 𝐹net, pressure + 𝐹net, shear
D𝑡 d d d
(6/18)
Now, we rewrite the forces term on the right, and the hard work we did in
the previous chapters is paying off.
The force due to gravity is of course the weight. We have:
1 ⃗ 1
𝐹weight = 𝑚𝑔⃗ = 𝜌 𝑔⃗
d d
(6/19)
Figure 6.3: In our study of fluid mechanics, we consider only forces due to gravity,
shear, or pressure.
Figure CC-0 Olivier Cleynen
1 In some special applications, additional forces may also apply, see §6.3.4 p. 121 further
118 down.
The force due to pressure was dealt with in chapter 4 (Effects of pressure). Back
then, we expressed it with the help of the gradient of pressure in equation 4/13
p. 77, which we repeat here:
1 ⃗
⃗𝑝
𝐹net, pressure = −∇
d
(6/20)
And finally, we had dealt with the shear force in chapter 5 (Effects of shear).
With somewhat effort, we had expressed it as a function of the divergent of
shear in equation 5/19 p. 95, which we repeat here:
1 ⃗
⃗ ⋅ 𝜏⃗𝑖𝑗
𝐹net, shear = ∇
d
(6/21)
Now, we can put together all of our findings back into equation 6/18, we
obtain the Cauchy equation:w
D𝑉⃗
𝜌 ⃗𝑝 + ∇
= 𝜌 𝑔⃗ − ∇ ⃗ ⋅ 𝜏⃗𝑖𝑗
D𝑡
(6/22)
from a fixed reference frame (the acceleration field D𝑉⃗ /D𝑡) as a the sum
a fluid particle. It expresses the time change of the velocity field measured
for us. In our search for the velocity field 𝑉⃗ , the changes in time and space
Nevertheless, while it is an excellent start, this equation isn’t detailed enough
of the shear tensor 𝜏⃗𝑖𝑗 and pressure 𝑝 are unknowns. Ideally, those two
terms should be expressed solely as a function of the flow’s other properties.
Obtaining such an expression is what Claude-Louis Navier and Gabriel Stokes
set themselves to in the 19th century: we follow their footsteps in the next
paragraphs.
The Navier-Stokes equation is the Cauchy equation (eq. 6/22) applied to
Newtonian fluids. In Newtonian fluids, which we encountered in chapter 5
(§5.4.4 p. 100), shear efforts are simply proportional to the rate of strain; thus,
We had seen with eq. 5/22 p. 97 that the norm ||⃗ 𝜏𝑖𝑗 || of shear component in
the shear component of eq. 6/22 can be re-expressed usefully.
119
1
vector field:
⃗ 2 𝑉⃗ + 𝜇 ∇
⃗ ⋅ 𝜏⃗𝑖𝑗 = 𝜇 ∇
∇ ⃗ ⋅ 𝑉⃗
⃗ ∇
3 ( ) (6/23)
⃗ 2 ) do not
The details of the notation (which includes the Laplacian operator ∇
interest us at the moment; we will explore them later on.
Adding this relationship between shear and the velocity field into the last
term of equation 6/22, we obtain the Navier-Stokes equation for compressible
flow:w
D𝑉⃗ 1
𝜌 = 𝜌 𝑔⃗ − ∇ ⃗ 2 𝑉⃗ + 𝜇 ∇
⃗ 𝑝 + 𝜇∇ ⃗ ⋅ 𝑉⃗
⃗ ∇
D𝑡 3 ( ) (6/24)
How much energy is expended or received by the fluid particles as they travel
through a complex, arbitrary flow? We answer this question with an energy
balance equation. Once again, we start from the analysis of transfers on an
= +
the rate of change the net flux the rate of work done
of energy inside of heat into on the element due to
𝐴 = 𝐵 + 𝐶
the fluid element the element body and surface forces
(6/25)
Let us first evaluate term 𝐶. The rate of work done on the particle is the dot
product of its velocity 𝑉⃗ and the net force 𝐹⃗net applying to it:
We replace the content of the parentheses with the right part of the Navier-
Stokes equation above, obtaining the scalar term:
1
𝐶 = 𝑉⃗ ⋅ [𝜌 𝑔⃗ − ∇ ⃗ 2 𝑉⃗ + 𝜇 ∇
⃗ 𝑝 + 𝜇∇ ⃗ ⋅ 𝑉⃗ d
⃗ ∇
3 ( )] (6/26)
We now turn to term 𝐵, the net flux of heat into the element. We attribute this
flux to two contributions (i.e. 𝐵 = 𝑄̇ radiation + 𝑄̇ conduction ). The first contributor
is the heat transfer 𝑄̇ radiation from the emission or absorption of radiation,
which we shyly express as:
heat 𝑄̇ conduction,𝑥 expressed as a function of the power per area 𝑞 (in W m−2 )
thermal conduction through the faces of the element causes a net flow of
𝜕𝑞𝑥
𝑄̇ conduction,𝑥 = 𝑞𝑥 − 𝑞𝑥 + d𝑥 d𝑦 d𝑧
[ ( 𝜕𝑥 )]
𝜕𝑞𝑥
=− d𝑥 d𝑦 d𝑧
𝜕𝑥
𝜕𝑞𝑥
(6/28)
=− d
𝜕𝑥
(6/29)
𝜕 2𝑇 𝜕 2𝑇 𝜕 2𝑇
=𝜅 + + d
[ (𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2 ]
(6/32)
Finally, term 𝐴, the rate of change of energy inside the fluid element, can
be expressed as a function of the specific internal energy 𝑖 specific kinetic
energy 𝑒𝑘 and :
D 1
𝐴=𝜌 ( 𝑖 + 𝑉 2 ) d
D𝑡 2
(6/33)
Abstruse Goose #275: how scien-
tists see the world
by an anonymous artist (CC-by-nc)
We are therefore able quantify the change of energy of fluid particle with a https://abstrusegoose.com/275
D 1 2 𝜕 2𝑇 𝜕 2𝑇 𝜕 2𝑇
𝜌 𝑖 + 𝑉 = 𝜌 𝑞̇ + 𝜅 + +
D𝑡 ( 2 ) [ (𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2 ]
1
radiation
+𝑉⃗ ⋅ [𝜌 𝑔⃗ − ∇ ⃗ 𝑝 + 𝜇∇ ⃗ 2 𝑉⃗ + 𝜇 ∇ ⃗ ∇⃗ ⋅ 𝑉⃗
3 ( )]
(6/34)
122
6.4 Equations for incompressible flow
In this equation, we see that if 𝜌 is uniform and constant, the first term
will vanish. Once this happens, 𝜌 can be simply dropped from the second
term. This leaves us with the (much simpler) mass balance equation for
incompressible flow, also called incompressible continuity equation:
⃗ ⋅ 𝑉⃗ = 0
∇ (6/35)
This equation states that “the divergent of velocity is zero” and it is a scalar
equation. In three Cartesian coordinates, it can be re-expressed as:
𝜕𝑢 𝜕𝑣 𝜕𝑤
+ + =0
𝜕𝑥 𝜕𝑦 𝜕𝑧
(6/36)
But if you don’t (for example, if you are missing 𝑤), how
them into equation 6/36, and hope it adds up to zero.
𝑤 to 𝑥, 𝑦 and 𝑡).
lot of unknown dependencies (e.g. the functions relating
D𝑉⃗
𝜌 ⃗𝑝 + ∇
= 𝜌 𝑔⃗ − ∇ ⃗ ⋅ 𝜏⃗𝑖𝑗
Video: worst one ever? The D𝑡
incompressible Navier-Stokes
In the 𝑥-direction, the net effect of shear on the six faces is a vector expressed
as the divergent ∇⃗ ⋅ 𝜏⃗𝑖𝑥 , which, in incompressible flow, can be expressed with
with the help of viscosity 𝜇 as:
∇⃗2 ≡ ∇⃗ ⋅∇⃗
⃗ 2𝐴 ≡ ∇
∇ ⃗ ⋅∇⃗𝐴
(6/38)
⃗ 2 𝐴𝑥
⎛ ∇ ⎞ ⃗ ⋅∇
⎛ ∇ ⃗ 𝐴𝑥 ⎞
(6/39)
⃗ 2 𝐴⃗ ≡ ⎜ ∇
∇ ⃗ 2𝐴 ⎟ = ⎜ ∇
⃗ ⋅∇
⃗ 𝐴𝑦 ⎟
⎜ 2 𝑦 ⎟ ⎜ ⎟
⃗ 𝐴𝑧
⎝ ∇ ⎠ ⃗ ⋅∇
⎝ ∇ ⃗ 𝐴𝑧 ⎠
(6/40)
Figure 6.4: Shear efforts on a cubic fluid particle (already represented in fig. 5.1 p. 93).
Figure CC-0 Olivier Cleynen
124
And now, we can re-write eq. 6/37 more elegantly and generalize to three
dimensions:
⃗ ⋅ 𝜏⃗𝑖𝑥 = 𝜇 ∇
∇ ⃗ 2 𝑢 𝑖⃗ = 𝜇 ∇
⃗ 2 𝑢⃗
∇ ⃗ 2 𝑣 𝑗⃗ = 𝜇 ∇
⃗ ⋅ 𝜏⃗𝑖𝑦 = 𝜇 ∇ ⃗ 2 𝑣⃗
∇ ⃗ 2 𝑤 𝑘⃗ = 𝜇 ∇
⃗ ⋅ 𝜏⃗𝑖𝑧 = 𝜇 ∇ ⃗ 2 𝑤⃗
∇ ⃗ 2 𝑉⃗
⃗ ⋅ 𝜏⃗𝑖𝑗 = 𝜇 ∇ (6/41)
With this new expression, we can come back to the Cauchy equation (eq. 6/22
p. 119), in which we can replace the shear term with eq. 6/41. This produces
the beautiful Navier-Stokes equation for incompressible flow:
D𝑉⃗
𝜌 = 𝜌 𝑔⃗ − ∇ ⃗ 2 𝑉⃗
⃗ 𝑝 + 𝜇∇
D𝑡
(6/42)
This simplified but still formidable equation describes the property fields of
all incompressible flows of Newtonian fluids. It expresses the acceleration
field (left-hand side) as the sum of three contributions (right-hand side): those
for in equation 6/42 are the velocity (vector) field 𝑉⃗ = (𝑢, 𝑣, 𝑤) = 𝑓1 (𝑥, 𝑦, 𝑧, 𝑡)
of gravity, gradient of pressure, and divergent of shear. The solutions we look
and the pressure field 𝑝 = 𝑓2 (𝑥, 𝑦, 𝑧, 𝑡), given a set of constraints to represent XKCD #435: scientific fields
sorted by purity
the problem at hand. by Randall Munroe (CC-by-nc)
https://xkcd.com/435
Though it is without doubt charming, equation 6/42 should be remembered
for what it is really: a three-dimensional system of coupled equations. In
Cartesian coordinates this complexity is more apparent:
𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑝 𝜕 2𝑢 𝜕 2𝑢 𝜕 2𝑢
𝜌 +𝑢 +𝑣 +𝑤 = 𝜌𝑔𝑥 − +𝜇 + +
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 ] 𝜕𝑥 [ (𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2 ]
𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝑝 𝜕 2𝑣 𝜕 2𝑣 𝜕 2𝑣
(6/43)
𝜌 +𝑢 +𝑣 +𝑤 = 𝜌𝑔𝑦 − +𝜇 + +
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 ] 𝜕𝑦 [ (𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2 ]
𝜕𝑤 𝜕𝑤 𝜕𝑤 𝜕𝑤 𝜕𝑝 𝜕 𝑤 𝜕 𝑤 𝜕 𝑤
(6/44)
2 2 2
𝜌 +𝑢 +𝑣 +𝑤 = 𝜌𝑔𝑧 − +𝜇 + +
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 ] 𝜕𝑧 [ (𝜕𝑥) 2 (𝜕𝑦) 2 (𝜕𝑧)2 ]
(6/45)
Today indeed, 150 years after it was first written, no general expression
has been found for velocity or pressure fields that would solve this vector
equation in the general case. Nevertheless, in this course we will use it
directly:
After this course, the reader might also engage into computational fluid
dynamics (cfd) a discipline entirely architectured around this equation, and
to which it purposes to find solutions as fields of discrete values.
Figure 6.5: The velocity (top) and acceleration (bottom) fields in the flow field of
the computed flow described in figure 4.4 p. 78. The velocity field is the result of a
computational fluid dynamics simulation (steady, Reynolds-Averaged: a modified
version of eq. 6/42). The acceleration field is obtained based on that solution, with
equation 6/10, but could also have been obtained with the right side of equation 6/42.
Figure CC-by by Arjun Neyyathala
126
6.4.3 The Bernoulli equation (again)
We had made clear in chapter 2 (Analysis of existing flows with one dimension) that
the Bernoulli equation was very limited in scope, and that it was always safer
to approach a problem from an energy equation instead (§2.6 p. 41). As a
reminder of this fact, and as an illustration of the bridges that can be built
between integral and derivative analysis, it can be instructive to derive the
Bernoulli equation directly from the Navier-Stokes equation. This derivation
Video: Bernoulli wants to help
is not difficult to follow; it is covered in Appendix A4.3 p. 253. Navier and Stokes
by Olivier Cleynen (CC-by)
https://youtu.be/qa5wvGOcg0Q
In our analysis of fluid flow from a derivative perspective, our five physical
principles from §1.7 have been condensed into three balance equations (often
loosely referred together to as the Navier-Stokes equations). Out of these, the
first two, for conservation of mass (6/35) and linear momentum (6/42) in
incompressible flows, are often enough to characterize most free flows, and Video: the very basics of CFD
velocity field 𝑉⃗ :
should in principle be enough to find the primary unknown, which is the by Olivier Cleynen (CC-by)
https://youtu.be/bcmbRzF67Ig
0=∇⃗ ⋅ 𝑉⃗
⃗
D𝑉
𝜌 = 𝜌 𝑔⃗ − ∇ ⃗ 2 𝑉⃗
⃗ 𝑝 + 𝜇∇
D𝑡
of each fall. In fluid mechanics, even though our analysis was carried out in
the same manner, we have yet to find such one general solution — or even to
prove that one exists at all.
It is therefore tempting to attack the above pair of equations from the nu-
small increments 𝛿𝑥, 𝛿𝑦, 𝛿𝑧 and 𝛿𝑡, we could re-express the 𝑥-component of
merical side, with a computer algorithm. If one discretizes space and time in
pressure, this equation 6/46 allows us to isolate and solve for 𝛿𝑢|𝑡 , and
If we start with a known (perhaps guessed) initial field for velocity and
therefore predict what the 𝑢 velocity field would look like after a time
increment 𝛿𝑡. The same can be done in the 𝑦- and 𝑧-directions. Repeating the
obtaining at every new time step the value of 𝑢, 𝑣 and 𝑤 at each position
process, we then proceed to the next time step and so on, marching in time,
𝑢 = 2𝑥 2 − 𝑦 2 + 𝑧 2
𝑣 = 3𝑥𝑦 + 3𝑦𝑧 + 𝑧
𝑤=?
Note: Unfortunately Olivier made an error in this video: at 4:00 the term 2𝑥 2 is
incorrectly derived into 2𝑥. The final result should have −7𝑥𝑧 instead of −5𝑥𝑧
as the first term. Many thanks to the students who double-checked and reported
the problem!
128
Playing with the Navier-Stokes equation
An incompressible flow with no gravity has the velocity field defined as
follows:
Does a function exist to describe the pressure field, and if so, what is it?
129
130
Problem sheet 6: Prediction of fluid flows
last edited June 12, 2020
by Olivier Cleynen — https://fluidmech.ninja/
The atmosphere has 𝑝atm. = 1 bar; 𝜌atm. = 1,225 kg m−3 ; 𝑇atm. = 11,3 °C; 𝜇atm. = 1,5 ⋅ 10−5 Pa s
Except otherwise indicated, assume that:
Air behaves as a perfect gas: 𝑅air =287 J kg−1 K−1 ; 𝛾air =1,4; 𝑐𝑝 air =1 005 J kg−1 K−1 ; 𝑐𝑣 air =718 J kg−1 K−1
Liquid water is incompressible: 𝜌water = 1 000 kg m−3 , 𝑐𝑝 water = 4 180 J kg−1 K−1
⃗ ⋅ 𝑉⃗ = 0
∇ (6/35)
D𝑉⃗
𝜌 = 𝜌 𝑔⃗ − ∇ ⃗ 2 𝑉⃗
⃗ 𝑝 + 𝜇∇
D𝑡
(6/42)
6.1 Quiz
Once you are done with reading the content of this chapter, you can go take
the associated quiz at https://elearning.ovgu.de/course/view.php?id=7199
In the winter semester, quizzes are not graded.
6.2.1. Write out the equation in its fully-developed form in three Cartesian coordinates;
6.2.2. State in which flow conditions the equation applies.
6.2.4. Describe a situation in which the the flow is unsteady, although some property
of the fluid, when measured from the point of view of a fluid particle, is not
changing with time.
131
6.4 Volumetric dilatation rate der. Munson & al. [29] 6.4
A flow is described by the following field (in si units):
𝑢 = 𝑥3 + 𝑦2 + 𝑧
𝑣 = 𝑥𝑦 + 𝑦𝑧 + 𝑧 3
𝑤 = −4𝑥 2 𝑧 − 𝑧 2 + 4
What is the volumetric dilatation rate field (the divergent of the velocity field)? What is
the value of this rate at {2;2;2}?
6.5 Incompressibility
Does the vector field 𝑉⃗ = (1,6 + 1,8𝑥)𝑖⃗ + (1,5 − 1,8𝑦)𝑗⃗ satisfy the continuity equation for
Çengel & al. [25] 9-28
6.6 Missing components Munson & al. [29] E6.2 + Çengel & al. [25] 9-4
Two flows are described by the following fields:
𝑢1 = 𝑥 2 + 𝑦 2 + 𝑧 2
𝑣1 = 𝑥𝑦 + 𝑦𝑧 + 𝑧
𝑤1 = ?
𝑢2 = 𝑎𝑥 2 + 𝑏𝑦 2 + 𝑐𝑧 2
𝑣2 = ?
𝑤2 = 𝑎𝑥𝑧 + 𝑏𝑦𝑧 2
Verify that this field satisfies the continuity equation for incompressible flow.
132
6.9 Pressure fields Çengel & al. [25] E9-13, White [22] 4.32 & 4.34
We consider the four (separate and independent) incompressible flows below:
133
Answers
6.2 1) Continuity: eq. 6/36 p. 123. Navier-Stokes: see eqs. 6/43, 6/44 and 6/45 p. 125;
2) Read §6.3.1 p. 116 for continuity, and §6/22 p. 119 for Navier-Stokes;
3) and 4) see §6.2.2 p. 112.
D𝑉⃗
D𝑡
= (0,4 + 0,64𝑥)𝑖⃗ + (−1,2 + 0,64𝑦)𝑗⃗. At the probe it takes the value 1,68𝑖⃗ + 0,08𝑗⃗
(length 1,682 m s−2 ).
6.3
6.5
Note: the constant (initial) value 𝑝𝑖𝑛𝑖 is sometimes implicitly written in the un-
known functions 𝑓 .
6.9
134
Fluid Dynamics
Chapter 7 – Pipe flows
last edited September 3, 2020
by Olivier Cleynen — https://fluidmech.ninja/
These notes are based on textbooks by White [22], Çengel & al.[25], Munson & al.[29], and de Nevers [17].
7.1 Motivation
In this chapter we focus on fluid flow in pipes. This topic allows us to
explore several important phenomena with only very modest mathematical
complexity. In particular, we are trying to answer two questions:
𝑢 = 𝑉:
tion (eqs. 1/24, 3/6) is enough to allow us to compute the change in velocity
𝑝0 = 𝑝 + 12 𝜌𝑉av.
2 remains constant.
average velocity and pressure change with cross-section area, but the total pressure
Thus, in this kind of simple flow, pressure increases everywhere the velocity
decreases, and vice-versa.
𝑝0 = cst. (7/3)
at constant altitude, in laminar inviscid straight pipe flow.
Inviscid flows are nice, but real flows are more interesting. Real flows feature
losses due to viscosity, and with viscous effects, one key assumption of
the Bernoulli equation breaks down. An additional term will appear in the
What does this extra term Δ𝑝loss depend on, and how can we quantify it?
Bernoulli equation, as we have seen in chapter 2 with equation 2/21 p. 43.
136
7.3 Parameters to quantify losses in pipes
Hydraulics is the oldest branch of fluid dynamics, and much of the notation
used to describe pressure losses predates modern applications. The most
widely-used parameters for quantifying losses due to friction in a duct are
the following:
The elevation loss which we note |Δ𝑙| (in the literature, often noted Δℎ), is
defined as
|Δ𝑝loss |
|Δ𝑙| ≡
𝜌𝑔
(7/4)
by the fluid flow in the duct, and is measured in meters. The reference
It represents the hydrostatic height loss (with a positive number) caused
Figure 7.2: Viscous laminar fluid flow in a one-dimensional pipe. This time, the
no-slip condition at the wall creates a viscosity gradient across the duct cross-section.
This in turn translates into pressure loss. Sudden duct geometry changes such as
represented here would also disturb the flow further, but the effect was neglected
here.
Figure CC-0 Olivier Cleynen
How can we now describe quantitatively the velocity profile and the pressure
loss? We need to clearly sketch the flow we are interested in, which we do in
figure 5.10.
We also need a powerful, extensive mathematical tool to describe the flow:
we turn to the Navier-Stokes equation which we derived in the previous
chapter as eq. 6/42 p. 125:
D𝑉⃗
𝜌 = 𝜌 𝑔⃗ − ∇ ⃗ 2 𝑉⃗
⃗ 𝑝 + 𝜇∇
D𝑡
(7/7)
Figure 7.3: Two-dimensional laminar flow between two plates, also called Poiseuille
flow. We already studied this flow case in fig. 5.10 p. 106; this time, we wish to derive
an expression for the velocity distribution.
Figure CC-0 Olivier Cleynen
138
Since we are applying this tool to the simple case of fully-developed, two-
dimensional incompressible fluid flow between two parallel plates (fig. 7.3),
we need only two Cartesian coordinates, so that the vector equation trans-
lates to:
𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑝 𝜕 2𝑢 𝜕 2𝑢
𝜌 +𝑢 +𝑣 = 𝜌𝑔𝑥 − +𝜇 +
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 ] 𝜕𝑥 [ (𝜕𝑥)2 (𝜕𝑦)2 ]
(7/8)
𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝑝 𝜕 2𝑣 𝜕 2𝑣
𝜌 +𝑢 +𝑣 = 𝜌𝑔𝑦 − +𝜇 +
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 ] 𝜕𝑦 [ (𝜕𝑥)2 (𝜕𝑦)2 ]
(7/9)
𝜕𝑝 𝜕 2𝑢
0=− +𝜇
𝜕𝑥 [ (𝜕𝑦)2 ]
(7/10)
𝜕𝑝
0 = 𝜌𝑔 −
𝜕𝑦
(7/11)
1 𝜕𝑝
𝑢= (𝑦 2 − 𝐻 2 )
2𝜇 ( 𝜕𝑥 )
(7/12)
̇ 2 𝐻 2𝐻 3 𝜕𝑝
= ∫ 𝑢𝑍 d𝑦 = −
𝑍 𝑍 0 3𝜇 ( 𝜕𝑥 )
𝜕𝑝 3 𝜇 ̇
=−
𝜕𝑥 2 𝑍𝐻3
(7/13)
In this section, the overall process is more important than the result: by
starting with the Navier-Stokes equations, and adding known constraints
that describe the flow of interest, we can predict analytically all of the
characteristics of a laminar flow.
139
Figure 7.4: A cylindrical coordinate system to study laminar flow in a cylindrical
duct.
Figure CC-0 Olivier Cleynen
We once again start from the Navier-Stokes vector equation, choosing this
time to develop it using cylindrical coordinates:
This mathematical arsenal does not frighten us, for the simplicity of the
0=0
XKCD #1230: polar coordinates
by Randall Munroe (CC-by-nc)
0=0
https://xkcd.com/1230
(7/17)
𝜕𝑝 1 𝜕 𝜕𝑣𝑧
(7/18)
0= − +𝜇 𝑟
𝜕𝑧 [ 𝑟 𝜕𝑟 ( 𝜕𝑟 )]
(7/19)
𝜕 𝜕𝑣𝑧 𝑟 𝜕𝑝
𝑟 =
𝜕𝑟 ( 𝜕𝑟 ) 𝜇 𝜕𝑧
𝜕𝑣𝑧 𝑟 2 𝜕𝑝
𝑟 = + 𝑘1
( 𝜕𝑟 ) 2 𝜇 ( 𝜕𝑧 )
𝑟 2 𝜕𝑝
𝑣𝑧 = + 𝑘1 ln 𝑟 + 𝑘2
4 𝜇 ( 𝜕𝑧 )
(7/20)
140
integration constants 𝑘1 and 𝑘2 .
We have to use boundary conditions so as to unburden ourselves from
sion for the velocity profile across a pipe of radius 𝑅 when the flow is laminar:
This simplifies eq. (7/20) and brings us to our objective, an extensive expres-
1 𝜕𝑝
𝑣𝑧 = 𝑢(𝑟) = − (𝑅 2 − 𝑟 2 )
4𝜇 ( 𝜕𝑧 )
(7/21)
This equation is parabolic (fig. 7.5). It tells us that in a pipe of given length 𝐿
and radius 𝑅, a given velocity profile will be achieved which is a function
only of the ratio Δ𝑝/𝜇.
Figure 7.5: The velocity profile across a cylindrical pipe featuring laminar viscous
flow.
Figure CC-0 Olivier Cleynen
We can also express the pressure gradient in the pipe as a function of the
volume flow rate. This is done through integration of velocity with respect
to density and cross-section area. We obtain:
𝜋𝜌𝐷 4 𝜕𝑝
𝑚̇ = −
128 𝜇 ( 𝜕𝑧 )
128 𝜇𝐿𝑚̇
Δ𝑝loss =−
𝜋 𝜌𝐷 4
(7/22)
This equation is interesting in several respects. For a given pipe length 𝐿 and
pressure drop Δ𝑝loss , the volume flow ̇ increases with the power 4 of the
diameter 𝐷. In other words, the volume flow is multiplied by 16 every time
the diameter is doubled.
We also notice that the pipe wall roughness does not appear in equation 7/22.
In a laminar flow, increasing the pipe roughness has no effect on the velocity
distribution in the pipe.
Out of curiosity, we may translate the result in equation 7/22 into a friction
factor equation, with the help of definition 7/5, obtaining
32𝑉av. 𝜇𝐿 𝜇 64
𝑓laminar cylinder flow = = 64 =
𝐿1
𝜌𝑉av.
2 𝐷2 𝜌𝑉av. 𝐷
(7/23)
𝐷2
[Re]𝐷
̇
in which we inserted the average velocity 𝑉av. = 𝜋𝑅
= − Δ𝑝 𝐷 . 2
2 32𝜇𝐿
141
7.5 Turbulent flow in pipes
Figure 7.6: Illustration published by Reynolds in 1883 showing the installation he set
up to investigate the onset of turbulence. Water flows from a transparent rectangular
tank down into a transparent drain pipe, to the right of the picture. Colored die
is injected at the center of the pipe inlet, allowing for the visualization of the flow
regime.
Image by Osborne Reynolds (1883, public domain)
142
Figure 7.7: Illustration published by Reynolds in 1883 showing two different flow
regimes observed in the installation from fig. 7.6.
Image by Osborne Reynolds (1883, public domain)
The significance of the Reynolds number extends far beyond pipe flow; we
shall explore this in chapter 8 (Engineering models).
• A strong increase in mass and energy transfer within the flow. Slow and
rapid fluid particles have much more interaction (especially momentum
transfer) than within laminar flow;
𝑣𝑟 = 𝑣𝑟 + 𝑣𝑟′
𝑣𝜃 = 𝑣𝜃 + 𝑣𝜃′
𝑣𝑧 = 𝑣𝑧 + 𝑣𝑧′
Video: highly-turbulent flow ex-
iting the flood discharge ducts of
In our case, 𝑣𝑟 and 𝑣𝜃 are both zero, but the fluctuations 𝑣𝑟′ and 𝑣𝜃′ are not, and
the Tarbela dam in northeastern
Pakistan
will cause 𝑣𝑧 to differ from the laminar flow case. The extent of turbulence is
by Y:Beauty Of Pakistan (styl)
https://youtu.be/13tBWzKajqw
[ 𝑣𝑖 ]
1
′2 2
𝐼 ≡
𝑣𝑖
(7/26)
𝑟 7
𝑢 (𝑟) = 𝑣𝑧 = 𝑣𝑧 max (1 −
1
𝑅)
(7/27)
• They result in shear and dissipation rates that are markedly higher
than laminar profiles.
144
Figure 7.8: Velocity profiles for laminar (A), and turbulent (B and C) flows in a
cylindrical pipe. B represents the time-averaged velocity distribution, while C shows
several arbitrary instantaneous distributions (blurred) as well as their average in
time. Turbulent flow in a pipe also features velocities in the radial and angular
directions, which are not shown here.
Figure CC-0 Olivier Cleynen
1 1 𝜖 2,51
√ = −2 log + √
𝑓 ( 3,7 𝐷 [Re]𝐷 𝑓 )
(7/28)
values for 𝑓 (and thus predict the pressure losses Δ𝑝loss ) if we know the
Moody diagram, fig. 7.9. This classic document allows us to obtain numerical
145
Figure 7.9: A Moody diagram, which presents values for 𝑓 measured experimentally,
as a function of the diameter-based Reynolds number [Re]𝐷 , for different values of
the relative roughness 𝜖/𝐷. This figure is reproduced with a larger scale as figure 7.11
p. 152.
Diagram CC-by-sa S Beck and R Collins, University of Sheffield
• When the flow is laminar, we can solve for the flow in the pipe. For
this, we write the Navier-Stokes equations and apply known boundary
• When the flow is turbulent, this analysis method does not work any-
more. Even though the conditions are in principle simple, the flow is
mesmerizingly complex. We resort to building models based on time-
averaged measurement data. Those models work for a wide number of
This equation is telling: for a given fluid (𝜇) and a given mass flow (𝑚), ̇ the
increase linearly with mass flow 𝑚, ̇ and with the power −4 (!) of the diame-
Again, it is visible here that in laminar pipe flow, losses per unit pipe length
ter 𝐷.
𝑚̇ 2 8
Δ𝑝loss = −𝐿 𝑓 2
𝐷 5 𝜋 𝜌
(7/31)
At very turbulent regimes (in the upper right area of the diagram), 𝑓 becomes
number and to the roughness of the pipe, as described in the Moody diagram.
pipe flow increase approximately with the square of mass flow 𝑚̇ and the
Based on this equation, we can see that pressure losses in highly-turbulent
power −5 of the diameter 𝐷. In between this regime and the laminar regime,
a variety of intermediary states are quantified using the Moody diagram.
147
7.6.5 Calculating pumping and turbining power
The tools above are all we need to calculate, given the geometry of an
installation, how much pumping or turbining power is involved in moving a
given mass flow of liquid though a pipe system. The steps are as follows:
corresponding Δ𝑝loss using the definition 7/5. Care must be taken with
the sign of Δ𝑝loss , which is always negative by definition, but very often
expressed as a positive number in the literature.
The power is recovered using equation 1/22 p. 19, 𝑃̇ device = Δ𝑝device 𝑚/𝜌.
̇
hydrostatic pressure differences may have either sign.
If this power is negative, the liquid is losing energy, and the device is
acting as a turbine. If the power is positive, the liquid is gaining energy,
and the device is acting as a pump.
148
See this solution worked out step by step on YouTube
https://youtu.be/3lyKby0fS-8 (CC-by Olivier Cleynen)
Turbining power
In the piping installation from the previous examples, what is the power
made available to the turbine?
149
150
Problem sheet 7: Pipe flows
last edited June 30, 2021
by Olivier Cleynen — https://fluidmech.ninja/
The atmosphere has 𝑝atm. = 1 bar; 𝜌atm. = 1,225 kg m−3 ; 𝑇atm. = 11,3 °C; 𝜇atm. = 1,5 ⋅ 10−5 Pa s
Except otherwise indicated, assume that:
Air behaves as a perfect gas: 𝑅air =287 J kg−1 K−1 ; 𝛾air =1,4; 𝑐𝑝 air =1 005 J kg−1 K−1 ; 𝑐𝑣 air =718 J kg−1 K−1
Liquid water is incompressible: 𝜌water = 1 000 kg m−3 , 𝑐𝑝 water = 4 180 J kg−1 K−1
In cylindrical pipe flow, we assume the flow is always laminar for [Re]𝐷 . 2 300, and
always turbulent for [Re]𝐷 & 4 000. The Darcy friction factor 𝑓 is defined as:
|Δ𝑝loss |
𝑓 ≡ 𝐿1
𝜌𝑉av.
2
(7/5)
𝐷2
Viscosities of various fluids are given in fig. 7.10. Pressure losses in cylindrical pipes
can be calculated with the help of the Moody diagram presented in fig. 7.11 p. 152.
2×10−2 2.4×10−5
10−2 2.2×10−5
9×10−3
8×10−3
7×10−3 ⟵ Crude Oil
6×10−3 Air ⟶
5×10−3
2×10−5
4×10−3
3×10−3
Viscosity 𝜇 of liquids in Pa s
Viscosity 𝜇 of gases in Pa s
2×10−3 1.8×10−5
CO2 ⟶
10−3 1.6×10−5
9×10−4
8×10−4
7×10−4
6×10−4
5×10−4
1.4×10−5
4×10−4 ⟵ Water
3×10−4
2×10−4 1.2×10−5
10−4 10−5
−20 0 20 40 60 80 100 120
Temperature 𝑇 in degree Celsius ( C)
◦
Figure 7.10: The viscosity of four fluids (crude oil, water, air, and C02) as a function of temperature.
The scale for liquids is logarithmic and displayed on the left; the scale for gases is linear and 151
displayed on the right.
Figure reproduced from figure 5.6 p. 99; CC-by by Arjun Neyyathala & Olivier Cleynen
Figure 7.11: A Moody diagram, which presents values for 𝑓 measured experimentally, as a
function of the diameter-based Reynolds number [Re]𝐷 , for different relative roughness values.
Diagram CC-by-sa S Beck and R Collins, University of Sheffield
152
7.1 Reading quiz
Once you are done with reading the content of this chapter, you can go take
the associated quiz at https://elearning.ovgu.de/course/view.php?id=7199
In the winter semester, quizzes are not graded.
̇ but
increase with increasing Reynolds number?
7.2.3. Why can the pressure losses Δ𝑝losses be calculated given the volume flow ,
not the other way around?
7.3 Air flow in a small pipe Munson & al. [29] E8.5
total length of the pipe is 10 km, its diameter is 0,5 m, and its roughness is 𝜖 = 0,5 mm. It
A long pipe is installed to carry water from one large reservoir to another (fig. 7.16). The
Figure 7.10 p. 151 quantifies the viscosity of various fluids, and fig. 7.11 p. 152 quantifies
losses in cylindrical pipes.
Your first assignment is to choose the dimensions of a system which should carry 3 m3 h−1
systems.
of water (10−3 Pa s) across a horizontal distance of 1 km. Fresh from reading through
chapter 7, you design the system to feature laminar flow only.
You begin with a revision of the relevant theory, starting, of course, from the Navier-
Stokes equations and a simple diagram (fig. 7.13), obtaining the velocity distribution for
laminar flow in a circular pipe:
1 𝜕𝑝
𝑣𝑧 = 𝑢(𝑟) = − (𝑅 2 − 𝑟 2 )
4𝜇 ( 𝜕𝑧 )
(7/21)
Figure 7.13: A cylindrical coordinate system to study laminar flow in a horizontal cylindrical
duct.
Figure CC-0 Olivier Cleynen
7.5.1. What is the minimum pipe diameter for which the flow will remain laminar?
pipe with laminar flow is expressed as a function of the volume flow ̇ and the
7.5.2. Starting from equation 7/21, show that the pressure loss per unit length in the
diameter 𝐷 as:
128 𝜇𝐿𝑚̇
Δ𝑝loss = −
𝜋 𝜌𝐷 4
(7/22)
7.5.3. With the diameter chosen above, what is the pressure loss in the pipe?
7.5.4. What is the pumping power required?
With those results in hand, you turn to your colleagues — but with a smile, they suggest
154 you try a design with turbulent flow instead.
7.5.5. What would be the pressure loss if an 8 cm-diameter plastic pipe was used?
7.5.7. What is one advantage of using a pipe with smaller diameter? (briefly justify
your answer, e.g. in 30 words or less)
Your design must safely carry 700 thousand barrels of oil (110 000 m3 ) per day along a
of designing one very large oil pipeline system (fig. 7.14).
length of 1 200 km. The crude oil has density 900 kg m−3 and its viscosity is quantified in
fig. 7.10 p. 151. The average temperature of the oil during the transit is 60 °C.
The landscape is flat for most of the journey, with a 200 km-wide mountain range in the
middle that reaches 1 400 m altitude.
Your team selects a cylindrical, smooth steel duct with 1,22 m diameter, average roughness
𝜖 = 0,15 mm. Because the pipeline passes through ecologically fragile areas, as well as a
seismically-active region, you decide to never exceed 200 psi (13,8 bar) of gauge pressure
you decide to never reach below 0,8 bar of absolute pressure in the pipeline.
in the pipeline. To prevent oil cavitation (a change of state with destructive consequences),
7.6.1. How much time does the average oil particle need to travel across the line?
7.6.3. How far apart should the pumping stations be laid out in the flat sections of the
pipeline?
7.6.4. How far apart should the pumping stations be laid out in the ascending section
of the pipeline?
7.6.5. Propose a pumping station arrangement, and calculate the power required for
each pump.
Figure 7.14: The Trans-Alaska Pipeline System, which inspired this problem. It was built in the
1970s, at tremendous financial, political and social cost.
Photo CC-by-sa by Luca Galuzzi – www.galuzzi.it
155
Before you start building the pipeline, the operator would like to know how the system
would perform at half-capacity (i.e. with half the volume flow).
7.6.6. If none of the other input data changes, what is the new pumping power?
7.6.7. Propose one reason why in practice, the pumping power may be higher than
you just calculated (briefly justify your answer, e.g. in 30 words or less).
Figure 7.15: Layout of the water pipe. For clarity, the diameter of the pipe and the vertical scale
are exaggerated.
Figure CC-0 Olivier Cleynen
The pipe connecting the reservoirs is made of concrete (𝜖 = 0,25 mm); it has a diameter
of 50 cm on the first half, and 100 cm on the second half. In the middle, the conical
expansion element induces a loss coefficient of 0,8. At the outlet (at point D), the pressure
The inlet is 14 m below the surface. The total pipe length is 400 m; the altitude change
is approximately equal to the corresponding hydrostatic pressure in the outlet tank.
7.7.1. Represent qualitatively (that is to say, showing the main trends, but without
displaying accurate values) the water pressure as a function of pipe distance,
when the pump is turned off.
7.7.2. On the same graph, represent qualitatively the water pressure when the pump
is switched on.
7.8 Piping and power of a water turbine from 2018-07 final examination
A water turbine is installed between two reservoirs in order to extract power from the
The pipe is made of reinforced concrete, with a total length 𝐿 = 0,8 km, a diameter 𝐷 =
flow of water. The water is guided to the turbine through a pipe, as shown in figure 7.16.
𝑑 = 1,2 m, and an interior surface roughness 𝜖 = 6 mm. The pipe has altitude variations
along its length, as indicated in figure 7.16. The turbine is designed to handle 5 000 L s−1
156 of water at 20 °C.
Figure 7.16: Layout of the water pipe. For clarity, the vertical scale and the diameter of the pipe
are greatly exaggerated.
7.8.1. On a diagram, represent qualitatively (i.e. without numerical data) the pressure
distribution along the length of the pipe.
7.8.2. What is the pressure drop due to friction losses generated by the water flow in
the pipe?
7.8.4. What would be the new power developed by the turbine if the volume flow was
divided by two?
157
Viscosity in centipoise
0 1,005
Percentage of alcohol by weight
10 1,538
20 2,183
30 2,71
40 2,91
50 2,87
60 2,67
70 2,37
80 2,008
90 1,61
100 1,2
• The viscosity of water and alcohol mixes is described in table 7.1 (use the nearest
relevant value);
• The pipe bends induce a loss coefficient factor 𝐾𝐿bend = 0,5 each;
• The pipe T-junction induces a loss coefficient factor 𝐾𝐿 = 0,3 in the line direction
and 1 in the branching flow;
The students wish to obtain the correct mix: one quarter vodka, three quarters tonic
water. For given levels of liquid in the bottles, is there a straw pipe network configuration
that will yield the correct mix, and if so, what is it?
158
Answers
1) Calculating inlet density with the perfect gas model, [Re]𝐷 = 14 263 (turbulent),
a Moody diagram read gives 𝑓 ≈ 0,029, so Δ𝑝friction = −1 292,6 Pa = −0,0129 bar.
7.3
7.4 |Δ𝑝alt. | = 𝜌𝑔(26 − 8 + 5 − 7) = 1,57 bar and |Δ𝑝friction | = 51,87 bar : 𝑊̇ pump = 5,345 MW.
3) |Δ𝑝|loss = 0,75 Pa
is the use of cylindrical (instead of rectangular) coordinates.
4) 𝑊̇ = 0,63 mW
5) |Δ𝑝|loss 2 = 4,82 kPa (8 000 times more)
6) 𝑊̇ 2 = 4,2 W
2) 𝑊̇ total = 10,02 MW
7.6 1) Approximately 12 days 18 hours
3) Δ𝑝𝑓 A→B = −3 735 Pa, Δ𝑝B→C = +71 Pa (the sum of losses due to friction and gains
due to decrease in kinetic energy), Δ𝑝𝑓 C→D = −117 Pa. Add hydrostatic pressure
7.7
7.9 The author cannot remember which exercise you are referring to.
159
160
Fluid Dynamics
Chapter 8 – Engineering models
last edited April 3, 2021
by Olivier Cleynen — https://fluidmech.ninja/
These notes are based on textbooks by White [22], Çengel & al.[25], Munson & al.[29], and de Nevers [17].
8.1 Motivation
In this chapter, we develop tools to analyze scale effects in fluid mechanics.
This study should allow us to answer two questions:
8.2.1 Principle
In many practical engineering situations which involve fluid flow, investi-
gating the flow in the real application is impossible or impractical. In those
cases it is common practice to build scaled-up or scaled-down versions of
the flow in a laboratory. This brings up the question: how should the model
flow properties be adapted to represent the original flow? For example, if Video: when their movement is
filmed and then viewed sped-up,
the model is half as small as the original, should the velocity be halved? Or fog banks can appear to flow like
perhaps doubled? water. What time-lapse rate is
required to achieve physical sim-
ilarity?
The answer to this problem is as follows: two flows are dynamically similar by Simon Christen (stvl)
(i.e. representative of one another) when their flow parameters are the same. https://vimeo.com/69445362
In order to understand what this means, we will need to look back at the
Navier-Stokes equation (the momentum balance equation we derived in
chapter 6), re-writing it in a non-dimensional form. Onwards! 161
Advice from an expert
Making models is useful when you have a large
machine, say, a jet airliner or a submarine,
and you want to study different configurations
without building a new full-scale machine each
time. But there are other applications: imagine
trying to study the flow of molten metal in a
furnace, the flow of blood in a beating heart,
or the flow of air around a mosquito’s wings.
Making models and understanding scale effects will allow you to investi-
gate fluid flow in all kind of inaccessible locations. That makes the math
worth the effort!
𝜕 𝑉⃗
𝜌 + 𝜌(𝑉⃗ ⋅ ∇
⃗ )𝑉⃗ = 𝜌 𝑔⃗ − ∇ ⃗ 2 𝑉⃗
⃗ 𝑝 + 𝜇∇
𝜕𝑡
(8/1)
this equation. The principle is to express each vector 𝐴⃗ as the multiple of its
What we would now like to do is separate the geometry from the scalars in
length 𝐴 and a non-dimensional vector 𝐴⃗∗ , which has the same direction as
𝐴⃗ but only unit length.
𝑡∗ ≡ 𝑓 𝑡 (8/2)
A flow with a very high frequency is highly unsteady, and the changes in
time of the velocity field will be relatively important. On the other hand, the
the velocity vector field divided by its own (scalar field) length 𝑉 :
𝑉⃗
𝑉⃗ ∗ ≡
𝑉
(8/3)
⃗ 𝑝 can be replaced by ∇
the velocity field. For example, in eq. 8/1 ∇ ⃗ (𝑝 − 𝑝∞ )
the pressure changes across a field, not their absolute value, that influence
(in which 𝑝∞ can be any constant faraway pressure). Now, the pressure field
𝑝 −𝑝∞ is non-dimensionalized by dividing it by a reference pressure difference
162
𝑝0 − 𝑝∞ , obtaining:
𝑝 − 𝑝∞
𝑝∗ ≡
𝑝0 − 𝑝∞
(8/4)
𝑔⃗
𝑔⃗∗ ≡
𝑔
(8/5)
⃗ ∗,
And finally, we define a new operator, the non-dimensional del ∇
⃗∗ ≡ 𝐿 ∇
∇ ⃗ (8/6)
by a reference length 𝐿.
and the scalar fields obtained with a non-dimensional divergent, are “scaled”
These new terms allow us to replace the constituents of equation 8/1 each by
a non-dimensional “unit” term multiplied by a scalar term representing its
length or value:
𝑡∗
𝑡=
𝑓
𝑉⃗ = 𝑉 𝑉⃗ ∗
𝑝 − 𝑝∞ = 𝑝 ∗ (𝑝0 − 𝑝∞ )
𝑔⃗ = 𝑔 𝑔⃗∗
1 ∗
⃗= ∇
∇ ⃗
𝐿
𝜕 ∗ 1 ⃗∗ 1 ∗ ∗ ⃗ ∗2
∇
𝜌 1 ∗ (𝑉 𝑉 ) + 𝜌 (𝑉 𝑉 ⋅ ∇ ) (𝑉 𝑉 ) = 𝜌𝑔 𝑔⃗ − ∇ [𝑝 (𝑝0 − 𝑝∞ ) + 𝑝∞ ] + 𝜇 2 (𝑉 𝑉⃗ ∗ )
⃗ ∗ ⃗ ⃗ ∗ ∗ ⃗
𝜕𝑓 𝑡 𝐿 𝐿 𝐿
𝜕 𝑉⃗ ∗ 1 1 ∗ ∗ 1 ∗2 ⃗ ∗
𝜌𝑉 𝑓 + 𝜌𝑉 𝑉 (𝑉⃗ ∗ ⋅ ∇ ⃗ ∗ 𝑉⃗ ∗ = 𝜌𝑔 𝑔⃗∗ − ∇
)
⃗ [𝑝 (𝑝0 − 𝑝∞ )] + 𝜇𝑉 ∇ ⃗ 𝑉
𝜕𝑡 ∗ 𝐿 𝐿 𝐿2
𝜕 𝑉⃗ ∗ 𝜌𝑉 2 𝑝 − 𝑝∞ 𝜇𝑉
[𝜌𝑉 𝑓 ] + (𝑉⃗ ∗ ⋅ ∇
⃗ ∗ 𝑉⃗ ∗ = [𝜌𝑔] 𝑔⃗∗ − 0
) [ ]
⃗ ∗𝑝∗ +
∇ ⃗ ∗2 𝑉⃗ ∗
∇
𝜕𝑡 ∗ [ 𝐿 ] 𝐿 [ 𝐿 2 ]
(8/7)
i.e. kg m−2 s−2 . Multiplying each by 𝜌𝑉𝐿 2 (of dimension m2 s2 kg−1 ), we obtain
sional (unit) vectors. These bracketed terms all have the same dimension,
𝑓 𝐿 𝜕 𝑉⃗ ∗ 𝑔𝐿 𝑝0 − 𝑝∞ 𝜇
mensional Navier-Stokes equa-
+ [1] (𝑉⃗ ∗ ⋅ ∇
⃗ ∗ 𝑉⃗ ∗ =
) 𝑔⃗∗ − ⃗ ∗𝑝∗ +
∇ ⃗ ∗2 𝑉⃗ ∗
∇
tion
[ 𝑉 ] 𝜕𝑡 ∗ [𝑉 ]
2 [ 𝜌𝑉 ]2 [ 𝜌𝑉 𝐿 ]
by Olivier Cleynen (CC-by)
https://youtu.be/KD3MIqUgN1A
(8/8)
163
Equation 8/8 does not bring any information on top of the original incom-
pressible Navier-Stokes equation (eq. 8/1). Instead, it merely separates it into
two distinct kinds of components. The first, in square brackets, are a scalar
fields (purely numbers), which indicate the magnitude of the acceleration
field. The others are unit vector fields (fields of oscillating vectors, all of
length 1, and noted with stars), which represent the geometry (direction) of
the acceleration field. In this form, we can more easily observe and quantify
the weight of the different terms relative to one another. At this point, it is
time to introduce the notion of flow parameters.
[Eu] ≡
𝜌 𝑉2
(8/10)
𝑉
[Fr] ≡ √
𝑔𝐿
(8/11)
𝜌𝑉 𝐿
[Re] ≡
𝜇
(8/12)
𝜕 𝑉⃗ ∗ ⃗∗ ⋅ ∇ 1 1
+ [1] 𝑉 ⃗ ∗ 𝑉⃗ ∗ = ⃗
𝑔 − ⃗ ∗𝑝∗ +
∇ ⃗ ∗2 𝑉⃗ ∗
∇
[ ] [ Re ]
∗
𝜕𝑡 ∗ 2
[St] [Eu]
Fr
𝜕 𝑉⃗ ∗ 1 1 ∗2 ⃗ ∗
+ [1] 𝑉⃗ ∗ ⋅ ∇
⃗ ∗ 𝑉⃗ ∗ = ⃗ ∗𝑝∗ +
⃗∗ − [Eu] ∇
2 𝑔
⃗ 𝑉
∇
𝜕𝑡 ∗
[St]
[Fr] [Re]
(8/13)
Equation 8/13 is an incredibly useful tool in the study of fluid mechanics, for
two reasons:
we must generate an incoming flow with the same four parameters [St],
[Eu], [Fr], and [Re].
164
Let us therefore take the time to explore the signification of these four
parameters.
The Strouhal number [St] ≡ 𝑓 𝐿/𝑉 (eq. 8/9) quantifies the influence of un-
√
than viscosity, convection or unsteadiness.
The Froude number [Fr] ≡ 𝑉 / 𝑔 𝐿 (eq. 8/11) quantifies the relative impor-
tance of gravity effects. In practice, gravity effects only play an impor-
tant role in free surface flows, such as waves on the surface of a water
reservoir. In most other cases, gravity contributes only to a hydrostatic
effect, which has little influence over the velocity field.
The Reynolds number [Re] ≡ 𝜌 𝑉 𝐿/𝜇 (eq. 8/12, also eqs. 1/28 p. 23 & 7/24
viscous effects (𝜇). When [Re] is very large, viscosity plays a negligible
role and the velocity field is mostly dictated by the inertia of the fluid.
We return to the significance of the Reynolds number in §8.2.5 below.
The Mach number [Ma] ≡ 𝑉 /𝑐 (eq. 1/10 p. 16) compares the flow speed
𝑉 with that of the molecules within the fluid particles (the speed of
sound 𝑐). [Ma] does not appear in equation 8/13, because we already
decided to restrict ourselves to non-compressible flows. If we hadn’t,
we would be re-expressing the pressure term in that equation as a
function of [Ma], quantifying the effect of changes in density.
These five flow parameters should be thought of scalar fields within the
studied flow domain: there is one distinct Reynolds number, one Mach
number etc. for each point in space and time. Nevertheless, when describing
fluid flows, the custom is to choose for each parameter a single representative
value for the whole flow. For example, when describing pipe flow, it is
customary to quantify a representative Reynolds number [Re]𝐷 based on
the average flow velocity 𝑉av. and the pipe diameter 𝐷 (as we have seen
over an aircraft wing is often based on the free-stream velocity 𝑉∞ and the
with eq. 7/25 p. 142), while the representative Reynolds number for flow
wing chord length. Similarly, the flight Mach number [Ma]∞ displayed on
air speed 𝑉∞ and the free-stream speed of sound 𝑐∞ , rather than particular
an aircraft cockpit instrument is computed using the relative free-stream
165
8.2.4 Flow parameters obtained as force ratios
Instead of the mathematical approach covered above, the concept of flow
parameter can be approached by comparing forces in fluid flows. This method
is described for reference in Appendix A5 p. 256.
• With low [Re], the viscosity 𝜇 plays an overwhelmingly large role, and
Video: half-century-old, but
timeless didactic exploration of
how the dynamics of fluids
change with the Reynolds num-
ber, with accompanying notes the velocity of fluid particles is largely determined by that of their own
by G. I. Taylor[3]
neighbors;
• With high [Re], the momentum 𝜌𝑉 of the fluid particles plays a more
by the National Committee for Fluid
Mechanics Films (ncfmf, 1967[21]) (styl)
important role than the viscosity 𝜇, and the inertia of fluid particles
https://youtu.be/51-6QCJTAjU
affects their trajectory much more than the velocity of their neighbors.
over the length 𝐿. Indeed, from a kinematic point of view, viscous effects are
flows: it can be thought of as the likeliness of the flow being turbulent
highly stabilizing: they tend to harmonize the velocity field and smooth out
disturbances. On the contrary, when these effects are overwhelmed by inertial
effects, velocity non-uniformities have much less opportunity to dissipate,
and at the slightest disturbance the flow will become and remain turbulent
(fig. 8.1). This is the reason why the quantification of a representative
Reynolds number is often the first step engineers and scientists take when
studying a fluid flow.
166
Figure 8.1: A viscous opaque fluid is dropped into a clearer receiving static fluid with
identical viscosity. The image shows four different experiments photographed after
yielding Reynolds numbers of 0,05, 10, 200 and 3 000 respectively.
the same amount of time has elapsed. The viscosity is decreased from left to right,
As described in eq. A/36 p. 256, a low Reynolds number indicates that viscous effects
dominate the acceleration field. As the Reynolds number increases, the nature of the
velocity field changes until it becomes clearly turbulent.
By the National Committee for Fluid Mechanics Films (ncfmf, 1967[21]), with accompanying notes by Taylor[3]
Image ©1961-1969 Educational Development Center, Inc., reproduced under Fair Use doctrine
A screen capture from film Low Reynolds Number Flow at http://web.mit.edu/hml/ncfmf.html
the flow around the real-size aircraft (here, a 48 m-wide Lockheed C-141 Starlifter),
Figure 8.2: In order for the flow around a wind tunnel model to be representative of
dynamic similarity must be obtained. The value of all flow parameters must be kept
identical. This is not always feasible in practice.
Wind tunnel photo by NASA (public domain)
Full-size aircraft photo CC-by by Peter Long
167
In the design and construction of models, practical constraints must be
balanced against the need to reproduce the dynamics of fluids accurately.
They include:
168
8.4 Comparing results: coefficients
8.4.1 Principle
The general rule that we follow when we compare measurements from
experiments at different scale is that
In other words, for scientists and engineers, the hard work is making a model
where the flow coefficients are the same (and thus the original flow and the
model are physically similar). Once this is done, the comparison of results is
easy: the force and power coefficients are the same on the model and in the
original flow.
This equation tells us that the norm of the net force vector, 𝐹net , is directly
related to a term with dimensions of 𝜌|𝑉⟂ |𝐴𝑉 . In our selection of a scale by
which to measure 𝐹net , it is therefore sensible to include a term proportional to
the the density 𝜌, a term proportional to the area 𝐴, and a term proportional to
the square of velocity 𝑉 . This “scale of fluid-induced force” is conventionally
measured using the force coefficient 𝐶𝐹 :
𝐹
𝐶𝐹 ≡ 1
𝜌𝑆𝑉 2
(8/15)
2
where 𝐹
𝜌 is a reference fluid density (kg m−3 );
is the considered fluid-induced force (N);
In effect, the force coefficient relates the magnitude of the force exerted by
(𝜌𝑆𝑉 2 = 𝑚𝑉̇ ).
the fluid on an object (𝐹 ) to the rate of flow of momentum towards the object
any particular value in any given case. The choice of terms is also worth
commenting:
169
• 𝐹 can be any fluid-induced force; generally we are interested in quan-
tifying either the drag 𝐹D (the force component parallel to the free-
stream velocity) or the lift 𝐹L (force component perpendicular to the
free-stream velocity);
Force coefficients are meaningful criteria to compare and relate what is going
on in the wind tunnel and on the full-size object: in each case, we scale the
measured force according to the relevant local flow conditions.
For example, a flow case around a car A may be studied using a model B.
then the flow dynamics will be identical. The drag force 𝐹D B measured on
If dynamic similarity is maintained (and that is by no means an easy task!),
the model can then be compared to the force 𝐹D A on the real car, using
coefficients: since 𝐶𝐹D B = 𝐶𝐹D A we have:
1 1 𝐹D B 1
Video: practical application of
2 2 𝜌 𝑆 𝑉 2
2
formula one car wind tunnel
2 𝐵 B B
𝜌𝐴 𝑆A 𝑉A2
model into their “real size” race
𝐹D A = 𝐹
car values, as in problem 8.5
( 𝜌𝐵 𝑆B 𝑉B2 ) D B
by the Sauber F1 team (styl)
https://youtu.be/FQxmOQmnaGw
170
During our investigation of the flow using a small-scale model, we may be
interested in measuring not only forces, but also other quantities such as
power — this would be an important parameter, for example, when studying
a pump, an aircraft engine, or a wind turbine.
We again go back to chapter 2 (Analysis of existing flows with one dimension), and Video: Understanding scale ef-
fects will save you a lot of money
in particular eq. 2/18 p. 40. This helps us recall that power gained or lost by and embarrassment: the caution-
ary tale of Howard Hughes
a fluid flowing steadily through a control volume could be expressed as:
𝑝 1
by Olivier Cleynen (CC-by)
[ ( 𝜌 2 )]
(8/16)
of the scalar Σnet 𝜌|𝑉⟂ |𝐴𝑉 2 . A meaningful ”scale” for the power of a machine
Thus, the power gained or lost by the fluid is directly related to the magnitude
power coefficient 𝐶P :
This “scale of fluid flow-related power” is conventionally measured using the
𝑊̇
𝐶P ≡ 1
𝜌𝑆𝑉 3
(8/17)
2
soon use the shear coefficient 𝑐𝑓 in the forthcoming chapter (eq. 10/6 p. 203).
in chapter 7 (eq. 7/6 p. 137), and we shall
XKCD #687: abusing dimen-
sional analysis
by Randall Munroe (CC-by-nc)
There are many other examples; in fact, in fluid mechanics, the non-dimen- https://xkcd.com/687
171
8.5 Solved problems
172
Problem sheet 8: Engineering models
last edited September 19, 2020
by Olivier Cleynen — https://fluidmech.ninja/
The atmosphere has 𝑝atm. = 1 bar; 𝜌atm. = 1,225 kg m−3 ; 𝑇atm. = 11,3 °C; 𝜇atm. = 1,5 ⋅ 10−5 Pa s
Except otherwise indicated, assume that:
Air behaves as a perfect gas: 𝑅air =287 J kg−1 K−1 ; 𝛾air =1,4; 𝑐𝑝 air =1 005 J kg−1 K−1 ; 𝑐𝑣 air =718 J kg−1 K−1
Liquid water is incompressible: 𝜌water = 1 000 kg m−3 , 𝑐𝑝 water = 4 180 J kg−1 K−1
𝜕 𝑉⃗ ∗ 1 1 ∗2 ⃗ ∗
+ [1] 𝑉⃗ ∗ ⋅ ∇
⃗ ∗ 𝑉⃗ ∗ = ⃗ ∗𝑝∗ +
⃗∗ − [Eu] ∇
2 𝑔
⃗ 𝑉
∇
𝜕𝑡 ∗
[St] (8/13)
[Fr] [Re]
𝐹 𝑊̇
𝐶𝐹 ≡ 𝐶P ≡
1
𝜌𝑆𝑉 2 1
𝜌𝑆𝑉 3
(8/17)
2 2
173
2×10−2 2.4×10−5
10−2 2.2×10−5
9×10−3
8×10−3
7×10−3 ⟵ Crude Oil
6×10−3 Air ⟶
5×10−3
2×10−5
4×10−3
3×10−3
Viscosity 𝜇 of liquids in Pa s
Viscosity 𝜇 of gases in Pa s
2×10−3 1.8×10−5
CO2 ⟶
10−3 1.6×10−5
9×10−4
8×10−4
7×10−4
6×10−4
5×10−4
1.4×10−5
4×10−4 ⟵ Water
3×10−4
2×10−4 1.2×10−5
10−4 10−5
−20 0 20 40 60 80 100 120
Temperature 𝑇 in degree Celsius (◦C)
Figure 8.3: The viscosity of four fluids (crude oil, water, air, and CO2 ) as a function of temperature.
The scale for liquids is logarithmic and displayed on the left; the scale for gases is linear and
displayed on the right.
Figure reproduced from figure 5.6 p. 99; CC-by by Arjun Neyyathala & Olivier Cleynen
8.2.1. Write the definition of the Reynolds number and indicatie the si units for each
term.
8.2.2. What is the consequence on the velocity field of having a low Reynolds number?
(Briefly justify your answer, e.g. in 30 words or less)
8.2.3. Give one example of a high-Reynolds number flow, and one of a low-Reynolds
174 number flow.
The standard golf ball has a diameter of 42,67 mm and a mass of 45,93 g (figure 8.4). A
typical maximum velocity for such balls is 200 km h−1 .
8.2.4. If the atmospheric conditions are identical, what flow velocity needs to be
generated during the experiment in order to reproduce the flow patterns around
the real ball?
8.2.5. Would the Mach number for the real ball then be reproduced?
8.2.6. If the temperature on the golf course is −5 °C, and the temperature in the wind
tunnel is 31 °C (with identical pressure, 1 bar), what should the new wind tunnel
velocity be?
8.2.7. If the enlarged model was made out of the same materials as the real ball, how
heavy would it be?
whose test section has a diameter of 1 m (in which, obviously, the model has to fit!).
A group of students wishes to study the flow field around the aircraft, with a wind tunnel
8.3.1. If the temperature inside the tunnel is maintained at 𝑇tunnel = 20 °C, propose a
combination of wind tunnel velocity and pressure which would enable the team
to adequately model the effects of viscosity.
8.3.2. Which flow conditions would be required in a water tunnel with the same
dimensions?
Realizing the conditions calculated in the questions above cannot be realized in practice,
the group of students tries instead to study the effect of compressibility (while accepting
that viscous effects may not be adequately modeled).
8.3.3. If the maximum velocity attainable in the wind tunnel is 80 m s−1 , which tunnel
air temperature is required for compressibility effects to be modeled?
175
8.4 Scale effects on a dragonfly
A dragonfly (sketched in fig. 8.5) has a 10 cm wingspan, a mass of 80 mg, and cruises
at 4 m s−1 , beating its four wings 20 times per second.
8.4.1. Which model size and flow velocity would you use for this experiment?
8.4.2. How many wing beats per second would then be required on the model?
8.4.3. What would be the lift developed by the model during the experiment?
8.4.4. How much mechanical power would the model require, compared to the real
dragonfly?
team races a car that is 5,1 m long, 1,8 m wide, has a 610 kW power plant, and a mass
You are leading the Aerodynamics team in a successful Formula One racing team. Your
of 750 kg (figure 8.6). The car can reach a top speed of 310 km h−1 ; at that speed, the drag
force is 7,1 kN.
In order to test different aerodynamic configurations for the car, your team invests in a
the choice between two model sizes: a 60 % model and a (smaller) 50 % model.
50-million-euro wind tunnel, in which you will run tests on a model. You currently have
8.5.1. What would be the frontal area of each model, in proportion to the frontal area
of the real car?
8.5.2. What would be the volume of each model, compared to the volume of the
real car?
8.5.3. How much less weight would the 50 % model have than the 60 % model?
176
Figure 8.6: A 2017 Formula One racing car, driven by Pascal Wehrlein
Photo CC-by by Jake Archibald
From here on, your team decides to use the 50 % car model.
8.5.4. If the ambient atmospheric conditions cannot be changed, which flow speed in
the wind tunnel is required, so that the air flow around the real car is reproduced
around the model?
You are preparing for a race where the air temperature will be 30 °C. The Formula One
regulations forbid you from running your wind tunnel faster than 50 m s−1 .
8.5.5. Your team considers modifying the wind tunnel air temperature, in order to
be controlled between −10 °C and 40 °C, but the pressure remains atmospheric
compensate for the limit in the air speed. If the temperature in the tunnel can
(1 bar), what is the maximum race-track speed that can be simulated in the wind
tunnel?
8.5.6. In that case, by which factor should the model drag force measurements be
multiplied in order to correspond to the real car?
8.5.7. The wind tunnel has a 2 m by 2 m square test cross-section. What is the power
required to bring stationary atmospheric air outside (𝑇atm. = 15 °C, 𝑉atm. = 0 m s−1 )
to the desired speed and temperature in the test section?
NB: this exercise is inspired by an informative and entertaining video by the Sauber F1 team about
their wind tunnel testing, which you are encouraged to watch at https://youtu.be/KC0E0wU6inU.
177
Answers
8.2 1) See equation 8/12 and subsequent comments;
2) See §8.2.5, in practice this can be summarized in two or three sentences;
3) High-Re: air flow around an airliner in cruise; Low-Re: air flow around a dust
7) 𝑚C = 73,9 kg
𝑉3 = 418 m s−1 !
2) In water, the density cannot be reasonably controlled, and we need a velocity
1) With e.g. a model of span 60 cm, match the Reynolds number: 𝑉2 = 0,67 m s−1 ;
2) Match the Strouhal number: 𝑓2 = 0,56 Hz. Mach, Froude and Euler numbers will
8.4
3) The third model would have 42 % less mass than the second;
4) Maintaining [Re] requires 𝑉3 = 620 km h−1 (fast!);
8.5
5) Now the fastest track speed that can be studied is 31,7 m s−1 ;
6) Multiply force measurements by 1,39 to scale up to reality;
7) The required work as power is 331 kW and the power to change temperature is
−6,654 MW (better keep the cooled air in a closed circuit than feed the tunnel with
outside air!)
178
Fluid Dynamics
Chapter 9 – Dealing with turbulence
last edited August 13, 2019
by Olivier Cleynen — https://fluidmech.ninja/
9.1 Motivation
Most flows of interest to engineers and scientists are turbulent. Fluid flow
in industrial and domestic piping, in engines and in turbomachinery, is
turbulent. Flow close to solid surfaces and in the wake of objects is turbulent
at all but the slowest speeds. Blood flow in our largest veins and arteries,
and air flow in our nostrils and tracheae, are turbulent. River flows, ocean
currents, and all but the calmest winds are turbulent.
Turbulence may be ubiquitous, but it remains an incredibly complex phe-
nomenon, and describing it accurately requires either extraordinarily power-
ful numerical computations, or advanced mathematics. Neither of those is
available in this course.
We are going to treat turbulence not as a topic of research, but instead merely
as an occurrence that, as engineers, we need to account for. We will try to
answer the following two questions: In a flow, how to quantify and measure
the degree of turbulence? And how to predict its degree and its rate of decay?
In answering those, we will keep in mind that at our level, turbulence char-
acterization is not an exact science: we expect that our models may be off
by a factor of 10 or so. An imprecise or inexact understanding is better than
none at all, and, at the very least, we are developing some familiarity with
turbulence that will be useful in further studies, especially when exploring
computational fluid dynamics or experimental fluid dynamics.
179
Figure 9.1: Plankton blooming in the Atlantic ocean reveals the complexity of the
covers approximately 500 km in this take.
flow passing over the coast of Argentina. The scale of the image is so that the height
Image by Jacques Descloitres, MODIS Rapid Response Team NASA GSFC (public domain)
ity 𝜇, and so the main parameter which determines how stable a given flow
are not significantly damped. The main dampening factor in fluids is viscos-
Typically, when [Re] exceeds 1 000, the flow is very likely to be or become
is because the magnitude of the Laplacian of the velocity field is 1 000 times
turbulent (figure 9.2). We have seen in chapter 8 (Engineering models) that this
smaller than the magnitude of its advective. (Dampening factors other than
viscosity sometimes also exist, such as density gradients or interaction with
soft solid surfaces: in those cases, other non-dimensional parameters are
used).
This instability is what gives turbulence its chaotic characteristic. If a rider-
less bicycle is rolled forward and left to itself, it will continue rolling and
eventually fall to the side. Which side, left or right, depends on the initial
180
conditions: even a minute modification to the start position is likely to
influence the result. The fall is deterministic, and can be calculated very
precisely, but with a very strong dependence on the initial conditions.
Figure 9.2: A column of hot air from burning incense rising through cold air is an
unstable situation. After a certain length, the flow breaks down into chaotic patterns.
The occurrence is predicted across all fluids, plume diameters, and velocities: every
time, the Reynolds number is the determining parameter.
Photo CC-by by Rafa Espada
Turbulent motion in fluids has the same properties. Fluid motion follows
laws which are fully deterministic (the Navier-Stokes equation, eq. 6/42), but
the exact patterns in situations where the Reynolds number is high cannot
be predicted because, much like the for the bicycle above, they depend very
Video: simulation of two misci-
minutely on the initial configuration. ble fluids of different densities
layered one on top of the other
Thus, in two identical turbulent flow experiments, the details of the flow will (color representing density). The
“perfect” uniform initial situa-
be different. In this sense, turbulent flow is chaotic (depending extremely tion is unstable and leads to
sensitively on initial conditions) but not random: it remains predictable, chaotic (hard to predict) pat-
terns whose details will depend
governed by well-known deterministic laws in which chance does not play strongly on minute changes in
the initial conditions. (A 2d dns
a role. The effective engineer will determine 1) what general characteristics simulation performed with Mi-
of turbulence do remain identical in both flows, and 2) how they affect the croHH)
by Chiel van Heerwaarden (CC-by)
main, global flow characteristics. https://vimeo.com/84518319
182
Figure 9.3: Not turbulence: surface waves on the ocean (top) are complex, but
not fully-chaotic, and they feature very little dissipation. Well-known oscillatory
patterns such as the von Kármán vortex street feature one dominant frequency and
one dominant vortex size: they are not turbulent either.
Sea wave photo CC-by-sa by Tiago Fioreze (cropped)
Wake photo CC-by-sa by Jürgen Wagner
[Re]𝑑 = 105 and 107 have nearly identical spread, length, and time-averaged
not lead to main flow pattern changes anymore. For example, jet plumes at
9.3.3 Mixing
Turbulence tremendously increases mixing. The large range of the scales of
motion (i.e. the many different sizes of vortices which occur simultaneously)
increases the contact surface between two mixing fluids, for example. This
makes turbulence a desired property in many chemical reactors, or in cases
where pollutants have to be dissipated (e.g. for exhaust gases).
The same features of turbulence greatly enhance heat transfer compared to
laminar flow. Most heat exchangers for which space is important, such as
radiators, feature turbulent flow.
The increased mixing also affects exchange of momentum, with increase in
the interaction between slower and faster fluid particles. This tends to widen
areas of interaction between fluids of different velocities, such as plumes,
exhaust stacks and shear layers.
one is the average flow (𝑢, 𝑣, 𝑤), and the other the instantaneous fluctuation
in figure 9.4. We thus decompose the velocity field into two components:
flow (𝑢 ′ , 𝑣 ′ , 𝑤 ′ ):
𝑢𝑖 ≡ 𝑢 𝑖 + 𝑢𝑖′
𝑢𝑖′ ≡ 0
(9/2)
(9/3)
the time-averaged temperature 𝑇 (blue curve) and the fluctuation 𝑇 ′ , whose average
𝑇 ′ is zero (red curve).
Figure CC-0 Olivier Cleynen
184
9.4.2 Turbulence intensity
1 1
1
𝐼 ≡ [ [(𝑢 ′2 ) + (𝑣 ′2 ) + (𝑤 ′2 )]]
2
𝑉 3
(9/4)
of the flow (somewhat more in unconfined flows, and somewhat less in very
confined flows). For example, wind flow around a building will feature eddies
It follows that one may quantify a large-scale eddy Reynolds number based
on those quantities:
𝜌𝑢Λ Λ
[Re]Λ ≡
𝜇
(9/6)
is approximately equal to 1:
𝜌𝑢𝜂 𝜂
[Re]𝜂 ≡ ≈1
𝜇
(9/7)
185
Figure 9.5: Surface-relative vorticity in the Atlantic ocean. Blue color indicates
clockwise rotation, and red color anticlockwise rotation. Rotating structures of
many different sizes can be observed. In homogeneous isentropic turbulence (and
in this flow case by approximation), the size of the largest and smallest vortices are
related to one another through the Reynolds number.
doi:10.1038/s41467-018-02983-w CC-by by Z. Su, J. Wang, P. Klein, A. F. Thompson & D. Menemenlis [33]
Based on this postulate, when the turbulence has been given time and space
enough to develop fully, is homogeneous, and isotropic (has identical proper-
ties in all three directions) —these are important restrictions—, Kolmogorov
and his peers showed using dimensional analysis that
𝐿min 𝜂
= = [Re]Λ
−3/4
𝐿max Λ
𝑢𝜂
(9/8)
= [Re]Λ
−1/4
𝑢Λ
𝑡𝜂
(9/9)
= [Re]Λ
−1/2
𝑡Λ
(9/10)
These three equations are a very important result. They show that as the
Reynolds number of a flow increases, the size and time scale of the smallest
structures in the flow decreases. The higher the Reynolds number, the more
complex and more minute the details of the flow become.
1
𝑘≡ ((𝑢 ′2 ) + (𝑣 ′2 ) + (𝑤 ′2 ))
2
(9/11)
186
Turbulent kinetic energy, measured in J kg−1 , represents the amount of energy
per unit mass contained in the chaotic (turbulent) component of the fluid
To convince oneself that the parameters are related, one may insert 𝐼 (eq. 9/4)
flow velocity.
that turbulence is simply left to decay, then 𝜖 is the time rate change of 𝑘:
kinetic energy is dissipating to heat. When no turbulence is produced, so
𝜕𝑘
𝜖=−
𝜕𝑡
(9/12)
The dissipation rate is measured in W kg−1 and represents the local amount
of turbulent kinetic energy that is currently being converted to heat through
viscosity.
Through dimensional analysis, Kolmogorov and his peers showed that in ho-
mogeneous, fully-developed and isotropic turbulence, the size, characteristic
velocity, and characteristic time scale of the smallest eddies could be related
to the dissipation rate with the relationships:
𝜇3 1 4
1
𝜂=
( 𝜌3 𝜖 )
(9/13)
𝜇1 4
1
𝑢𝜂 =
(𝜌 𝜖 )
(9/14)
𝜇1 2
1
𝑡𝜂 =
(𝜌 𝜖 )
(9/15)
In the same way, we may define the dissipation density 𝐷 as 𝜕𝜖/𝜕𝑙, so that
the dissipation rate would be recovered as:
Λ
𝜖=∫ 𝐷 d𝑙 (9/17)
𝜂 187
figure 9.6, where the distribution of 𝑘 and 𝜖 across the scales of eddies is
These two integral equations are only useful to understand the meaning of
plotted.
Figure 9.6: Distribution of turbulent kinetic energy (left) and of turbulent dissipation
rate (right) in fully-developed homogeneous isotropic turbulence. The top diagrams
diagrams, the horizontal axis displays 1/𝑙, so that the small-scale eddies are on the
are in linear scale, while the bottom diagrams are in logarithmic scale. In those
right side, and large-scale eddies are on the left side.
turbulence; their features (in particular, the curves’ slopes and the ratios between Λ
Those energy and dissipation distributions are for the simplest occurrences of
and 𝜖) are used as reference cases in the study of more complex cases.
Figure CC-by-sa by Olivier Cleynen
In this figure 9.6, one may see that most of the energy is contained in the large
eddies, while most of the dissipation occurs in the small eddies.
for any given flow by solving for the change in time of the unknowns 𝑢, 𝑣
We have seen in §6.5 that in principle, the flow of fluids can be computed
algorithm. Halving each of 𝛿𝑥, 𝛿𝑦, 𝛿𝑧 and 𝛿𝑡 multiplies the total number
low speeds ([Re] = 4 ⋅ 105 , so 𝑉 ≈
flow over an airfoil at relatively
of the time step increases the total number of equations to be solved by the
50 km h−1 ). Because the com-
of equations by 16, so that soon enough the designer of the simulation will solved, 35 million cpu-hours
plete details of the flow are
performed in eq. 9/2 earlier is useful when we wish to simulate the flow
while 𝑢 ′ is the component which is too small, or occurring too quickly, for
the simulation to capture. Inserting the definition 9/2 into the 𝑥-component
of the Navier-Stokes equation for incompressible flow, we obtain:
Taking the average of this equation —thus expressing the dynamics of the
flow as we calculate them with a finite, coarse grid— yields, after some
intimidating but easily conquerable algebra:
𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑢 ′ 𝜕𝑢 ′ 𝜕𝑢 ′
𝜌 +𝑢 +𝑣 +𝑤 + 𝜌 𝑢′ + 𝑣′ + 𝑤′
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 ] [ 𝜕𝑥 𝜕𝑦 𝜕𝑧 ]
𝜕𝑝 𝜕 𝑢
2
𝜕 𝑢
2
𝜕 𝑢
2
= 𝜌𝑔𝑥 − +𝜇 + +
𝜕𝑥 [ (𝜕𝑥)2 (𝜕𝑦) 2 (𝜕𝑧)2 ]
(9/18)
189
Equation 9/18 is the 𝑥-component of the Reynolds-averaged Navier-Stokes
equation (rans). It shows that when one observes the flow in terms of the
sum of an average and an instantaneous component, the dynamics cannot be
expressed solely according to the average component. Comparing eqs. 9/18
term, called the Reynolds stress, is often re-written as 𝜌 𝜕𝑢𝑖′ 𝑢𝑗′ /𝜕𝑗. In turbulent
and 6/43 we find that an additional term has appeared on the left side. This
𝑢 ′ and 𝑣 ′ ) are strongly correlated: they are each zero on average, but their
flow, it is not zero, because the instantaneous fluctuations of velocity (e.g.
is, a local value for 𝜌 𝜕𝑢𝑖′ 𝑢𝑗′ /𝜕𝑗 is estimated everywhere, depending on the
be approximated in bulk with schemes named turbulence models. That
average values (𝑢𝑖 ). The most well-known method for doing this is the
𝑘-epsilon turbulence model, which involves solving partly arbitrary transport
equations for both 𝑘 and 𝜖. The delights, shortcomings and mysteries of
that method and more are left for the reader to discover in a good hands-on
course on the youngest and most promising area of this discipline, cfd.
• Reference works
Unfortunately, no book truly aimed at engineers is known to the au-
thor. The following books provide in-depth insight over the physics of
turbulence:
191
192
Problem sheet 9: Dealing with turbulence
last edited August 13, 2019
by Olivier Cleynen — https://fluidmech.ninja/
These notes are based on textbooks by White [22], Çengel & al.[25], Munson & al.[29], and de Nevers [17].
The atmosphere has 𝑝atm. = 1 bar; 𝜌atm. = 1,225 kg m−3 ; 𝑇atm. = 11,3 °C; 𝜇atm. = 1,5 ⋅ 10−5 Pa s
Except otherwise indicated, assume that:
Air behaves as a perfect gas: 𝑅air =287 J kg−1 K−1 ; 𝛾air =1,4; 𝑐𝑝 air =1 005 J kg−1 K−1 ; 𝑐𝑣 air =718 J kg−1 K−1
Liquid water is incompressible: 𝜌water = 1 000 kg m−3 , 𝑐𝑝 water = 4 180 J kg−1 K−1
Turbulence intensity 𝐼 :
1 1
1
𝐼 ≡ [ [(𝑢 ′2 ) + (𝑣 ′2 ) + (𝑤 ′2 )]]
2
𝑉 3
(9/4)
= [Re]Λ
−1/4
𝑢Λ
𝑡𝜂
(9/9)
= [Re]Λ
−1/2
𝑡Λ
(9/10)
𝜇3 1 4
1
𝜂=
( 𝜌3 𝜖 )
(9/13)
𝜇1 4
1
𝑢𝜂 =
(𝜌 𝜖 )
(9/14)
𝜇1 2
1
𝑡𝜂 =
(𝜌 𝜖 )
(9/15)
193
9.1 Hypothetical flow
We imagine a turbulent flow described at some point with the equations (in m s−1 )
From De Nevers [17] Ex 18.1
𝑢 = 10 + sin 𝑡
𝑣=0
(9/19)
𝑤=0
(9/20)
(9/21)
(No real turbulent flow can be described by equations this simple — but this is a nice first
Hint: ∫ sin2 𝑥 d𝑥 = 12 (𝑥 + 2 )
sin 2𝑥
+𝑏
average velocity is 0,82 m s−1 . The pressure drop caused by both friction on the walls and
turbulent dissipation is measured at −0,0286 Pa m−1 .
9.2.1. What is the non-turbulent kinetic energy per unit mass of the flow?
9.2.2. At what average rate does this kinetic energy degrade into heat?
9.2.3. If there was no heat transfer, what would be the rate of temperature increase of
the air?
Measurements are carried out to measure the turbulent intensity through the channel.
Those are displayed in figure 9.7.
Figure 9.8: Summer clouds (Cumulus humilis) form when hot moist air convected from the ground
is cooled down when it rises.
Photo CC-by-sa by en:Wikipedia User:Dwindrim
9.3.1. What is approximately the size of the smallest eddies in the cloud?
9.3.2. What is approximately the dissipation power, per unit mass of air and for the
entire cloud?
𝐷2 = 100 m?
9.3.3. What will those three values become once the cloud has grown to a diameter of
𝐿 = 2 m (figure 9.9). The tank is filled with a water-like liquid and vigorously stirred
A tank used to store chemical reactants has roughly the size of a cube of side length
9.4.1. What is approximately the size of the smallest eddies in the tank?
9.4.2. What is approximately the specific dissipation power?
A full-scale simulation (dns) of the flow was carried out, which required 500 hours of
computing time on a supercomputer. Now the same simulation is to be carried out again,
for the same flow, but with a fluid whose viscosity is half that of water.
196
Answers
9.1 p. 194
9.1.1 𝑢 = 10 m s−1
9.1.2 𝑢 ′ = sin(𝑡)
9.1.3 𝑢 ′ = 0 m s−1 (as always)
9.1.4 𝐼𝑥 = 7,07 %
9.1.5 𝐼 = 4,08 %
9.1.6 𝑘 = 0,25 J kg−1
9.2 p. 194
9.3 p. 195
9.4 p. 195
197
198
Fluid Dynamics
Chapter 10 – Flow near walls
last edited September 19, 2020
by Olivier Cleynen — https://fluidmech.ninja/
These notes are based on textbooks by White [22], Çengel & al.[25], Munson & al.[29], and de Nevers [17].
10.1 Motivation
In this chapter, we focus on fluid flow close to solid walls. In these regions,
viscous effects dominate the dynamics of fluids. This study should allow us
to answer two questions:
10.2.1 Rationale
At the very beginning of the 20th century, Ludwig Prandtl observed that for
most ordinary fluid flows, viscous effects played almost no role outside of a
very small layer of fluid along solid surfaces. In this area, shear between the
zero-velocity solid wall and the outer flow dominates the flow structure. He
named this zone the boundary layer.w
We indeed observe that around any solid object within a fluid flow, there
exists a thin zone which is significantly slowed down because of the object’s
presence. This deceleration can be visualized by measuring the velocity
profile (fig. 10.1).
(termed 𝛿, as we will see in §10.2.3) is defined as the distance where the fluid
The boundary layer is a concept, a thin invisible layer whose upper limit
200
10.2.2 Why do we study the boundary layer?
Expending our energy on solving such a minuscule area of the flow may
seem counter-productive, yet three great stakes are at play here:
𝜌 DD𝑡𝑉 = −∇
effects can be safely neglected. Fluid flow can then be described with
⃗
https://youtu.be/cgWIYSnvIEg
Figure 10.3: Fluid flow around a wing profile. When analyzing the flow, whether
analytically or within a computational fluid dynamics (cfd) simulation, the flow
domain is frequently split into three distinct areas. In the boundary layer (B), fluid
flow is dominated by viscosity. Outside of the boundary layer (A), viscous effects
are very small, and the flow can be approximately solved using the Euler equation.
Lastly, in the turbulent wake (C), characterization of the flow is very difficult and
typically necessitates experimental investigations.
Figure CC-by-sa Olivier Cleynen
𝛿 ≡ 𝑦|𝑢=0,99𝑈 (10/1)
the ∞ limit), as long as the upper limit exceeds the boundary layer thickness.
In practice, this integral can be calculated on a finite interval (instead of using
𝛿
𝑢 𝑢
𝛿 ∗∗ ≡ ∫ 1 − d𝑦
𝑈 ( 𝑈)
(10/3)
0
with the same remark regarding the upper limit. The momentum thickness,
in particular when compared to the displacement thickness, is an important
parameter used in prediction models for boundary layer separation.
for a quantification of the shear term 𝜏wall . Since we are working with the
Once these three thicknesses have been quantified, we are generally looking
202
hypothesis that the fluid is Newtonian, we merely have to know 𝑢(𝑦) to
quantify shear, according equation 5/22 which we wrote way back p. 97:
𝜕𝑢
𝜏wall 𝑦𝑥 = 𝜇
𝜕𝑦
(10/4)
In a boundary layer, the shear 𝜏wall will decrease with longitudinal distance 𝑥,
because the velocity gradient above it also decreases. Consequently, 𝜏wall will
become a function of 𝑥, so that the entire shear force will be obtained by
integration (reusing eq. 5/3 p. 92):
𝜏wall
𝑐𝑓(𝑥) ≡ 1
𝜌𝑈 2
(10/6)
2
distance 𝑥.
The shear coefficient, just like the shear, remains a function of the flow-wise
𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑝 𝜕 2𝑢 𝜕 2𝑢
𝜌 +𝑢 +𝑣 = 𝜌𝑔𝑥 − +𝜇 +
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 ] 𝜕𝑥 [ (𝜕𝑥)2 (𝜕𝑦)2 ]
(10/7)
𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝑝 𝜕 2𝑣 𝜕 2𝑣
𝜌 +𝑢 +𝑣 = 𝜌𝑔𝑦 − +𝜇 +
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 ] 𝜕𝑦 [ (𝜕𝑥)2 (𝜕𝑦)2 ]
(10/8)
Building from these two equations, we are going to add three simplifications,
which are hypotheses based on experimental observation of fluid flow in
boundary layers:
d𝑝 d𝑈
= −𝜌𝑈
d𝑥 d𝑥
(10/10)
𝜕𝑢 𝜕𝑢 1 𝜕𝑝 𝜇 𝜕 2 𝑢
𝑢 +𝑣 =− +
𝜕𝑥 𝜕𝑦 𝜌 𝜕𝑥 𝜌 (𝜕𝑦)2
d𝑈 𝜇 𝜕 2 𝑢
=𝑈 +
d𝑥 𝜌 (𝜕𝑦)2
(10/11)
layer, 𝑢(𝑥,𝑦) . Unfortunately, over a century after it has been written, we still
The main unknown in this system is the longitudinal speed profile across the
Blasius was able to show that 𝑢 is a function such that 𝑢/𝑈 = 𝑓(𝜂) ′
Based on this work, it can be shown that for a laminar boundary layer flowing
along a smooth wall, the four parameters of interest for the engineer are
204
Figure 10.5: The velocity profile obtained by Blasius (an exact solution to the Navier-
Stokes equations simplified with laminar boundary-layer hypothesis).
Figure CC-0 Olivier Cleynen
We observe that the distance 𝑥transition at which the boundary layer changes
layer only, not the outer flow.
number [Re]𝑥 ≡ 𝜌𝑈 𝑥/𝜇. The most commonly accepted prediction for the
decreased. In practice this distance depends on the distance-based Reynolds
𝑢 𝑦 7
≈( )
1
𝑈 𝛿
(10/20)
This profile has a much flatter geometry near the wall than its laminar
counterpart (fig. 10.7).
In the same way that we have worked with the laminar boundary layer
profiles, we can derive models for our characteristics of interest from this
velocity profile:
𝛿 0,16
≈
𝑥 [Re] 17
Video: a look inside a turbu-
lent boundary layer, with a com-
(10/21)
𝑥
pletely resolved (dns) fluid flow
𝛿 ∗
0,02
simulation
≈
𝑥 [Re] 17
by Linné Flow Centre
& Serc KTH [26, 27] (styl)
https://youtu.be/4KeaAhVoPIw (10/22)
𝑥
𝛿 ∗∗ 0,016
≈
𝑥 1 (10/23)
[Re]𝑥7
0,027
𝑐𝑓(𝑥) ≈ 1 (10/24)
[Re]𝑥7
206
Figure 10.7: Comparison of laminar and turbulent boundary layer profiles, both
boundary layers are much thicker than laminar ones, and that 𝑢 is only a time-
scaled to the same size. It is important to remember that in practice turbulent
average velocity.
Figure CC-0 Olivier Cleynen
When the main flow speed 𝑈 along the wall is varied, we observe that
desired trajectory to the fluid. Video: flow separation: what it
is, how to predict and avoid it
by Olivier Cleynen (CC-by)
speed gradient (d𝑈 / d𝑥 > 0), and the flatter the profile becomes. Conversely,
https://youtu.be/zB5g4aB1Sus
the geometry of the boundary layer changes. The greater the longitudinal
when the longitudinal speed gradient is negative, the boundary layer velocity
profile straightens up. When it becomes perfectly vertical at the wall, it
is such that streamlines separate from the wall: this is called separationw
(fig. 10.8).
The occurrence of separation can be predicted if we have a robust model
for the velocity profile inside the boundary layer. For this, we go back to
fundamentals, stating that at the separation point, the shear effort on
the surface must be zero:
𝜕𝑢
𝜏wall at separation = 0 = 𝜇
( 𝜕𝑦 )@𝑦=0
(10/25)
𝜕 2𝑢 d𝑝 d𝑈
𝜇 = = −𝜌𝑈
( (𝜕𝑦)2 )@𝑦=0 d𝑥 d𝑥
(10/26)
Thus, as we progressively increase the term d𝑝/ d𝑥, the term 𝜕 2 𝑢/(𝜕𝑦)2
reaches higher (positive) values on the wall surface. Nevertheless, we know
207
and flowing towards a region of increasing pressure. For clarity, the 𝑦-scale is greatly
Figure 10.8: Separation of the boundary layer. The main flow is from left to right,
exaggerated.
Figure CC-by Olivier Cleynen
that it must take a negative value at the exterior boundary of the boundary
at the wall (𝑦 = 0) the term 𝜕𝑢/𝜕𝑦 tends towards ever smaller values. Given
The existence of the inflection point within the boundary layer tells us that
enough distance 𝑥, it will reach zero value, and the boundary layer will
separate (fig. 10.9). Therefore, the longitudinal pressure gradient, which in
practice determines the longitudinal velocity gradient, is the key factor in
the analytical prediction of separation.
We shall remember two crucial points regarding the separation of boundary
layers:
Figure 10.9: The inflection point within a boundary layer about to separate.
Figure CC-by Olivier Cleynen
208
2. Laminar boundary layers are much more sensitive to separation
than turbulent boundary layers (fig. 10.10).
A widely-used technique to reduce or delay the occurrence of separa-
tion is to make boundary layers turbulent, using low-height artificial
obstacles positioned in the flow. By doing so, we increase shear-based
friction (which increases with turbulence) as a trade-off for better
resistance to stall.
Figure 10.10: The effect of decreasing Reynolds number on flow attachment over an
airfoil at constant angle of attack, with the transition point highlighted. Laminar
boundary layers are much more prone to separation than turbulent boundary layers.
Figure CC-by-sa Olivier Cleynen, based on Barlow & Pope 1999 [11]
209
10.7 Solved problems
board above their car. They drive at 10 m s−1 ; the board is 3 m long and
A successful fluid dynamics professor advertises for their course using a
1,5 m high.
Will the boundary layer on the board become turbulent? How thick will
it become?
correct, but improperly calculated. The correct result is 0,107 N. Many thanks
Note: Unfortunately Olivier made an error in this video: the final expression is
210
See this solution worked out step by step on YouTube
https://youtu.be/6i_yu1BKkVY (CC-by Olivier Cleynen)
211
212
Problem sheet 10: Flow near walls
last edited July 9, 2020
by Olivier Cleynen — https://fluidmech.ninja/
The atmosphere has 𝑝atm. = 1 bar; 𝜌atm. = 1,225 kg m−3 ; 𝑇atm. = 11,3 °C; 𝜇atm. = 1,5 ⋅ 10−5 Pa s
Except otherwise indicated, assume that:
Air behaves as a perfect gas: 𝑅air =287 J kg−1 K−1 ; 𝛾air =1,4; 𝑐𝑝 air =1 005 J kg−1 K−1 ; 𝑐𝑣 air =718 J kg−1 K−1
Liquid water is incompressible: 𝜌water = 1 000 kg m−3 , 𝑐𝑝 water = 4 180 J kg−1 K−1
In boundary layer flow, we assume that transition occurs at [Re]𝑥 & 5 ⋅ 105 .
The wall shear coefficient 𝑐𝑓 , a function of distance 𝑥,
is defined based on the free-stream flow velocity 𝑈 :
𝜏wall
𝑐𝑓(𝑥) ≡ 1
𝜌𝑈 2
(10/6)
2
Exact solutions to the laminar boundary layer along a smooth surface yield:
𝛿 4,91 𝛿∗ 1,72
=√ =√
𝑥 𝑥
(10/16)
[Re]𝑥 [Re]𝑥
𝛿 ∗∗
0,664 0,664
=√ 𝑐𝑓(𝑥) = √
𝑥
(10/18)
[Re]𝑥 [Re]𝑥
Solutions to the turbulent boundary layer along a smooth surface yield the following
time-averaged characteristics:
𝛿 0,16 𝛿∗ 0,02
≈ ≈
𝑥 [Re] 17 𝑥 [Re] 71
(10/22)
𝑥 𝑥
𝛿 ∗∗ 0,016 0,027
≈ 𝑐𝑓(𝑥) ≈
𝑥 1 1 (10/24)
[Re] 7 𝑥 [Re]𝑥7
213
2×10−2 2.4×10−5
10−2 2.2×10−5
9×10−3
8×10−3
7×10−3 ⟵ Crude Oil
6×10−3 Air ⟶
5×10−3
2×10−5
4×10−3
3×10−3
Viscosity 𝜇 of liquids in Pa s
Viscosity 𝜇 of gases in Pa s
2×10−3 1.8×10−5
CO2 ⟶
10−3 1.6×10−5
9×10−4
8×10−4
7×10−4
6×10−4
5×10−4
1.4×10−5
4×10−4 ⟵ Water
3×10−4
2×10−4 1.2×10−5
10−4 10−5
−20 0 20 40 60 80 100 120
Temperature 𝑇 in degree Celsius (◦C)
Figure 10.11: The viscosity of four fluids (crude oil, water, air, and CO2 ) as a function of tempera-
ture. The scale for liquids is logarithmic and displayed on the left; the scale for gases is linear and
displayed on the right.
Figure reproduced from figure 5.6 p. 99; CC-by by Arjun Neyyathala & Olivier Cleynen
10.2.2. Draw a few streamlines, indicate the boundary layer thickness 𝛿, and the
displacement thickness 𝛿 ∗ .
10.2.3. Explain shortly (e.g. in 30 words or less) how the transition to turbulent regime
can be triggered.
10.2.4. Explain shortly (e.g. in 30 words or less) how the transition to turbulent regime
could instead be delayed.
Figure 10.12: A thin plate positioned parallel to an incoming uniform flow. Two configurations
are studied in this exercise.
Figure CC-0 Olivier Cleynen
10.3.1. What is the shear force exerted on the top surface of the plate for each of
10.3.2. What are the shear forces when the fluid is water of viscosity 𝜇water = 1 ⋅ 10−3 Pa s?
a biplane with a 12 m wingspan (fig. 10.13). It had two wings of chord length 1,98 m
The Wright Flyer I, the first airplane capable of sustained controlled flight (1903), was
fly at very low angles of attack. Its flight speed was approximately 40 km h−1 .
stacked one on top of the other. The wing profile was extremely thin and it could only
10.4.1. If the flow over the wings can be treated as if they were flat plates, what is the
power necessary to compensate the shear exerted by the airflow on the wings
during flight?
10.4.2. Which other forms of drag would also be found on the aircraft? (give a brief
answer, e.g. in 30 words or less)
215
Figure 10.13: The Wright Flyer I, first modern airplane in history. Built with meticulous care and
impeccable engineering methodology by two bicycle makers, it made history in December 1903.
Photo by Orville Wright, 1908 (public domain)
Figure 10.14: The Airbus A340-600, a large airliner first flown in 2001.
Photo CC-by-sa by Iberia Airlines (retouched)
The cylindrical part of the fuselage has diameter 5,6 m and length 65 m.
10.5.1. What is approximately the maximum boundary layer thickness around the
fuselage?
10.5.2. What is approximately the average shear applying on the fuselage skin?
10.5.3. Estimate the power dissipated to friction on the cylindrical part of the fuselage.
10.5.4. In practice, in which circumstances could flow separation occur on the fuselage
skin? (give a brief answer, e.g. in 30 words or less)
216
Figure 10.15: Comparison of the thickness distribution of two uncambered wing profiles: an
ordinary medium-speed naca 0009 profile, and a “laminar” naca 66-009 profile.
Figure © Bertin & Cummings 2010 [24]
217
Figure 10.17: Values of the section drag coefficient 𝐶𝑑 ≡ 𝑑
2 𝑐𝜌𝑉
1 2 as a function of the section lift
coefficient 𝐶𝑙 ≡ 𝑙
2 𝑐𝜌𝑉
1 2 for both airfoils presented in fig. 10.15.
Figure © Bertin & Cummings 2010 [24], based on data by Abott & Von Doenhoff 1949 [1]
𝜕𝑢 𝜕𝑢 d𝑈 𝜇 𝜕 2 𝑢
𝑢 +𝑣 =𝑈 +
𝜕𝑥 𝜕𝑦 d𝑥 𝜌 (𝜕𝑦)2
(10/12)
𝜕𝑢 𝜕𝑣
+ =0
𝜕𝑥 𝜕𝑦
(10/13)
Identify these two equations, list the conditions in which they apply, and explain shortly
(e.g. in 30 words or less) why a boundary layer cannot separate when a favorable pressure
gradient is applied along the wall.
218
Answers
10.1 1) At trailing edge [Re]𝑥 = 5 348 thus the layer is laminar everywhere. 𝛿 will grow
from 0 to 2,01 cm (eq. 10/15 p. 205);
2) For water: 𝛿trailing edge = 4,91 mm.
𝑦-direction is greatly exaggerated, and that the outer velocity 𝑈 is identical for
10.2 1) See fig. 10.6 p. 206. At the leading-edge the velocity is uniform. Note that the
both regimes;
2) See fig. 10.4 p. 202. Note that streamlines penetrate the boundary layer;
3) and 4) See §10.4 p. 205.
10.3 𝑥transition, air = 4,898 m and 𝑥transition, water = 0,4 m. In a laminar boundary layer, insert-
√ √ 𝑥transition
𝐹𝜏 = 0,664𝐿𝑈 1,5 𝜌𝜇 [ 𝑥 ]0
ing equation 10/18 into equation 10/6 into equation 10/5 yields
.
𝐹𝜏 = 0,01575𝐿𝜌 7 𝑈 7 𝜇 7 [𝑥 7 ]𝑥
6 𝑥trailing edge
In a turbulent boundary layer, we use equation 10/24 instead and get
6 13 1
. These expressions allow the calculation of the
transition
forces below, for the top surface of the plate:
1) (air) First case: 𝐹 = 3,445 ⋅ 10−3 N; second case 𝐹 = 8,438 ⋅ 10−3 N (who would
have thought eh?);
10.5 1) 𝑥transition = 7,47 cm (the laminar part is negligible). With the equations developed
in exercise 7.3, we get 𝐹 = 24,979 kN and 𝑊̇ = 6,09 MW. Quite a jump from the
Wright Flyer I!
2) When the longitudinal pressure gradient is zero, the boundary layer cannot
separate. Thus separation from the fuselage skin can only happen if the fuselage is
flown at an angle relative to the flight direction (e.g. during a low-speed maneuver).
219
220
Fluid Dynamics
Chapter 11 – Large- and small-scale flows
last edited June 26, 2019
by Olivier Cleynen — https://fluidmech.ninja/
11.1 Motivation
This exploratory chapter is not a critical component of fluid dynamics; in-
stead, it is meant as a brief overview of two extreme cases: flows for which
viscous effects are negligible, and flows for which they are dominant. This
exploration should allow us to answer two questions:
𝜕 𝑉⃗ ∗ 1 1 ∗2 ⃗ ∗
+ [1] 𝑉⃗ ∗ ⋅ ∇
⃗ ∗ 𝑉⃗ ∗ = ⃗ ∗𝑝∗ +
⃗∗ − [Eu] ∇
2 𝑔
⃗ 𝑉
∇
𝜕𝑡 ∗
[St]
[Fr] [Re]
(11/1)
221
becomes negligible relative to the other four. Thus, our governing equation
can be reduced as follows:
𝜕 𝑉⃗ ∗ 1
+ [1] 𝑉⃗ ∗ ⋅ ∇
⃗ ∗ 𝑉⃗ ∗ ≈ ⃗ ∗𝑝∗
𝑔⃗∗ − [Eu] ∇
𝜕𝑡 ∗
[Fr]2
[St] (11/2)
Now, converting eq. 11/2 back to dimensional terms, the governing momen-
tum equation for large-scale flow becomes:
D𝑉⃗
𝜌 ⃗𝑝
= 𝜌 𝑔⃗ − ∇
D𝑡
(11/3)
We see that with the starting proposition that [Re] was large, we have re-
moved altogether the viscous (last) term from the Navier-Stokes equation.
Flows governed by this equation are called inviscid flows. Equation 11/3 is
named the Euler equation; it stipulates that the acceleration field is driven
only by gravity and by the pressure field.
222
11.3 Plotting velocity with functions
respected.
⃗ × 𝑉⃗ = 0⃗
∇ (11/4)
by definition, for an irrotational flow.
⃗ 𝜙 ≡ 𝑉⃗
∇ (11/5)
≡𝑣
𝜕𝑦
(11/7)
⃗ × 𝜓⃗ = 𝑉⃗
∇ (11/8)
In summary, we have shifted the problem from looking for 𝑢 and 𝑣, to looking
for 𝜓 and 𝜙. The existence of such functions ensures that flows can be added
and subtracted from one another yet will always result in mass-conserving,
mathematically-describable flows. If such two functions are known, then the
velocity components can be obtained (recovered) easily either in Cartesian
coordinates,
𝜕𝜙 𝜕𝜓
𝑢= =
𝜕𝑥 𝜕𝑦
(11/11)
𝜕𝜙 𝜕𝜓
𝑣= =−
𝜕𝑦 𝜕𝑥
(11/12)
or angular coordinates:
𝜕𝜙 1 𝜕𝜓
𝑣𝑟 = =
𝜕𝑟 𝑟 𝜕𝜃
1 𝜕𝜙 𝜕𝜓
(11/13)
𝑣𝜃 = =−
𝑟 𝜕𝜃 𝜕𝑟
(11/14)
• strictly steady;
• inviscid;
• incompressible;
and so it follows that if the solution to a potential flow is known, the pressure
is known everywhere, and the forces due to pressure can be calculated with
relative ease.
Nevertheless, from a science and engineering point of view, potential flows
have only limited value, because they are entirely unable to account for
turbulence, which we have seen is an integral feature of high-[Re] flows.
We should therefore use them only with great caution. Potential flows help
us model large-scale structures with very little computational cost, but this
comes with strong limitations.
𝜙 = 𝑉 𝑟 cos 𝜃
𝜓 = 𝑉 𝑟 sin 𝜃
(11/16)
(11/17)
̇
the flow:
𝜙= ln 𝑟
2𝜋
̇
(11/18)
𝜓= 𝜃
2𝜋
(11/19)
225
velocity 𝑣𝜃 = 𝑓 (𝑟, 𝜃) on the flow, in addition to which there may exist a
• Irrotational vortices, rotational patterns which impart a rotational
radial component 𝑣𝑟 :
Γ
𝜙= 𝜃
2𝜋
Γ
(11/20)
𝜓 = − ln 𝑟
2𝜋
(11/21)
𝜓 = −𝐾
𝑟
(11/23)
̇
in which 𝐾 is a constant proportional to the source/sink volume flow rate .
It was found in the 17th Century that combining a doublet with uniform flow
resulted in flow patterns that imiated “perfect” flow around a cylinder: a
flow where the fluid flows smoothly and steadily everywhere (fig. 11.3). The
stream function of that flow is:
𝑅2
𝜓 = 𝑈∞ sin 𝜃 𝑟 −
( 𝑟 )
(11/24)
1 𝜕𝜓 𝑅2
𝑣𝑟 = = 𝑈∞ cos 𝜃 1 − 2
𝑟 𝜕𝜃 ( 𝑟 )
(11/25)
𝜕𝜓 𝑅2
𝑣𝜃 = − = −𝑈∞ sin 𝜃 1 + 2
𝜕𝑟 ( 𝑟 )
(11/26)
𝑣𝑟 |𝑟=𝑅 = 0
𝑣𝜃 |𝑟=𝑅 = −2𝑈∞ sin 𝜃
(11/27)
(11/28)
Since the Bernoulli equation can be applied along any streamline in this
(steady, constant-energy, inviscid, incompressible) flow, we can express the
Figure 11.2: Left: a simple uniform steady flow; Right: a doublet, the result of a
source and a sink brought very close one to another
Figure CC-0 Olivier Cleynen
226
Figure 11.3: The addition of a doublet and a uniform flow produces streamlines for
an (idealized) flow around a cylinder.
Figure CC-by-sa by Commons User:Kraaiennest
fluid on the cylinder per unit width 𝐿, in each of the two directions 𝑥 and 𝑦:
Now, a relatively simple integration gives us the net forces exerted by the
𝐹net,𝑥 2𝜋
= −∫ 𝑝s cos 𝜃 𝑅 d𝜃 = 0
𝐿
(11/30)
0
𝐹net,𝑦 2𝜋
= −∫ 𝑝s sin 𝜃 𝑅 d𝜃 = 0
𝐿
(11/31)
0
The results are interesting, and at the time they were obtained by their author,
Jean le Rond D’Alembert, were devastating: both lift and drag are zero. This
inability to reproduce the well-known phenomena of drag is often called the
d’Alembert paradox.
To find out why the solution is not realistic, we can plot the resulting surce
pressure distribution graphically, and compared to experimental measure-
ments: this is done in fig. 11.4. Good agreement is obtained on the leading
edge of the cylinder; but as the pressure gradient becomes unfavorable, in
practice the boundary layer separates –a phenomenon that cannot be de-
scribed with inviscid flow— and a low-pressure area forms on the downstream
side of the cylinder.
227
Figure 11.4: Pressure distribution (relative to the far-flow pressure) on the surface of
a cylinder, with flow from left to right. On the left is the potential flow case, purely
symmetrical. On the right (in blue) is a measurement made at a high Reynolds
number. Boundary layer separation occurs on the second half of the cylinder, which
prevents the recovery of leading-edge pressure values, and increases drag.
Figure CC-by-sa Commons User:BoH & Olivier Cleynen
𝑅2 Γ
𝜓 = 𝑈∞ sin 𝜃 𝑟 − − ln 𝑟
( 𝑟 ) 2𝜋
(11/32)
Video: watch the Brazilian foot-
ball team show their French
counterparts how circulation (in-
duced through friction by ball ro-
With this function, several key characteristics of the flow field can be obtained.
tation) is associated to dynamic The first is the velocity field at the cylinder surface:
𝑣𝑟 |𝑟=𝑅 = 0
lift on a circular body
Γ
by TF1, 1997 (styl)
(11/33)
𝑣𝜃 |𝑟=𝑅 = −2𝑈∞ sin 𝜃 +
https://youtu.be/oGeMZ3t8jn4
2𝜋𝑅
(11/34)
𝐹net,𝑥 2𝜋
= −∫ 𝑝s cos 𝜃 𝑅 d𝜃 = 0
𝐿
(11/35)
0
𝐹net,𝑦 2𝜋
= −∫ 𝑝s sin 𝜃 𝑅 d𝜃 = −𝜌 𝑈∞ Γ
𝐿
(11/36)
0
flow— but that lift occurs which is proportional to the free-stream velocity 𝑈
We thus find out that the drag is once again zero —as for any potential
cause real flows to differ from the ideal case described here, and it turns out
that rotating cylinders are a horribly uneconomical and unpractical way of
generating lift.
229
Figure 11.6: Potential flow around an airfoil without (top) and with (bottom) circu-
can only result in a net vertical force if an irrotational vortex (with circulation Γ) is
lation. Much like potential flow around a cylinder, potential flow around an airfoil
added on top of the flow. Only one value for Γ will generate a realistic flow, with the
rear stagnation point coinciding with the trailing edge, a occurrence named Kutta
condition.
Figure CC-0 Olivier Cleynen
Figure 11.7: Potential flow allows all velocities to be inverted without any change in
the flow geometry. Here the flow around an airfoil is reversed, displaying unphysical
behavior.
Figure CC-0 Olivier Cleynen
Figure 11.8: A numerical model of flow around an airfoil. In the left figure, the
velocity vectors are represented relative to a stationary background. In the right
figure, the velocity of the free-stream flow has been subtracted from each vector,
bringing the circulation phenomenon into evidence.
Figures 1 & 2 CC-by-sa by en:Wikipedia User:Crowsnest
230
Figure 11.9: From top to bottom, the lift distribution over the wings of a glider is
modeled with increasingly complex (and accurate) lift and circulation distributions
along the span. The Lifting-line theory is a method associating each element of
lift with a certain amount of circulation. The effect of each span-wise change of
circulation is then mapped onto the flow field as a trailing vortex.
Figures 1, 2 & 3 CC-by-sa Olivier Cleynen
231
11.4 Flow at very small scales
At the complete opposite of the spectrum, we find flow at very small scales:
flows the representative length 𝐿 is extremely small, which makes for small
flows around bacteria, dust particles, and inside very small ducts. In those
values of the Reynolds number. Such flows are termed creeping or Stokes
flows. What are their main characteristics?
Looking back once again at the non-dimensional Navier-Stokes equation for
incompressible flow derived as eq. 8/13 p. 164,
𝜕 𝑉⃗ ∗ ⃗∗ ⋅ ∇ 1 1 ∗2 ⃗ ∗
+ [1] 𝑉 ⃗ ∗ 𝑉⃗ ∗ = ⃗
𝑔 ∗
− ⃗ ∗𝑝∗ +
∇ ⃗ 𝑉
∇
𝜕𝑡 ∗ 2
[St] [Eu] (11/37)
[Fr] [Re]
⃗ ∗2 𝑉⃗ ∗ then becomes
smaller than 1. The relative weight of the term (1/[Re]) ∇
we see that creeping flow will occur when the Reynolds number is much
In addition to cases where [Re] ≪ 1, we focus our interest on flows for which:
overwhelming.
With these characteristics, the terms associated with the [St] (Strouhal) and
[Fr] (Froude) numbers become very small with respect to the other terms, and
our non-dimensionalized Navier-Stokes equation (eq. 11/37) is approximately
1 ∗2 ⃗ ∗
reduced to:
0⃗ ≈ −[Eu] ∇
⃗ ∗𝑝∗ + ⃗ 𝑉
∇ (11/38)
[Re]
∇ ⃗ 2 𝑉⃗
⃗ 𝑝 = 𝜇∇ (11/39)
In this type of flow, the pressure field is entirely dictated by the Laplacian
which the representative length 𝐿 is very small, spend their lives in such
of velocity, and the fluid density has no importance. Micro-organisms, for
flows (fig. 11.10). At the human scale, we can visualize the effects of these
flows by moving an object slowly in highly-viscous fluids (e.g. a spoon in
honey), or by swimming in a pool filled with plastic balls. The inertial effects
are almost inexistent, drag is extremely important, and the object geometry
has comparatively small influence.
In 1851, George Gabriel Stokes worked through equation 11/39 for flow
allowed him to show that the drag 𝐹D sphere applying on a sphere of diameter
around a sphere, and obtained an analytical solution for the flow field. This
232
Reynolds numbers, since their scale 𝐿 is very small. For them, viscosity effects
Figure 11.10: Micro-organisms carry themselves through fluids at extremely low
dominate inertial effects.
Photo by Yutaka Tsutsumi, M.D., Fujita Health University School of Medicine
Inserting this equation 11/40 into the definition of the drag coefficient 𝐶𝐹 D ≡
𝐹D / 12 𝜌𝑆frontal 𝑈∞2 (from eq. 8/15 p. 169) then yields:
𝐹D sphere 24𝜇 24
𝐶𝐹 D = = =
1
𝜌𝑈∞2 𝜋4 𝐷 2 𝜌 𝑈∞ 𝐷 [Re]𝐷
(11/41)
2
These equations are specific to flow around spheres, but the trends they
describe apply well to most bodies evolving in highly-viscous flows, such
as dust or liquid particles traveling through the atmosphere. Drag is only
proportional to the speed (as opposed to low-viscosity flows in which it
grows with velocity squared), and it does not depend on fluid density.
Figure 11.11: Flow at very low Reynolds numbers around a sphere. In this regime,
the drag force is proportional to the velocity.
Figure CC-by-sa by Olivier Cleynen & Commons User:Kraaiennest
233
234
Problem sheet 11: Large- and small-scale flows
last edited June 26, 2019
by Olivier Cleynen — https://fluidmech.ninja/
These notes are based on textbooks by White [22], Çengel & al.[25], Munson & al.[29], and de Nevers [17].
The atmosphere has 𝑝atm. = 1 bar; 𝜌atm. = 1,225 kg m−3 ; 𝑇atm. = 11,3 °C; 𝜇atm. = 1,5 ⋅ 10−5 Pa s
Except otherwise indicated, assume that:
Air behaves as a perfect gas: 𝑅air =287 J kg−1 K−1 ; 𝛾air =1,4; 𝑐𝑝 air =1 005 J kg−1 K−1 ; 𝑐𝑣 air =718 J kg−1 K−1
Liquid water is incompressible: 𝜌water = 1 000 kg m−3 , 𝑐𝑝 water = 4 180 J kg−1 K−1
11.1 Volcanic ash from the Eyjafjallajökull Çengel & al. [25] E10.2
In 2010, a volcano with a complicated name and unpredictable mood decided to ground
We consider a microscopic ash particle released at very high altitude (−50 °C, 0,55 bar,
the entire European airline industry for five days.
11.1.2. Will this terminal velocity increase or decrease as the particle progresses towards
the ground? (briefly justify your answer, e.g. in 30 words or less)
water drop with diameter 42,4 µm is falling through air at 25 °C and 1 bar.
A rainy day provides yet another opportunity for exploring fluid dynamics (fig. 11.12). A
11.2.2. Which velocity will it reach once its diameter will have doubled?
235
fall. When their diameter is lower than 2 mm, water drops are approximately spherical (B). As
Figure 11.12: A sketched diagram showing the geometry of water drops of various sizes in free
they grow beyond this size, their shape changes and they eventually break-up (C-E). They never
display the “classical” shape displayed in A, which is caused only by surface tension effects when
they drip from solid surfaces.
Figure CC-by-sa by Ryan Wilson
11.3 Idealized flow over a hangar roof based on White [22] P8.54
Certain flows in which both compressibility and viscosity effects are negligible can be
described using the potential flow assumption (the hypothesis that the flow is everywhere
irrotational). If we compute the two-dimensional laminar steady fluid flow around a
cylinder profile, we obtain the velocities in polar coordinates as:
𝑅2
𝑣𝑟 = 𝑉∞ cos 𝜃 1 − 2
( 𝑟 )
(11/25)
𝑅2
𝑣𝜃 = −𝑉∞ sin 𝜃 1 + 2
( 𝑟 )
(11/26)
In this exercise, we study the air flow over a hangar roof with this model. We use the
equations above to describe the air velocity everywhere, pretending the as the wind
blows about a large semi-cylindrical solid structure — an idealized description of an
is 𝑅 = 20 m.
hangar with a semi-cylindrical geometry, as shown in fig. 11.13. The radius of the hangar
236
Figure 11.13: A semi-cylindrical hangar roof. Wind with uniform velocity 𝑈 flows perpendicular
to the cylinder axis.
Figure CC-0 Olivier Cleynen
11.3.1. Starting from eqs. 11/25 and 11/26, show that the pressure 𝑝𝑠 on the surface on
1
the roof is distributed as:
11.3.2. The pressure inside the hangar is set to 𝑝∞ . What is the total lift force on the
hangar?
(a couple of hints to help with the algebra: ∫ sin 𝑥 d𝑥 = − cos 𝑥+𝑘 and ∫ sin3 𝑥 d𝑥 =
(see also problem 4.6 p. 87)
1
3
cos3 𝑥 − cos 𝑥 + 𝑘).
11.3.4. Describe briefly (e.g. in 30 words or less) two reasons why the results above
would not correspond to reality.
11.4 Cabling of the Wright Flyer derived from Munson & al. [29] 9.106
The Wright Flyer I, the first powered and controlled aircraft in history, was subjected
to multiple types of drag. We have already studied viscous friction on its thin wings in
exercise 7.4. The data in figure 11.14 provides the opportunity to quantify drag due to
A network of metal cables with diameter 1,27 mm criss-crossed the aircraft in order to
pressure.
237
Figure 11.14: Experimental measurements of the drag coefficient applying to a cylinder and to a
sphere as a function of the diameter-based Reynolds number [Re]𝐷 , shown together with schematic
depictions of the flow around the cylinder. By convention, the drag coefficient 𝐶D ≡ 𝐶𝐹 D ≡ 1 𝜌𝑆𝑈
𝐹D
2
(eq. 8/15 p. 169) compares the drag force 𝐹D with the frontal area 𝑆.
2 ∞
238
11.5 Ping pong ball Munson & al. [29] E9.16
A series of experiments is conducted in a wind tunnel on a large cast iron ball with a
smooth surface; the results are shown in fig. 11.15. These measurement data are used to
11.5.3. How would the drag and lift applying on the ball evolve if the air viscosity was
progressively decreased to zero?
Figure 11.15: Experimental measurements of the lift and drag coefficients applying on a rotating
sphere in an steady uniform flow.
Figure © from Munson & al.[29]
239
11.6 Flow field of a tornado Çengel & al. [25] E9-5, E9-14 & E10-3
In this problem, we attempt to model a very large-scale flow: that of a tornado (fig. 11.16).
We begin by pretending the tornado is one perfectly straight, stationary structure. We
divide the flow into two regions: a core cylinder that rotates almost like a solid body,
and an outer region where flow spins in an irrotational matter. This model is called the
Rankine vortex (displayed in fig. 11.17) and is used widely as a simple, first approximation
to model flows as large as a hurricane and as small as turbulence-induced vortices.
Figure 11.17: Modeled angular velocity in a vortex, according to the Rankine vortex model
Figure CC-by-sa by en:Wikipedia User:Justin1569
in which Γ is the circulation (measured in s−1 ) and remains constant and uniform.
11.6.1. The mass balance equation for incompressible flow (eq. 6/35 p. 123) is developed
in cylindrical coordinates as follows:
1 𝜕𝑟𝑣𝑟 1 𝜕𝑣𝜃 𝜕𝑣𝑧
+ + =0
𝑟 𝜕𝑟 𝑟 𝜕𝜃 𝜕𝑧
(11/44)
According to this mass balance equation, what form must the radial velocity 𝑣𝑟
240 take?
Among all the possibilities for 𝑣𝑟 , we choose the simplest form, so that from now on, we
model radial velocity as:
𝑣𝑟 = 0 (11/45)
11.6.2. The momentum balance equation for incompressible flow (eq. 6/42 p. 125) is
developed in cylindrical coordinates are as follows:
Starting from those equations, show that the pressure distribution in the outer
region of the tornado can be expressed as:
1 1
𝑝 = 𝑝∞ − 𝜌 Γ2 2
2 𝑟
(11/49)
We now turn to the core of the tornado, which we model as if it were a rotating solid (a
vortex core).
11.6.6. According to the model, what is the lowest pressure attained by the air?
50 km h−1 ?
11.6.7. According to the model, at what distance from the core are winds lower than
(curious students may play with the above model by adding a non-zero radial velocity,
and look up the phenomenon of vortex stretching) 241
11.7 Lift on a symmetrical object non-examinable
Briefly explain (e.g. with answers 30 words or less) how lift can be generated on a sphere
or a cylinder,
11.8 Air flow over a wing profile From Munson & al. [29] 9.109
The characteristics of a thin, flat-bottomed airfoil are examined by a group of students
in a wind tunnel. The first investigations focus on the boundary layer, and the research
group evaluate the boundary layer thickness and make sure that it is fully attached.
Measurements of the longitudinal speed 𝑢 just above the boundary layer on the top
Once this is done, the group proceeds with speed measurements all around the airfoil.
242
Answers
to obtain 𝑈 = 𝑔𝜌sphere 18𝜇 = 0,1146 m s−1 : unbearably slow when you are stuck in an
11.1 1) At terminal velocity, the weight of the sphere equals the drag. This allows us
𝐷2
airport! With 𝑈 , check that the Reynolds number indeed corresponds to creeping
flow: [Re]𝐷 = 0,334.
11.2 Same as previous exercise: 𝑈1 = 4,578 ⋅ 10−2 m s−1 and 𝑈2 = 0,183 m s−1 , with Reynolds
numbers of 0,113 and 0,906 respectively (thus creeping flow hypothesis valid).
11.3 1) Integrate the vertical component of force due to pressure: 𝐹L roof = 1,575 MN.
243
244
Appendix
last edited February 12, 2021
A1 Notation 246
A2 Vector operations 247
A2.1 Vector dot product 247
A2.2 Vector cross product 247
A3 Field operators 250
A3.1 Gradient 250
A3.2 Divergent 250
A3.3 Advective 251
A3.4 Laplacian 251
A3.5 Curl 252
A4 Derivations of the Bernoulli equation 253
A4.1 The Bernoulli equation from the energy equation 253
A4.2 The Bernoulli equation from the integral momentum equation 253
A4.3 The Bernoulli equation from the Navier-Stokes equation 253
A5 Flow parameters as force ratios 256
A5.1 Acceleration vs. viscous forces: the Reynolds number 256
A5.2 Acceleration vs. gravity force: the Froude number 257
A5.3 Acceleration vs. elastic forces: the Mach number 257
A5.4 Other force ratios 258
A6 Details of the winter 2020-2021 final examination (updated Febru-
ary 2021) 259
A7 Example of previous examinations 261
A8 References 292
A1 Notation
≡ By definition. The ≡ symbol sets the definition of the term on its left
(which does not depend on previous equations).
̇ (dot above symbol) Time rate: ̇ ≡ d𝑡d . For example, 𝑄̇ is the rate of heat
(in watts) representing a heat quantity 𝑄 (in joules) every second.
vectors Vectors are written with an arrow. Velocity is 𝑉⃗ ≡ (𝑢, 𝑣, 𝑤), alter-
natively written 𝑢𝑖 ≡ (𝑢, 𝑣, 𝑤). The norm of a vector 𝐴⃗ (positive or
⃗ its length (always positive) is ||𝐴||.
negative) is |𝐴|, ⃗
units Units are typed in roman (normal) font and colored gray (1 kg). In
sentences units are fully-spelled and conjugated (one hundred watts).
The liter is noted L to increase readability (1 L ≡ 10−3 m3 ). Units in
equations are those from système international (si) unless otherwise
indicated.
ceded by a dot, integers are written in groups of three (1,234 ⋅ 103 = 1 234).
numbers The decimal separator is a comma, the decimal exponent is pre-
246
A2 Vector operations
For a step-by-step revision of those notions and many more, written in a
progressive, nonjudgmental way, with plenty of worked-out exercises, you
can try John Bird’s Higher Engineering Mathematics [15].
𝑎⃗ ⋅ 𝑏⃗ = |𝑎| ⃗ cos 𝜃
⃗ |𝑏| (A/1)
⃗
where 𝜃 is the angle separating the two vectors 𝑎⃗ and 𝑏.
⃗
In this document, the dot product is always written with a median dot (𝑎⃗ ⋅ 𝑏),
but in other literature, it is sometimes written with the × symbol. Take care
⃗ 𝑏 , 𝑦𝑏 , 𝑧𝑏 }
not to confuse it with the vector cross product (see §A2.2 p. 248).
It can be shown that the dot product of two vectors 𝑎{𝑥 ⃗ 𝑎 , 𝑦𝑎 , 𝑧𝑎 } and 𝑏{𝑥
𝑎⃗ ⋅ 𝑏⃗ = 𝑥𝑎 𝑥𝑏 + 𝑦𝑎 𝑦𝑏 + 𝑧𝑎 𝑧𝑏
can be quantified as:
(A/2)
The dot product of two vectors is the same regardless of the order in which
𝑎⃗ ⋅ 𝑏⃗ = 𝑏⃗ ⋅ 𝑎⃗
they are multiplied:
(A/3)
−𝑎 ⋅ →
(→
− ⃗
−𝑏) = −(𝑎⃗ ⋅ 𝑏) (A/4)
⃗
Figure A.1: Two vectors 𝑎⃗ and 𝑏. 247
A2.2 Vector cross product
The cross productw of two vectors is a vector written as:
𝑎⃗ ∧ 𝑏⃗ = 𝑐⃗ (A/5)
𝑐 = 𝑎 𝑏 sin 𝜃
its length is equal to
(A/6)
⃗
its direction is perpendicular to 𝑎⃗ and 𝑏;
In this document, the cross product is written with a wedge symbol (𝑎⃗ ∧ 𝑏) ⃗
sions.
but in the literature, it is often written with the symbol ×. Make sure you do
⃗ 𝑏 , 𝑦𝑏 , 𝑧𝑏 } is:
𝑏{𝑥
It can be shown that the cross product
| 𝑖⃗ 𝑗⃗ 𝑘⃗ ||
|
| |
𝑐⃗ = |𝑥𝑎 𝑦𝑎 𝑧𝑎 ||
|
|𝑥𝑏 𝑦𝑏 𝑧𝑏 ||
(A/7)
|
So that one obtains:
|𝑦 𝑧 | |𝑥 𝑧 | |𝑥 𝑦 |
𝑐⃗ = || 𝑎 𝑎 || 𝑖⃗ − || 𝑎 𝑎 || 𝑗⃗ + || 𝑎 𝑎 || 𝑘⃗
|𝑦𝑏 𝑧𝑏 | |𝑥𝑏 𝑧𝑏 | |𝑥𝑏 𝑦𝑏 |
(A/8)
The two vectors 𝑎⃗ ∧ 𝑏⃗ and 𝑏⃗ ∧ 𝑎⃗ are pointing away one from the other (fig. A.3):
𝑏⃗ ∧ 𝑎⃗ = − (𝑎⃗ ∧ 𝑏)
⃗ (A/10)
Figure A.2: Two vectors 𝑎⃗ and 𝑏.⃗ The vector product 𝑎⃗ ∧ 𝑏⃗ has length the product of
the lengths 𝑏⟂ et 𝑎. In the case shown here, the vector 𝑐⃗ = 𝑎⃗ ∧ 𝑏⃗ is going into through
the document plane, going away from the reader.
248
Figure A.3: The vectors 𝑎⃗ ∧ 𝑏⃗ and 𝑏⃗ ∧ 𝑎⃗ have the same length but are pointing
directions opposite one from the other (the first away from the reader, and the other
towards the reader).
If any vector changes direction, the cross product also changes direction
→
− ⃗
𝑎⃗ ∧ −𝑏 = − (𝑎⃗ ∧ 𝑏)
(fig. A.4):
(A/11)
A3.1 Gradient
𝜕 𝜕 𝜕
defined as:
⃗ ≡ 𝑖⃗ + 𝑗⃗ + 𝑘⃗
∇
𝜕𝑥 𝜕𝑦 𝜕𝑧
(A/12)
𝜕𝐴 ⃗ 𝜕𝐴 ⃗ 𝜕𝐴 ⃗ ⎛⎜ ⎞
𝜕𝐴
⃗𝐴 ≡ ⎟
𝜕𝑥
∇ 𝑖+ 𝑗+ 𝑘= 𝜕𝐴
𝜕𝑥 𝜕𝑦 𝜕𝑧 ⎜ 𝜕𝑦
⎟
⎝ ⎠
(A/13)
𝜕𝐴
𝜕𝑧
⃗ 𝑝:
For example, the gradient of a pressure field is the vector field ∇
𝜕𝑝 ⃗ 𝜕𝑝 ⃗ 𝜕𝑝 ⃗ ⎛⎜ ⎞
𝜕𝑝
⃗𝑝 ≡ ⎟
𝜕𝑥
∇ 𝑖+ 𝑗+ 𝑘=
𝜕𝑝
𝜕𝑥 𝜕𝑦 𝜕𝑧 ⎜ 𝜕𝑦
⎟
⎝ ⎠
(A/14)
𝜕𝑝
𝜕𝑧
A3.2 Divergent
𝜕 ⃗ 𝜕 ⃗ 𝜕 ⃗
⃗⋅ ≡
∇ 𝑖⋅ + 𝑗 ⋅ + 𝑘⋅
𝜕𝑥 𝜕𝑦 𝜕𝑧
(A/15)
⎛ 𝜕𝐴𝑥𝑥
+ 𝜕𝑦𝑦𝑥 + 𝜕𝐴𝜕𝑧𝑧𝑥
𝜕𝐴
⎞ ⎛ ⃗ ⋅ 𝐴⃗𝑖𝑥
∇ ⎞
⎜ 𝜕𝑥
⎟ ⎜ ⎟
⃗ ⋅ 𝐴⃗𝑖𝑗 ≡ ⎜
∇
𝜕𝐴𝑥𝑦
+ 𝜕𝑦𝑦𝑦 + 𝜕𝑧𝑧𝑦
𝜕𝐴 𝜕𝐴
⎟=⎜ ⃗ ⋅ 𝐴⃗𝑖𝑦
∇ ⎟
⎜ ⎟ ⎜ ⎟
𝜕𝑥
+ 𝜕𝑦𝑦𝑧 + 𝜕𝐴𝜕𝑧𝑧𝑧 ⃗ ⋅ 𝐴⃗𝑖𝑧
∇
(A/18)
𝜕𝐴𝑥𝑧 𝜕𝐴
⎝ 𝜕𝑥 ⎠ ⎝ ⎠
⃗ ⋅ 𝑉⃗ :
For example, the divergent of a velocity field is the scalar field ∇
𝜕𝑢 𝜕𝑣 𝜕𝑤
⃗ ⋅ 𝑉⃗ ≡
∇ + +
𝜕𝑥 𝜕𝑦 𝜕𝑧
(A/19)
250
A3.3 Advective
𝜕 𝜕 𝜕
𝑉⃗ ⋅ ∇
⃗≡𝑢 +𝑣 +𝑤
𝜕𝑥 𝜕𝑦 𝜕𝑧
(A/20)
𝜕𝐴 𝜕𝐴 𝜕𝐴
(𝑉⃗ ⋅ ∇
⃗ )𝐴 = 𝑢 +𝑣 +𝑤
𝜕𝑥 𝜕𝑦 𝜕𝑧
(A/21)
⃗
It can also be applied to a vector field 𝐴:
𝜕 𝐴⃗ 𝜕 𝐴⃗ 𝜕 𝐴⃗
(𝑉⃗ ⋅ ∇
⃗ )𝐴⃗ = 𝑢 +𝑣 +𝑤
𝜕𝑥 𝜕𝑦 𝜕𝑧
(A/22)
⎛ 𝑢 𝜕𝑥 + 𝑣 𝜕𝑦 + 𝑤 𝜕𝐴
𝜕𝐴𝑥 𝜕𝐴𝑥 ⎞
⎜ ⎟
𝑥
𝜕𝑧
=⎜ 𝑢 𝜕𝑥𝑦 + 𝑣 𝜕𝑦𝑦 + 𝑤 𝜕𝑧𝑦 ⎟
𝜕𝐴 𝜕𝐴 𝜕𝐴
⎜ 𝑢 𝜕𝐴 + 𝑣 𝜕𝐴 + 𝑤 𝜕𝐴 ⎟
(A/23)
⎝ ⎠
𝑧 𝑧 𝑧
𝜕𝑥 𝜕𝑦 𝜕𝑧
A3.4 Laplacian
⃗2 ≡ ∇
∇ ⃗ ⋅∇
⃗ (A/24)
⃗ 2𝐴 ≡ ∇
∇ ⃗ ⋅∇⃗𝐴
𝜕 𝐴 𝜕 2𝐴 𝜕 2𝐴
(A/25)
2
= + +
(𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2
(A/26)
When applied to a vector field, the general expression uses the curl operator
(we never use this expression in this course), and produces a vector field:
⃗ 2 𝐴⃗ ≡ (∇
∇ ⃗ ) 𝐴⃗ − ∇
⃗ ⋅∇ ⃗ ⃗
⃗× ∇
( × 𝐴) (A/27)
⎛ ⃗ 2 𝐴𝑥
∇ ⎞ ⃗ ⋅∇
⎛ ∇ ⃗ 𝐴𝑥 ⎞
∇𝐴=≡⎜
⃗ 2 ⃗ ⃗ 2 𝐴𝑦
∇ ⎟ = ⎜ ∇ ⃗ ⋅∇ ⃗ 𝐴𝑦 ⎟
⎜ ⎟ ⎜ ⎟
⎝ ⃗ 2 𝐴𝑧
∇ ⎠ ⃗ ⋅∇
⎝ ∇ ⃗ 𝐴𝑧 ⎠
(A/28)
⎛ 𝜕 2 𝐴𝑥
+ (𝜕𝑦) 2 + (𝜕𝑧)2 ⎞
𝜕 2 𝐴𝑥 𝜕 2 𝐴𝑥
⎜ (𝜕𝑥)2
𝜕2𝐴 ⎟
=⎜ 𝜕 2 𝐴𝑦
+ (𝜕𝑦)𝑦2 + (𝜕𝑧)𝑦2 ⎟
𝜕2𝐴
⎜ 𝜕 2 𝐴𝑧 ⎟
(𝜕𝑥)2
+ (𝜕𝑦) 2 + (𝜕𝑧)2
(A/29)
𝜕 2 𝐴𝑧 𝜕 2 𝐴𝑧
⎝ (𝜕𝑥)2 ⎠
251
⃗ 2 𝑉⃗ :
For example, the Laplacian of a velocity field is the vector field ∇
⎛ ⃗ 2𝑢
∇ ⎞ ⎛⎜ + + ⎞
𝜕2𝑢 𝜕2𝑢 𝜕2𝑢
(𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2
⎟
∇𝑉 ≡⎜
⃗ 2 ⃗ ⃗ 2𝑣
∇ ⎟=⎜ 𝜕2𝑣
+ 𝜕2𝑣
+ 𝜕2𝑣
⎟
⎜ ⎟ ⎜ ⎟
⃗ 2𝑤
(𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2
⎝ ∇ ⎠ ⎝ + +
(A/30)
𝜕2𝑤 𝜕2𝑤 𝜕2𝑤
(𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2 ⎠
A3.5 Curl
⃗ × 𝑉⃗ :
For example, the curl of velocity is the vector field ∇
| 𝑖⃗ 𝑗⃗ 𝑘⃗ ||
| 𝜕𝑤 𝜕𝑣 ⃗ 𝜕𝑤 𝜕𝑢 ⃗ 𝜕𝑣 𝜕𝑢 ⃗
⃗ × 𝑉⃗ = || 𝜕
∇ 𝜕 𝜕 |
| = − 𝑖 + − + 𝑗 + − 𝑘
| 𝜕𝑥 𝜕𝑧 | ( 𝜕𝑦 𝜕𝑧 ) ( 𝜕𝑥 𝜕𝑧 ) ( 𝜕𝑥 𝜕𝑦 )
|𝑢 𝑤 ||
𝜕𝑦
| 𝑣
(A/33)
252
A4 Derivations of the Bernoulli equation
d
d𝐹⃗pressure + d𝐹⃗shear + d𝐹⃗gravity = 𝜌 𝑉⃗ d + 𝜌𝑉𝐴 d𝑉⃗
d𝑡 ∭CV
along a streamline, where the velocity 𝑉⃗ is aligned (by definition) with the streamline.
The projection of the net force due to gravity d𝐹⃗gravity on the streamline
segment d𝑠 has norm d𝐹⃗gravity ⋅ d⃗𝑠 = −𝑔𝜌𝐴 d𝑧, while the net force due to
pressure is aligned with the streamline and has norm d𝐹pressure,𝑠 = −𝐴 d𝑝.
Along this streamline, we thus have the following scalar equation, which we
integrate from points 1 to 2:
−𝐴 d𝑝 − 𝜌𝑔𝐴 d𝑧 = 𝜌𝑉𝐴 d𝑉
1
− d𝑝 − 𝑔 d𝑧 = 𝑉 d𝑉
𝜌
2
1 2 2
−∫ d𝑝 − ∫ 𝑔 d𝑧 = ∫ 𝑉 d𝑉
1 𝜌 1 1
(eq. 6/42 p. 125) onto an infinitesimal portion of trajectory d⃗𝑠. Once all terms
We are now going to project every component of the Navier-Stokes equation
𝜕 𝑉⃗
𝜌 + 𝜌(𝑉⃗ ⋅ ∇
⃗ )𝑉⃗ = 𝜌 𝑔⃗ − ∇ ⃗ 2 𝑉⃗
⃗ 𝑝 + 𝜇∇
𝜕𝑡
𝜕 𝑉⃗
𝜌 ⋅ d⃗𝑠 + 𝜌(𝑉⃗ ⋅ ∇
⃗ )𝑉⃗ ⋅ d⃗𝑠 = 𝜌 𝑔⃗ ⋅ d⃗𝑠 − ∇ ⃗ 2 𝑉⃗ ⋅ d⃗𝑠
⃗ 𝑝 ⋅ d⃗𝑠 + 𝜇 ∇
𝜕𝑡
This equation can then be integrated from point 1 to point 2 along the
pathline:
2 2 2
𝜌∫ 𝑉 d𝑉 = − ∫ 𝜌𝑔 d𝑧 − ∫ d𝑝
1 1 1
254
Thus, we can see that if we follow a particle along its path, in a steady,
incompressible, frictionless flow with no heat or work transfer, its change in
kinetic energy is due only to the result of gravity and pressure, in accordance
with the Navier-Stokes equation.
255
A5 Flow parameters as force ratios
This topic is well covered in Massey [6]
mass grows proportionally to the product of its density 𝜌 and its volume 𝐿3 .
Meanwhile, its acceleration relates how much its velocity 𝑉 will change over
a time interval Δ𝑡: it may be expressed as a ratio Δ𝑉 /Δ𝑡. In turn, the time
interval Δ𝑡 may be expressed as the representative length 𝐿 divided by the
velocity 𝑉 , so that the acceleration may be represented as proportional to the
ratio 𝑉 Δ𝑉 /𝐿. Thus we obtain:
𝑉 Δ𝑉
|net force| = |mass × acceleration| ∼ 𝜌𝐿3
𝐿
⃗
|𝐹net | ∼ 𝜌𝐿 𝑉 Δ𝑉
2
to the shear effort and a representative acting surface 𝐿2 . The shear can
We now observe the viscous force acting on a particle: it is proportional
The magnitude of the viscous force can now be compared to the net force:
256
and we recognize the ratio as the Reynolds number (8/12 p. 164). We thus see
that the Reynolds number can be interpreted as the inverse of the influence
of viscosity. The larger [Re] is, and the smaller the influence of the viscous
forces will be on the trajectory of fluid particles.
The magnitude of this force can now be compared to the net force:
and here we recognize the square of the Froude number (8/11 p. 164). We thus
see that the Froude number can be interpreted as the inverse of the influence
of weight on the flow. The larger [Fr] is, and the smaller the influence of
gravity will be on the trajectory of fluid particles.
of the fluid (formally defined as 𝐾 ≡ 𝜌 𝜕𝑝/𝜕𝜌); the elastic force can therefore
be modeled as proportional to 𝐾 𝐿2 :
The magnitude of this force can now be compared to the net force:
because the value of 𝐾 in a given fluid varies considerably not only according
This ratio is known as the Cauchy number; it is not immediately useful
𝐾 |reversible = 𝑐 2 𝜌 (A/38)
|net force| 𝜌𝑉 2 𝑉 2
∼ = 2 = [Ma]
2
|elasticity force|reversible 𝐾 𝑐
(A/39)
257
and here we recognize the square of the Mach number (1/10 p. 16). We thus
see that the Mach number can be interpreted as the influence of elasticity on
the flow. The larger [Ma] is, and the smaller the influence of elastic forces
will be on the trajectory of fluid particles.
258
A6 Details of the winter 2020-2021 final
examination (updated February 2021)
The final examination for this course in the winter semester 2020-2021 is an
open-book, take-home exam which will take place on February 18, 2021
from 14:00 to 16:00. The most important information is as follows:
• This is a take-home exam: you may consult your notes, books, use Video: The companion video to
software and resources offline or online. this exam briefing (2020)
by Olivier Cleynen (CC-by)
https://youtu.be/6vnK7-VsKXc
• You must take this exam alone: you cannot interact with anyone online
or offline during the exam.
• The exam lasts 90 minutes. Additionally, 30 minutes are provided for
the the scanning and upload of your answer.
• You will receive the assignment per email shortly before the start. You
must scan your answer with your student card on each page, and send
it as a single-PDF file to fluidmech@ovgu.de using your academic email
before the end of the exam.
• In the winter semester, this online examination grade is the only grade
you receive in this course.
Three problems will be given, all mandatory. The first problem (10 pts is
exercise 6.2 p.131. The two other problems (45 pts each) are extracted from
the list above, and modified slightly. Typically, the input data is changed, as
well as the problem geometry. The method for solving the problems remains
the same.
A formula sheet is provided. It is the sum of the preambles of every problem
sheet in the lecture notes. It includes the Moody diagram and the viscosity
1 In exercise 4.6, only the calculation of the vertical force 𝐹top is examinable. 259
diagram used in the problem sheets. You should definitely have a calculator
with you, to facilitate calculations.
The criteria for grading your answers are:
Examinations from previous years, and their full solution, are available on
the course website (https://fluidmech.ninja/). Since the course content has
changed over time, you might find a few differences:
• Viscosity values were read in a different diagram, and may not match
values read in the 2020 viscosity diagram.
You are welcome (and in fact encouraged!) to ask me questions of all sorts
about the exam. You may contact me as described in the introduction, page 7.
I wish you to have productive and joyful revisions!
Olivier
February 2021
(updated February 12, after cancellations following covid-19 restrictions)
260
A7 Example of previous examinations
The following pages present the final examinations for this course in 2020
and 2021, and their full solutions.
261
Fluid Mechanics examination — July 11, 2019
Continuity equation for incompressible flow:
⃗ ⋅ 𝑉⃗ = 0
∇
Fluid Mechanics for Master Students
(7)
Duration: 2 h – Use of calculator is authorized; documents are not authorized.
Solve problem 1, plus three other problems among problems 2 to 6.
Navier-Stokes equation for incompressible flow:
D𝑉⃗
𝜌 = 𝜌 𝑔⃗ − ∇ ⃗ 2 𝑉⃗
⃗ 𝑝 + 𝜇∇
D𝑡
Except otherwise indicated, we assume that:
(8)
Fluids are Newtonian
The atmosphere has 𝑝atm. = 1 bar; 𝜌atm. = 1,225 kg m−3 ; 𝑇atm. = 11,3 ◦C; 𝜇atm. = 1,5 ⋅ 10−5 Pa s
Air behaves as a perfect gas: 𝑅air =287 J kg−1 K−1 ; 𝛾air =1,4; 𝑐𝑝 air =1 005 J kg−1 K−1 ; 𝑐𝑣 air =718 J kg−1 K−1
In a highly-viscous (creeping) steady flow, the drag 𝐹𝐷 exerted on a spherical body of
Liquid water is incompressible: 𝜌water = 1 000 kg m−3 , 𝑐𝑝 water = 4 180 J kg−1 K−1 diameter 𝐷 at by flow at velocity 𝑉∞
is quantified as:
𝐹𝐷sphere = 3𝜋𝜇𝑉∞ 𝐷 (9)
In cylindrical pipe flow, we assume the flow is always laminar for [Re]𝐷 . 2 300,
Balance of mass in a considered volume with steady flow:
0 = Σ [𝜌𝑉⟂ 𝐴]incoming + Σ [𝜌𝑉⟂ 𝐴]outgoing (1) and always turbulent for [Re]𝐷 & 4 000. The Darcy friction factor 𝑓 is defined as:
where 𝑉⟂ is negative inwards, positive outwards.
|Δ𝑝loss |
𝑓 ≡ 𝐿1
𝜌𝑉av.
2
(10)
𝐷2
The loss coefficient 𝐾𝐿 is defined as:
Balance of momentum in a considered volume with steady flow:
𝐹⃗net on fluid = Σ [𝜌𝑉⟂ 𝐴𝑉⃗ ]incoming + Σ [𝜌𝑉⟂ 𝐴𝑉⃗ ]outgoing |Δ𝑝loss |
𝐾𝐿 ≡
(2)
where 𝑉⟂ is negative inwards, positive outwards. 1
𝜌𝑉av.
2
(11)
2
Viscosities of various fluids are given in fig. 1. Pressure losses in cylindrical pipes
Balance of energy in a considered volume with steady flow:
𝑝 1
𝑄̇ net + 𝑊̇ shaft, net
can be calculated with the help of the Moody diagram presented in fig. 2 p.4.
= Σ 𝑚̇ 𝑖 + + 𝑉 2 + 𝑔𝑧
[ ( 𝜌 2 )]in
𝑝 1 2
+Σ 𝑚̇ 𝑖 + + 𝑉 + 𝑔𝑧
[ ( 𝜌 2 )]
(3) Non-dimensional incompressible Navier-Stokes equation:
𝜕 𝑉⃗ ∗ 1 1 ∗2 ⃗ ∗
where 𝑚̇ is negative inwards, positive outwards.
+ [1] 𝑉⃗ ∗ ⋅ ∇
⃗ ∗ 𝑉⃗ ∗ = ⃗ ∗𝑝∗ +
𝑔⃗∗ − [Eu] ∇ ⃗ 𝑉
∇
out
𝜕𝑡 ∗ [Fr]2
[St] (12)
[Re]
in which [St] ≡ 𝑓𝑉𝐿 , [Eu] ≡ 𝑝𝜌0 −𝑝
𝑉2
∞
, [Fr] ≡ √𝑉 and [Re] ≡ 𝜌 𝑉𝜇 𝐿 .
𝑔𝐿
The force coefficient 𝐶𝐹 and power coefficient 𝐶P are defined as:
d
Mass balance through an arbitrary volume:
0 = 𝜌 d + ∬ 𝜌 (𝑉⃗rel ⋅ 𝑛)
⃗ d𝐴
𝑊̇
d𝑡 ∭CV 𝐶𝐹 ≡
𝐹
𝐶P ≡
(4)
1
𝜌𝑆𝑉 2 1
𝜌𝑆𝑉 3
CS (13)
2 2
d The speed of sound 𝑐 in air is modeled as:
Momentum balance through an arbitrary volume:
𝐹⃗net = 𝜌 𝑉⃗ d + ∬ 𝜌 𝑉⃗ (𝑉⃗rel ⋅ 𝑛)
⃗ d𝐴 √
d𝑡 ∭CV 𝑐 = 𝛾 𝑅𝑇
(5)
CS (14)
d
Angular momentum balance through an arbitrary volume:
𝑀⃗ net,X = 𝑟⃗X𝑚 ∧ 𝜌 𝑉⃗ d + ∬ 𝑟⃗X𝑚 ∧ 𝜌 (𝑉⃗rel ⋅ 𝑛)
⃗ 𝑉⃗ d𝐴
d𝑡 ∭CV
(6)
CS
1 2
In boundary layer flow, we assume that transition occurs at [Re]𝑥 & 5 ⋅ 105 .
The wall shear coefficient 𝑐𝑓 , a function of distance 𝑥,
is defined based on the free-stream flow velocity 𝑈 :
𝜏wall
𝑐𝑓(𝑥) ≡ 1
𝜌𝑈 2
(15)
2
Exact solutions to the laminar boundary layer along a smooth surface yield:
𝛿 4,91 𝛿∗ 1,72
= √ = √
𝑥 𝑥
(16)
[Re]𝑥 [Re]𝑥
𝛿 ∗∗ 0,664 0,664
= √ 𝑐𝑓(𝑥) = √
𝑥
(17)
[Re]𝑥 [Re]𝑥
Solutions to the turbulent boundary layer along a smooth surface yield the following
time-averaged characteristics:
𝛿 0,16 𝛿∗ 0,02
≈ ≈
𝑥 1
𝑥 1 (18)
[Re]𝑥7 [Re]𝑥7
𝛿 ∗∗ 0,016 0,027
≈ 𝑐𝑓(𝑥) ≈
𝑥 1 1 (19)
[Re]𝑥7 [Re]𝑥7
In a highly-viscous (creeping) steady flow, the drag 𝐹D exerted on a spherical body of
diameter 𝐷 at by flow at velocity 𝑈∞ is quantified as:
𝐹D sphere = 3𝜋𝜇𝑈∞ 𝐷 (20)
Figure 1 – Viscosity of various fluids at a pressure of 1 bar (in practice viscosity is almost independent of
Figure 2 – A Moody diagram, which presents values for 𝑓 measured experimentally, as a function of the
diameter-based Reynolds number [Re]𝐷 , for different relative roughness values.
pressure).
Figure © White 2008
Diagram CC-by-sa S Beck and R Collins, University of Sheffield
3 4
1 Governing equation
1.1. [5 pts] Write out equation (8), the Navier-Stokes equation for incompressible flow,
in its fully-developed form in three Cartesian coordinates.
1.2. [5 pts] Write out equation (7), the continuity equation for incompressible flow, in
its fully-developed form in three Cartesian coordinates.
2 Observation window in a water tank
A water tank used in a laboratory is filled with stationary water (fig. 3). A window is
installed on one of the walls of the canal, to enable observation. The window is hinged
The window has a height of 1,5 m and a width of 3,5 m. The walls of the tank are inclined
on its top face.
Solve problem 1,
and three other problems among problems 2 to 6. with an angle 𝜃 = 70° relative to horizontal.
The following marking guidelines will be used:
• Answers to questions starting with “show that” should be fully-developed and
continuous;
• In all other questions, the correct result with the correct unit is enough to obtain Figure 3 – A door installed on the wall of a water tank.
full points;
• Illegible or ambiguous answers are always discarded.
2.1. [15 pts] What is the magnitude of the net force applying on the tank window?
2.2. [10 pts] At what distance away from the hinge does this force apply?
Water is added to the tank, so that the water level increases.
2.3. [5 pts] How will the distance calculated above change as water is added? (briefly
justify your answer, e.g. in 30 words or less)
5 6
3 Piping leading to a turbine 4 Boundary layer on a flat plate
A thin and smooth plate with width 𝑊 = 0,6 m and length 𝐿 = 2 m is placed with a
zero angle of attack in atmospheric air flow incoming at 21 m s−1 , as shown in figure 5.
A pipe leads water from one reservoir to a turbine, which discharges into another
reservoir, as shown in figure 4.
We would like to study the shear exerted by the flow over the top surface of the plate.
Figure 5 – A thin plate positioned parallel to an incoming uniform flow.
Figure CC-0 o.c.
4.1. [5 pts] At what distance 𝑥tr. along the plate, approximately, will the boundary layer
transit and become turbulent?
4.2. [10 pts] Starting from equation (21), which quantifies the friction factor 𝑐𝑓 (see
definition 15) in a laminar boundary layer,
0,664
Figure 4 – Layout of the water pipe. For clarity, in this figure, the vertical scale is greatly exaggerated. In
𝑐𝑓(𝑥) = √
the vertical scale, the diameter of the pipe is also greatly exaggerated.
(21)
The pipe is made of coarse concrete (roughness 0,25 mm) and carries 800 L s−1 of water
[Re]𝑥
at 20 ◦C. It has a diameter of 1,1 m and features four elbow bends with sharp angles, each show that the shear force 𝐹𝜏 laminar exerted in the laminar section of the boundary
inducing a loss coefficient 𝐾𝐿 of 0,75. layer is:
√
𝐹𝜏 laminar = 0,664 𝜌𝜇 𝑈 2 𝑊 𝑥tr.2
1
3.1. [10 pts] Represent qualitatively (i.e. without numerical data) the pressure distribu-
3
(22)
4.3. [5 pts] What is the shear force exerted on the top surface of the plate by the laminar
tion along the length of the pipe, both when the turbine is shut down (without any
flow), and when it is operating.
3.2. [15 pts] What is the hydraulic power available to the turbine?
section of the boundary layer?
4.4. [5 pts] What is the shear force exerted on the top surface of the plate by the
The outlet tank on the right is very large, so that its water level does not vary. The source turbulent section of the boundary layer?
4.5. [5 pts] Would the boundary layer become thicker if the velocity was increased?
water tank on the left, however, sees its height decrease as the water is emptied through
the turbine. Ultimately, as the water level decreases, the water stops flowing entirely.
(briefly justify your answer, e.g. in 30 words or less).
3.3. [5 pts] When the water stops flowing, what will be the height of the water level in
the source tank on the left?
7 8
The width of the profile (perpendicular to the flow, in the 𝑧-direction) is 70 cm. The water
has uniform temperature and density (20 ◦C, 999 kg m−3 ) and the pressure is uniform
5 Velocity measurements in a tunnel
across the measurement surface.
A group of students proceeds with speed measurements in a water tunnel. The objective
5.1. [20 pts] What is the drag force applying on the profile?
is to measure the drag applying on a an object with constant cross-section, positioned
across the tunnel test section (fig. 6).
5.2. [10 pts] If water was replaced with a fluid with higher viscosity, how would you
expect the drag force to change? (briefly justify your answer, e.g. in 30 words or
less)
6 Lift and drag on a rotating football
A group of fluid dynamicists investigates the air flow around a football. In particular,
flying through the air. The football has diameter 22 cm, a weight of 430 g; it is traveling
they are interested in the forces applying on the ball when it has been kicked and is
at 70 km h−1 .
completely across the tunnel (in the 𝑧-direction). The horizontal velocity distributions upstream and
Figure 6 – An object with constant cross-section positioned across a water tunnel. The object spans
sphere has a diameter of 1,1 m. Drag force measurements are carried out in the tunnel.
In order to observe the flow, they install a steel sphere in a wind tunnel (figure 7). The
downstream of the profile are also shown.
Figure CC-0 o.c.
Upstream of the object, the water flow velocity is uniform (𝑢1 = 𝑈 = 3,2 m s−1 ).
Downstream of the object, horizontal velocity measurements are made every 5 cm across
the flow; the following results are obtained:
vertical position 𝑦 (cm) horizontal speed 𝑢2 (m s−1 )
0 3,2
5 3,2
10 3,15
15 3,14
20 3,03
25 2,92
30 2,81
35 2,87
40 2,89 travels with speed 𝑉tunnel . Force measurements are carried out on the ball.
Figure 7 – A steel sphere positioned in a wind tunnel. The sphere is maintained stationary, while the air
45 2,97
50 3,19
Figure CC-0 o.c.
55 3,2
60 3,2
9 10
6.1. [5 pts] What is the wind tunnel speed required, so that the flow around the real
football is reproduced around the sphere in the tunnel?
6.2. [5 pts] With the speed calculated above, by which factor should the drag force
measured in the wind tunnel be multiplied, in order to obtain the drag force on the
real football?
The fluid dynamicists now investigate the effect of spin on the ball. When the football
is rotated along a horizontal axis during travel, a lift force exerts laterally on the ball,
curving its trajectory. This is represented, from above, in figure 8.
In order to quantify this effect, the wind tunnel sphere is rotated in the wind tunnel, and
measurements are carried out; the results are plotted in figure 9.
Figure 9 – Experimental measurements of the lift and drag coefficients applying on a rotating sphere in an
steady uniform flow.
Figure © from Munson & al. 2013
6.3. [10 pts] How many rotations per second are required in order to generate a lift
force of 3,1 N on the real football when it travels?
6.4. [5 pts] What is then the corresponding drag force ?
6.5. [5 pts] Propose and quantify one possibility for the football player to double the
lift force applying on the ball.
Figure 8 – Trajectory of a rotating football in free flight, as seen from above. A lift force exerts towards the
left, and deviates the trajectory towards the left.
Figure CC-0 o.c.
11 12
Solution: Fluid Mechanics examination — July 11, 2019 2 Observation window in a water tank
Fluid Mechanics for Master Students
fluidmech.ninja 2.1 Net force
We define coordinates 𝑟 and 𝑧, and the length 𝐿1 , as shown in the figure below.
1 Governing equation
1.1 N-S equation question
𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑝 𝜕 2𝑢 𝜕 2𝑢 𝜕 2𝑢
𝜌 +𝑢 +𝑣 +𝑤 = 𝜌𝑔𝑥 − +𝜇 + +
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 ] 𝜕𝑥 [ (𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2 ]
𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝑝 𝜕 2𝑣 𝜕 2𝑣 𝜕 2𝑣
(1)
𝜌 +𝑢 +𝑣 +𝑤 = 𝜌𝑔𝑦 − +𝜇 + +
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 ] 𝜕𝑦 [ (𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2 ]
𝜕𝑤 𝜕𝑤 𝜕𝑤 𝜕𝑤 𝜕𝑝 𝜕 2𝑤 𝜕 2𝑤 𝜕 2𝑤
(2)
𝜌 +𝑢 +𝑣 +𝑤 = 𝜌𝑔𝑧 − +𝜇 + +
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 ] 𝜕𝑧 [ (𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2 ]
(3)
The force on a small section of door with length d𝑟 and width 𝑊 is d𝐹 . On the complete
door, the force applying due to the net pressure 𝑝net of water and air is:
1.2 Continuity question
𝜕𝑢 𝜕𝑣 𝜕𝑤 𝑟=𝑅max
+ + = 0 𝐹net = ∫ d𝐹
𝜕𝑥 𝜕𝑦 𝜕𝑧
(4) (5)
𝑟=0
𝑟=𝑅max
= ∫ 𝑝net d𝑆 (6)
𝑟=0
𝑟=𝑅max
= ∫ 𝑝net 𝑊 d𝑟 (7)
𝑟=0
𝑟=𝑅max
= ∫ 𝜌 𝑔 𝑧 𝑊 d𝑟 (8)
𝑟=0
A coordinate transform is needed to solve the integral, expressing 𝑧 as a function of 𝑅.
This is obtained by geometry:
𝑧 = 𝑍min + 𝑟 sin 𝜃 (9)
1 2
Inserting eq. 9 into eq. 8, we continue with: 2.3 Change in distance
𝑟=𝑅max
𝐹net = ∫ 𝜌𝑔 (𝑍min + 𝑟 sin 𝜃) 𝑊 d𝑟 Increasing water height translates (only) in an increase of the value of 𝐿1 in the equations
above. Both 𝑀net and 𝐹net increase together with 𝐿1 (eqs. 14 & 22), so it is not immediately
(10)
𝑟=0
𝑟=𝑅max
= 𝜌𝑔𝑊∫ (𝑍min + 𝑟 sin 𝜃) d𝑟 (11) apparent how 𝑅𝐹 changes in eq. 26.
𝑟=0
𝑟=𝑅max
= 𝜌𝑔𝑊∫ (𝐿1 sin 𝜃 + 𝑟 sin 𝜃) d𝑟 (12) Several possibilities can be used to find the answer:
𝑟=0
1 𝑟=𝑅max
= 𝜌 𝑔 𝑊 sin 𝜃 [𝐿1 𝑟 + 𝑟 2 ] • Substituting 𝑥𝐿1 instead of 𝐿1 and comparing the two radiuses, one can write:
2 𝑟=0
(13)
= 𝜌 𝑔 𝑊 sin 𝜃 (𝐿1 𝑅max + 0,5𝑅max 2 )
𝑅𝐹 2 < 𝑅𝐹 1
(14)
= 103 × 9,81 × 3,5 × sin(70°) (0,8 × 1,5 + 0,5 × 1,52 )
(29)
(15)
= 75 014 N 1
𝑥𝐿1 𝑅max
2
+ 1 𝑅3 1
𝐿1 𝑅max + 3 𝑅max
2 1 3
2 3 max
< 2
(16)
𝐹net = 75,01 kN 𝑥𝐿1 𝑅max + 21 𝑅max
2 𝐿1 𝑅max + 21 𝑅max
2
(30)
(17)
1 < 𝑥 (31)
2.2 Distance from hinge
(it is even possible to show, using this equation 30, that 𝑅𝐹 tends towards 𝑅max /2 as
𝐿1 increases);
We first calculate the moment exerting about the hinge due to the net pressure of air and
water, using the same notation as above:
𝑟=𝑅max
𝑀net = ∫ 𝑟 d𝐹
• It is possible to observe graphically that the center of application of the force
(18)
𝑟=0
increase in 𝐿1 ;
moves closer to the hinge when the net pressure distribution is changed due to the
𝑟=𝑅max
= ∫ 𝑟𝜌 𝑔 𝑧 𝑊 d𝑟 (19)
𝑟=0
𝑟=𝑅max
= ∫ 𝑟𝜌 𝑔 (𝑍min + 𝑟 sin 𝜃) 𝑊 d𝑟 (20)
𝑟=0
𝑟=𝑅max
= 𝜌𝑔𝑊∫ 𝑟 (𝐿1 sin 𝜃 + 𝑟 sin 𝜃) d𝑟 (21)
𝑟=0
1 1 𝑟=𝑅max
= 𝜌 𝑔 𝑊 sin 𝜃 [ 𝐿1 𝑟 2 + 𝑟 3 ]
2 3 𝑟=0
1
(22)
= 103 × 9,81 × 3,5 × sin(70°) (0,5 × 0,8 × 1,52 + × 1,53 )
3 • It is also possible to calculate manually one or several new values for 𝑅𝐹 ;
(23)
= 65 335 N m (24)
𝑀net = 65,34 kN m (25) All those methods will provide some evidence that the distance 𝑅𝐹 will in fact decrease
when 𝐿1 is increased.
The distance away from the hinge 𝑅𝐹 is obtained by dividing the moment by the force:
𝑀net
𝑅 =
𝐹net
𝐹 (26)
75 014
=
65 335
(27)
𝑅𝐹 = 0,87 m (28)
3 4
The Reynolds number is
3 Piping leading to a turbine
𝜌𝑉av. 𝐷
[Re]𝐷 =
𝜇
(36)
103 × 0,842 × 1,1
=
3.1 Pressure distribution
10−5
(37)
[Re]𝐷 = 9,251 ⋅ 105 (38)
The relative roughness is
𝜖 0,25 ⋅ 10−3
=
𝐷 1,1
𝜖
(39)
= 2,27 ⋅ 10−4
𝐷
(40)
With those values, the Moody diagram reads:
𝑓 = 0,0158 (41)
Finally, the wall friction losses along the pipe are calculated as:
1 2 𝐿
Δ𝑝𝑓 = −𝑓 𝜌𝑉av.
2 𝐷
(42)
1 4 ⋅ 103
= −0,0158 103 0,8422
2 1,1
(43)
3.2 Turbine power Δ𝑝𝑓 = −2,03 ⋅ 104 Pa (44)
We want to calculate three pressure drops: • Pressure drop due to losses in the four bends, Δ𝑝bends :
• Pressure drop due to wall friction losses along the pipe, Δ𝑝𝑓 : The average velocity 1 2
Δ𝑝bends = −4 × 𝐾𝐿 𝜌𝑉av.
2
1
(45)
= −4 × 0,75 × × 103 × 0,8422
2
in the pipe is
̇
(46)
𝑉 = Δ𝑝bends = −1,06 ⋅ 103 Pa
𝑆
(47)
̇
av. (32)
= 𝐷2 • Pressure drop due to hydrostatic pressure change across the turbine, Δ𝑝ℎ :
𝜋 4
(33)
0,8
=
𝜋 × 1,14
2
Δ𝑝ℎ = 𝜌𝑔(Δ𝑧)
(34)
𝑉av. = 0,842 m s−1
(48)
(35) = 103 × 9,81 [4 − (25 + 51)] (49)
Δ𝑝ℎ = −7,06 ⋅ 105 Pa (50)
5 6
Finally, the turbine hydraulic power is obtained as:
4 Boundary layer on a flat plate
̇
𝑊̇ turbine = (Δ𝑝turbine )
= ̇ (Δ𝑝ℎ − Δ𝑝𝑓 − Δ𝑝bends )
(51)
(52) 4.1 Transition point
= 0,8 × [−7,06 ⋅ 105 − (−2,03 ⋅ 104 ) − (−1,06 ⋅ 10−3 )] (53) The transition point occurs at [Re]𝑥 ≈ 5 ⋅ 105 . Solving for 𝑥tr. , we have:
= −5,479 ⋅ 105 W
𝜌𝑈 𝑥tr.
𝑊̇ turbine = −547,9 kW [Re]𝑥tr. =
(54)
𝜇
(56)
[Re]𝑥tr. 𝜇
(55)
𝑥tr. =
𝜌𝑈
(57)
5 ⋅ 105 × 1,5 ⋅ 10−5
=
3.3 Residual water height
1,225 × 21
(58)
𝑥tr. = 0,29 m
residual water height in the left tank will be 8 m.
The water will flow until air is entrained (“sucked”) into the pipe inlet. At this point, the (59)
4.2 Shear in laminar section
We start with the given equation and implement the definition (15) of the formula sheet,
as well as the definition of the distance-based Reynolds number [Re]𝑥 :
0,664
𝑐𝑓(𝑥) = √ (60)
[Re]𝑥
𝜏wall, laminar 0,664
= √
1
𝜌𝑈 2 𝜌𝑈 𝑥
(61)
2 𝜇
𝜌𝑈 𝑥 2 1 2
−1
𝜏wall, laminar = 0,664 𝜌𝑈
( 𝜇 ) 2
√
(62)
𝜏wall, laminar = 0,332 𝜌𝜇 𝑈 1,5 𝑥 − 2
1
(63)
The shear force is the integral of the shear with respect to area:
𝐹shear, laminar = ∫ 𝜏wall, laminar d𝑆 (64)
We split the total area covered by the laminar boundary layer in strips of width 𝑊 and
length d𝑥, with 𝑥 ranging from 0 (leading edge) to 𝑥tr. (where the laminar part of the
7 8
boundary ends), obtaining: 4.5 Thickness
𝑥=𝑥tr.
𝐹shear, laminar = ∫ 𝜏wall 𝑊 d𝑥 (65)
and 18 for 𝛿 (one for the laminar section, the other for the turbulent section). In both, the
𝑥=0
Models for the boundary layer thickness are given in the formula sheet as equations 16
𝑥=𝑥tr.
√
= ∫ 0,332 𝜌𝜇 𝑈 1,5 𝑥 − 2 𝑊 d𝑥
1
(66)
𝑥=0 Reynolds number [Re]𝑥 appears in the denominator (the lower part of the fraction). As
√ 𝑥tr.
𝑈 is increased, [Re]𝑥 will increase too, and consequently, the thickness of the boundary
= 0,332 𝜌𝜇 𝑈 1,5 𝑊 ∫ 𝑥 −0,5 d𝑥 (67)
0
√ 1
𝑥
= 0,332 𝜌𝜇 𝑈 1,5 𝑊 𝑥 −0,5+1
layer will decrease.
[ −0,5 + 1 ]0
tr.
√
(68)
= 0,332 𝜌𝜇 𝑈 1,5 𝑊 2 𝑥tr.0,5
√
(69)
𝐹𝜏 laminar = 0,664 𝜌𝜇 𝑈 2 𝑊 𝑥tr.2
3 1
(70)
4.3 Value of shear in the laminar section
We simply insert values into eq. 70:
√
𝐹𝜏 laminar = 0,664 1,225 × 1,5 ⋅ 10−5 21 2 × 0,6 × 0,292 2
3 1
(71)
𝐹𝜏 laminar = 0,0887 N (72)
4.4 Shear in turbulent section
The process is the same in the turbulent part of the layer, with 𝜏wall, turbulent derived from
equation (19) in the formula sheet:
1 𝜌𝑈
− 71
𝜏wall, turbulent = 0,027 𝜌𝑈 2 𝑥−7
( 𝜇 )
1
2
(73)
This is integrated with respect to area, with 𝑥 ranging from 𝑥tr. (where the turbulent
section begins) to 𝑥max (the trailing edge of the plate):
𝑥=𝑥max
𝐹shear, turbulent = ∫ 𝜏wall 𝑊 d𝑥 (74)
𝑥=𝑥tr.
𝑥max
= 0,0135 𝜌 7 𝑈 7 𝜇 7 𝑊 ∫ 𝑥 − 7 d𝑥
6 13 1 1
(75)
𝑥tr.
𝑥max
= 0,01575 𝜌 7 𝑈 7 𝜇 7 𝑊 [𝑥 7 ]𝑥tr.
6 13 1 6
(76)
= 0,01575 × 1,225 7 × 21 7 × (1,5 ⋅ 10−5 ) 7 × 0,6 × (2 7 − 0,292 7 )
6 13 1 6 6
(77)
𝐹shear, turbulent = 0,961 N (78)
9 10
We insert the expression for ℎ1 obtained above in this last expression, continuing as:
5 Velocity measurements in a tunnel
1 ℎ2 ℎ2
−𝐹net = −𝜌𝐿 𝑢2(𝑦) d𝑦 + 𝜌𝐿 ∫ 𝑢2(𝑦) d𝑦
𝑈 ∫0
2
(87)
0
5.1 Drag force ℎ2
= 𝜌𝐿 ∫ (𝑢2(𝑦) − 𝑈1 𝑢2(𝑦) ) d𝑦
2
(88)
0
We build a control volume around the object:
Instead of a function 𝑢2 = 𝑓 (𝑦), we have discrete values. The integral is therefore
approximated as:
−𝐹net = 𝜌𝐿 ∑ (𝑢22 − 𝑈1 𝑢2 ) δ𝑦 (89)
𝑦
= 𝜌𝐿 ∑ [𝑢2 (𝑢2 − 𝑈1 )] δ𝑦 (90)
𝑦
−𝐹net = 𝜌𝐿 ∑ [3,2 (3,2 − 3,2)
𝑦
+3,2 (3,2 − 3,2)
+3,15 (3,15 − 3,2)
+3,14 (3,14 − 3,2)
+3,03 (3,03 − 3,2)
We use a mass balance equation to quantify the height ℎ1 of the inlet:
+2,92 (2,92 − 3,2)
d
0 = 𝜌 d + ∬ 𝜌 (𝑉⃗rel ⋅ 𝑛)
⃗ d𝐴 +2,81 (2,81 − 3,2)
d𝑡 ∭CV
(79)
+2,87 (2,87 − 3,2)
= − ∬ 𝜌|𝑉in | d𝐴 + ∬ 𝜌|𝑉out | d𝐴
CS
+2,89 (2,89 − 3,2)
(80)
ℎ1 ℎ2
= −𝜌𝐿 ∫ 𝑈 d𝑦 + 𝜌𝐿 ∫ 𝑢2(𝑦) d𝑦 (81) +2,97 (2,97 − 3,2)
0 0
+3,19 (3,19 − 3,2)
ℎ2
= −𝜌𝐿𝑈 ℎ1 + 𝜌𝐿 ∫ 𝑢2(𝑦) d𝑦
+3,2 (3,2 − 3,2)
(82)
0
1 ℎ2
ℎ1 = 𝑢2(𝑦) d𝑦
𝑈 ∫0 +3,2 (3,2 − 3,2) ] δ𝑦
(83)
(91)
= 999 × 0,7 × (−5,3325) × 0,05 (92)
scalar equation in the 𝑥-direction: = −186,45 N
The drag force is quantified using a momentum balance equation, which reduces to a
(93)
𝐹net = 186,45 N
d
𝐹⃗net = 𝜌 𝑉⃗ d + ∬ 𝜌 𝑉⃗ (𝑉⃗rel ⋅ 𝑛)
⃗ d𝐴
d𝑡 ∭CV
(94)
(84)
−𝐹net = − ∬ 𝜌|𝑉in |2 d𝐴 + ∬ 𝜌|𝑉out |2 d𝐴 𝐹net is the net force exerted on the fluid by the object. It is positive in the 𝑥-direction.
CS
(85)
ℎ2
= −𝜌𝐿ℎ1 𝑈12 + 𝜌𝐿 ∫ 𝑢2(𝑦) d𝑦 opposite direction (flow-wise direction): 𝐹drag = −186,45 N.
The drag force is the force exerted on the object on the fluid, and so is pointing in the
2
(86)
0
11 12
5.2 Dependence on viscosity 6 Lift and drag on a rotating football
Viscosity does not appear in equation (90) above. Nevertheless, an increase in viscosity
will translate into higher shear, and so it is likely that the object will affect a larger 6.1 Required wind tunnel speed
of 𝑢2 in the tabled measurement values). The expression for 𝐹drag will not change, but its
amount of fluid around itself. This will result in a larger velocity deficit (reduced values
The two flows will have identical behavior if the Reynolds numbers are equal. With 1
denoting the real football, and 2 denoting the wind tunnel sphere, we have:
value will increase.
[Re]1 = [Re]2
𝜌𝑉1 𝐷1 𝜌𝑉2 𝐷2
(95)
=
𝜇 𝜇
(96)
𝐷1
𝑉2 = 𝑉1
𝐷2
(97)
70 0,22
=
3,6 1,1
(98)
= 3,89 m s−1 (99)
𝑉2 = 14 km h−1 (100)
6.2 Ratio of forces
Since the two flows are dynamically similar, the force coefficients are the same. Consid-
ering the lift coefficients,
𝐶𝐿1 = 𝐶𝐿2
𝐿 𝐿2
(101)
= 1
1
1
𝜌𝑆 𝑉 2
𝜌𝑆 2 𝑉1
2
(102)
2 1 1 2
𝐿1 𝐿2
=
𝐷12 𝑉12 𝐷22 𝑉12
(103)
𝐿1 𝐷2 𝑉 2
= 1 1
𝐿2 𝐷22 𝑉22
(104)
𝑉2 𝑉2
= 22 12
𝑉1 𝑉2
(105)
𝐿1
= 1
𝐿2
(106)
So, the forces will be identical on both the wind tunnel model and the real football.
13 14
6.3 Rotation speed
The desired lift force is 𝐿1 = 3,1 N. This corresponds to a lift coefficient of:
𝐿1
𝐶𝐿1 = 1
𝜌𝑆1 𝑉12
(107)
2
𝐿1
=
1
𝜌𝜋 𝐷41 𝑉12
2 (108)
2
3,1
=
0,5 × 1,225 × 𝜋 × 0,222
× 19,442
(109)
4
𝐶𝐿1 = 0,352 (110)
Inputting this value in figure 9, one corresponding value of 𝜔𝐷/2𝑈 is 1,51. This allows
us to obtain a value for 𝜔 (other higher values also work):
𝜔1 𝐷1
= 1,51
2𝑈1
(111)
1,51 × 2𝑈1
𝜔1 =
𝐷1
(112)
1,51 × 2 × 19,44
=
0,22
(113)
= 266 rad s−1 (114)
𝜔1 = 42,5 rotations/s (115)
6.4 Drag force
The chosen value of 𝜔𝐷/2𝑈 corresponds to a drag coefficient reading of 0,56 in figure 9.
Inputting this in the definition for the drag coefficient, we can solve for the drag 𝐹𝐷 :
𝐹𝐷1
0,56 = 𝐶𝐷1 = 1
𝜌𝑆 1 𝑉1
2
(116)
2
1 𝐷2
𝐹𝐷1 = 0,56 × 𝜌𝜋 1 𝑉 2
2 4 1
(117)
0,222
= 0,56 × 0,5 × 1,225 × 𝜋 × 19,442
4
(118)
𝐹𝐷1 = 4,93 N (119)
6.5 Doubling of lift force
√ √
The force is “easily” doubled by multiplying the speed 𝑉1 by a factor 2: 𝑉3 = 2𝑉1 .
We obtain the same lift coefficient (𝐶𝐿3 = 𝐶𝐿1 = 0,352). The rotation speed has to be
adapted according to the expression 112: we obtain 𝜔3 = 60,1 rotations/s.
15
Fluid dynamics examination — September 21, 2020 Shear force on a flat solid surface:
𝐹shear, direction 𝑖 = ∬ 𝜏direction 𝑖 d𝑆 (7)
Fluid Dynamics for Engineers by Olivier Cleynen 𝑆
Shear in the direction 𝑗, on a plane perpendicular to direction 𝑖:
Solve problem 1, plus three other problems among problems 2 to 6.
Duration: 2 h – Use of calculator is authorized; documents are not authorized.
𝜕𝑉𝑗
||⃗
𝜏𝑖𝑗 || = 𝜇
𝜕𝑖
(8)
Except otherwise indicated, assume that:
The atmosphere has 𝑝atm. = 1 bar; 𝜌atm. = 1,225 kg m−3 ; 𝑇atm. = 11,3 ◦C; 𝜇atm. = 1,5 ⋅ 10−5 Pa s
Air behaves as a perfect gas: 𝑅air =287 J kg−1 K−1 ; 𝛾air =1,4; 𝑐𝑝 air =1 005 J kg−1 K−1 ; 𝑐𝑣 air =718 J kg−1 K−1
Continuity equation for incompressible flow:
Liquid water is incompressible: 𝜌water = 1 000 kg m−3 , 𝑐𝑝 water = 4 180 J kg−1 K−1
⃗ ⋅ 𝑉⃗ = 0
∇ (9)
Navier-Stokes equation for incompressible flow:
D𝑉⃗
𝜌 = 𝜌 𝑔⃗ − ∇ ⃗ 2 𝑉⃗
⃗ 𝑝 + 𝜇∇
D𝑡
Balance of mass in a fixed control volume with steady flow:
0 = Σ [𝜌𝑉⟂ 𝐴]incoming + Σ [𝜌𝑉⟂ 𝐴]outgoing
(10)
(1)
where 𝑉⟂ is negative inwards, positive outwards.
In a highly-viscous (creeping) steady flow, the drag 𝐹𝐷 exerted on a spherical body of
diameter 𝐷 at by flow at velocity 𝑉∞ is quantified as:
Balance of momentum in a fixed control volume with steady flow:
𝐹⃗net on fluid = Σ [𝜌𝑉⟂ 𝐴𝑉⃗ ] + Σ [𝜌𝑉⟂ 𝐴𝑉⃗ ]
𝐹𝐷sphere = 3𝜋𝜇𝑉∞ 𝐷
(2)
(11)
where 𝑉⟂ is negative inwards, positive outwards.
incoming outgoing
Balance of energy in a fixed control volume with steady flow: In cylindrical pipe flow, we assume the flow is always laminar for [Re]𝐷 . 2 300,
𝑝 1 and always turbulent for [Re]𝐷 & 4 000. The Darcy friction factor 𝑓 is defined as:
𝑄̇ net + 𝑊̇ shaft, net = Σ 𝑚̇ 𝑖 + + 𝑉 2 + 𝑔𝑧
[ ( 𝜌 2 )]in |Δ𝑝loss |
𝑓 ≡
𝑝 1 𝐿1
𝜌𝑉av.
+Σ 𝑚̇ 𝑖 + + 𝑉 2 + 𝑔𝑧
2
(12)
[ ( 𝜌 )]out
𝐷2
2
The loss coefficient 𝐾𝐿 is defined as:
(3)
where 𝑚̇ is negative inwards, positive outwards.
|Δ𝑝loss |
𝐾𝐿 ≡ 1
𝜌𝑉av.
2
(13)
2
Viscosities of various fluids are given in fig. 1 p. 4. Pressure losses in cylindrical pipes
Mass balance through an arbitrary volume:
d
can be calculated with the help of the Moody diagram presented in fig. 2 p.5.
0 = 𝜌 d + ∬ 𝜌 (𝑉⃗rel ⋅ 𝑛)
⃗ d𝐴
d𝑡 ∭CV
(4)
CS
Momentum balance through an arbitrary volume: The non-dimensional incompressible Navier-Stokes equation:
d 𝜕 𝑉⃗ ∗ 1 1 ∗2 ⃗ ∗
𝐹⃗net = 𝜌 𝑉⃗ d + ∬ 𝜌 𝑉⃗ (𝑉⃗rel ⋅ 𝑛)
⃗ d𝐴 + [1] 𝑉⃗ ∗ ⋅ ∇
⃗ ∗ 𝑉⃗ ∗ = ⃗ ∗𝑝∗ + ⃗ 𝑉
d𝑡 ∭CV 𝜕𝑡 ∗
𝑔⃗∗ − [Eu] ∇ ∇
(5)
[Fr]2
[St] (14)
CS [Re]
in which [St] ≡ 𝑓𝑉𝐿 , [Eu] ≡ 𝑝𝜌0 −𝑝
𝑉2
∞
, [Fr] ≡ √𝑉 and [Re] ≡ 𝜌 𝑉𝜇 𝐿 .
𝑔𝐿
Angular momentum balance through an arbitrary volume:
d
𝑀⃗ net,X = 𝑟⃗X𝑚 ∧ 𝜌 𝑉⃗ d + ∬ 𝑟⃗X𝑚 ∧ 𝜌 (𝑉⃗rel ⋅ 𝑛)
⃗ 𝑉⃗ d𝐴
d𝑡 ∭CV
(6)
CS
1 2
The force coefficient 𝐶𝐹 and power coefficient 𝐶P are defined as:
𝐹 𝑊̇
𝐶𝐹 ≡ 𝐶P ≡
1
𝜌𝑆𝑉 2 1
𝜌𝑆𝑉 3
(15)
2 2
The speed of sound 𝑐 in air is modeled as:
√
𝑐 = 𝛾 𝑅𝑇 (16)
2×10−2 2.4×10−5
In boundary layer flow, we assume that transition occurs at [Re]𝑥 ≈ 5 ⋅ 105 .
The wall shear coefficient 𝑐𝑓 , a function of distance 𝑥, is defined using the free-stream
flow velocity 𝑈 : 10−2 2.2×10−5
9×10−3
𝜏wall 8×10−3
𝑐𝑓(𝑥) ≡ 7×10−3 ⟵ Crude Oil
1
𝜌𝑈 2 6×10−3 Air ⟶
(17)
5×10−3
2
2×10−5
Exact solutions to the laminar boundary layer along a smooth surface yield: 4×10−3
𝛿 4,91 𝛿∗ 1,72 3×10−3
Viscosity 𝜇 of liquids in Pa s
Viscosity 𝜇 of gases in Pa s
= √ = √
𝑥 𝑥 1.8×10−5
2×10−3
(18)
[Re]𝑥 [Re]𝑥
𝛿 ∗∗ 0,664 0,664
= √ 𝑐𝑓(𝑥) = √ CO2 ⟶
𝑥
(19)
[Re]𝑥 [Re]𝑥
10−3 1.6×10−5
9×10−4
8×10−4
7×10−4
Solutions to the turbulent boundary layer along a smooth surface yield the following
time-averaged characteristics: 6×10−4
𝛿 0,16 𝛿∗ 0,02 5×10−4
≈ ≈ 1.4×10−5
𝑥 1
𝑥 1 (20) 4×10−4 ⟵ Water
3×10−4
[Re]𝑥7 [Re]𝑥7
𝛿 ∗∗ 0,016 0,027
≈ 𝑐𝑓(𝑥) ≈
𝑥 2×10−4 1.2×10−5
1 1 (21)
[Re]𝑥7 [Re]𝑥7
10−4 10−5
−20 0 20 40 60 80 100 120
Temperature 𝑇 in degree Celsius (◦C)
Figure 1 – The viscosity of four fluids (crude oil, water, air, and C02) as a function of temperature. The scale
for liquids is logarithmic and displayed on the left; the scale for gases is linear and displayed on the right.
Figure CC-by by Arjun Neyyathala & Olivier Cleynen
3 4
Solve problem 1,
and three other problems among problems 2 to 6.
The following marking guidelines will be used:
• Answers to questions starting with “show that” should be fully-developed and
continuous;
• In all other questions, the correct result with the correct unit is enough to obtain
full points;
• Illegible or ambiguous answers are always discarded.
Figure 2 – A Moody diagram, which presents values for 𝑓 measured experimentally, as a function of the
diameter-based Reynolds number [Re]𝐷 , for different relative roughness values.
Diagram CC-by-sa S Beck and R Collins, University of Sheffield
5 6
1 Navier-Stokes equation 3 Governing equations
1.1. [5 pts] Write out equation (10), the Navier-Stokes equation for incompressible flow, We consider a two-dimensional fluid flow described with the following velocity field,
in its fully-developed form in three Cartesian coordinates. described in Cartesian coordinates 𝑥 and 𝑦 in m s−1 :
1.2. [5 pts] In which flow conditions does this equation apply? 𝑉⃗ = (2𝐴𝑥 − 𝐶)𝑖⃗ + (−2𝐴𝑦 − 𝐵𝑥)𝑗⃗
where 𝐴, 𝐵, and 𝐶 are all constants.
3.1. [5 pts] Show that this flow satisfies the continuity equation for incompressible flow
(equation 9).
3.2. [10 pts] What is the acceleration field corresponding to this flow?
3.3. [5 pts] What is the value of acceleration at a point of coordinates (3; 3)?
2 Pressure forces on the panels of a barge
3.4. [10 pts] Does a function exist to describe the pressure field of this flow, and if so,
A large barge is being built with the dimensions shown in figure 3. Once completed, it
will be floated in a lake. The bottom panel of the barge will then sit horizontally, 2 m
what is it?
below the surface of the water.
Figure 3 – Basic layout of a barge
2.1. [15 pts] What is the magnitude of the force resulting from pressure efforts on each
of the panels labeled A, B and C?
2.2. [5 pts] How would the force on panel C change if the angle 𝜃 was increased? (briefly
justify your answer, e.g. in 30 words or less)
2.3. [10 pts] What is the weight of the barge?
7 8
4 Pipe installation with a pump-turbine 5 Friction on a flat plate
Engineers in a chemical company would like to pump 50 L s−1 of water at 20 ◦C from A laboratory is developing a special oil with Newtonian fluid characteristics. An experi-
reservoir A to reservoir B in the installation drawn below in figure 4. The network is ment is set up to measure its viscosity. For this, a thin layer of oil is poured at the bottom
built with a cylindrical pipe made out of concrete (surface roughness 𝜖 = 0,25 mm), with of a tank with rectangular walls. A flat plate is moved perfectly horizontally, sitting flush
a diameter of 12 cm. The Y-junction induces a loss coefficient of 0,2. with the surface of the oil. A force measurement is carried out.
The plate has width 𝑊 = 40 cm and length 𝐿 = 60 cm. It is moved at 15 cm s−1 at the
surface of the oil, which is at a distance 𝐷 = 4 mm above the bottom of the tank.
Figure 4 – Basic layout of the piping network
The installation lengths are as follows:
𝐿1 = 15 m 𝐿5 = 16 m Figure 5 – Flat plate moved horizontally at the surface of a layer of oil, at the bottom of a tank
𝐿2 = 12 m 𝐿6 = 6 m
𝐿3 = 8 m 𝐿7 = 5 m
𝐿4 = 14 m
The flow below the plate is smooth, steady, and laminar, so that the velocity distribution
in the oil between the plate and the bottom of the tank is entirely uniform.
4.1. [10 pts] What is the hydraulic power that the pump must provide, in order to deliver
the required volume flow? 5.1. [10 pts] What is the relationship between the viscosity of the oil and the drag force
4.2. [10 pts] On a diagram, represent qualitatively (i.e. without numerical data) the
due to shear on the bottom side of the plate?
pressure distribution along the length of the pipe, indicating both the top and the The magnitude of the drag force resulting from the shear exerted by the oil is measured
bottom paths followed by the water. as 𝐹drag = 0,47 N.
4.3. [10 pts] If the water was to be transferred back from reservoir B to reservoir A, the
5.2. [5 pts] What is the viscosity of the oil?
pump/turbine device would be operated as a turbine. In that case, what would be
the hydraulic power available to the turbine?
9 10
On the top surface of the plate, in the air, there is enough space for a boundary layer to
6 Pickup truck with snow plow
develop. The scientists in the laboratory would like to check that the drag force generated
A pickup truck is equipped with a snow plow blade. It travels steadily at 30 km h−1 in
by this boundary layer does not influence the measurement. The air in the room has
properties 1 bar and 20 ◦C.
20 cm of snow with density 𝜌snow ground = 400 kg m−3 , with the blade angled at 𝛼 = 60°.
5.3. [10 pts] Starting with equation 19 for the friction coefficient in a laminar boundary
density 𝜌snow compacted = 600 kg m−3 at velocity 𝑉⃗snow/truck relative to the truck. The cross-
The snow is deflected along the blade, compacted by the movement, and it exits with
sectional area of the rejected snow is 𝐴outlet = 0,55 m2 , measured in a plane perpendicular
layer:
0,664 to 𝑉⃗snow/truck .
𝑐𝑓(𝑥) = √ (19)
[Re]𝑥
show that the relationship between the shear force on the top side of the plate and
the properties of the air is:
𝑝air 𝜇air 𝐿 2 32
1
𝐹top = 0,664𝑊 𝑉
( 𝑅𝑇air ) plate
(22)
5.4. [5 pts] What is the magnitude of the force exerted by the air on the top surface of
the plate?
Figure 6 – A pickup truck using a large blade to clear snow
Pickup drawing CC-0 by en:Wikipedia User:Wikideas1; diagram CC-0 Olivier Cleynen
6.1. [15 pts] What is the force exerted on the blade by the snow?
6.2. [5 pts] What is the power required for the truck to plow the snow?
6.3. [10 pts] How would the power change if the angle 𝛼 was reduced? (briefly justify
your answer, e.g. in 30 words or less)
11 12
Solution: Fluid Dynamics examination — Sept. 21, 2020 2 Pressure forces on the panels of a barge
Fluid Dynamics for Engineers by Olivier Cleynen
https://fluidmech.ninja/ 2.1 Magnitude of the forces
We define coordinates 𝑟, 𝑟𝐶 , and 𝑧, as well as the lengths 𝐿𝐴 , 𝐿𝐵 and 𝐿𝐶 , as shown in the
1 Governing equation figure below.
1.1 N-S equation question
𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑢 𝜕𝑝 𝜕 2𝑢 𝜕 2𝑢 𝜕 2𝑢
𝜌 +𝑢 +𝑣 +𝑤 = 𝜌𝑔𝑥 − +𝜇 + +
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 ] 𝜕𝑥 [ (𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2 ]
𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝑝 𝜕 2𝑣 𝜕 2𝑣 𝜕 2𝑣
(1)
𝜌 +𝑢 +𝑣 +𝑤 = 𝜌𝑔𝑦 − +𝜇 + +
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 ] 𝜕𝑦 [ (𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2 ]
𝜕𝑤 𝜕𝑤 𝜕𝑤 𝜕𝑤 𝜕𝑝 𝜕 2𝑤 𝜕 2𝑤 𝜕 2𝑤
(2)
𝜌 +𝑢 +𝑣 +𝑤 = 𝜌𝑔𝑧 − +𝜇 + +
[ 𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑧 ] 𝜕𝑧 [ (𝜕𝑥)2 (𝜕𝑦)2 (𝜕𝑧)2 ]
(3)
1.2 Conditions for this equation On each wall, the net pressure applying is that due to water and air from both sides,
This equation applies to all incompressible flows of a Newtonian fluid. 𝑝net = 𝑝water + 𝑝atm. − 𝑝atm. (4)
= 𝜌𝑔𝑧 (5)
Wall A
The force on a small horizontal strip of panel with height d𝑟 and width 𝐿A is d𝐹A . On the
complete door, the force applying due to the net pressure 𝑝net of water and air is:
𝑟=𝑅max
𝐹netA = ∫ d𝐹A (6)
𝑟=0
𝑟=𝑅max
= ∫ 𝑝net d𝑆 (7)
𝑟=0
𝑟=𝑅max
= ∫ 𝑝net 𝐿A d𝑟 (8)
𝑟=0
𝑟=𝑅max
= ∫ 𝜌 𝑔 𝑧 𝐿A d𝑟 (9)
𝑟=0
1 2
A coordinate transform is needed to solve the integral, expressing depth 𝑧 as a function Wall C
of 𝑟. This is obtained by geometry: 𝐹net C is found with a similar calculation. This time, the relationship between 𝑟C and 𝑧 is
𝑧 = 𝑍max −𝑟
a little more complicated, and found by trigonometry:
(10)
𝑧 = 𝑍max − 𝑟C sin 𝜃 (26)
Inserting eq. 10 into eq. 9, we continue with:
𝑟=𝑅max
𝐹net A = ∫ 𝜌 𝑔 (𝑍max − 𝑟) 𝐿A d𝑟
Picking up eq. 9 applied to panel C and inserting eqs. 10 and 26, we get:
(11)
𝑟=0
𝑟C =𝑅Cmax
𝐹netC = ∫ 𝜌 𝑔 (𝑍max − 𝑟C sin 𝜃) 𝐿C d𝑟C
𝑟=𝑅max
= 𝜌 𝑔 𝐿A ∫ (𝑍max − 𝑟) d𝑟 (27)
𝑟C =0
(12)
𝑟=0
1 𝑟=𝑅max
𝑟C =𝑅Cmax
= 𝜌 𝑔 𝐿A [𝑍max 𝑟 − 𝑟 2 ] = 𝜌 𝑔 𝐿C ∫ (𝑍max − 𝑟C sin 𝜃) d𝑟C
2 𝑟=0
(13) (28)
𝑟C =0
= 𝜌 𝑔 𝐿A (𝑍max 𝑅max − 0,5𝑅max 2 ) 1 𝑟C =𝑅Cmax
= 𝜌 𝑔 𝐿C 𝑍max 𝑟C − sin 𝜃𝑟C2 ]
[ 2
(14)
= 103 × 9,81 × 20 × (2 × 2 − 0,5 × 22 )
(29)
𝑟C =0
1
= 𝜌 𝑔 𝐿C 𝑍max 𝑅Cmax − sin 𝜃𝑅 2 )
(15)
= 3,924 ⋅ 105 N ( 2
(30)
𝑍max 1 𝑍2
(16) Cmax
𝐹net A = 392,4 kN = 𝜌𝑔𝐿 𝑍 − sin 𝜃 max 2
( max sin 𝜃 2 (sin 𝜃) )
(17) (31)
1 1
C
= 𝜌 𝑔 𝐿C 𝑍2 − 𝑍2
sin 𝜃 ( max 2 max )
(32)
1 1 2
= 𝜌 𝑔 𝐿C 𝑍
Wall B
𝐹net B 𝐿B is a function of 𝑟, which can be sin 𝜃 2 max
(33)
1
is found with a similar calculation. This time,
= 103 × 9,81 × 6 × × 0,5 × 22
sin 30°
found by trigonometry: (34)
𝑟 = 2,3544 ⋅ 105 N
𝐿B =
(35)
tan 𝜃 𝐹net C = 235,4 kN
(18)
(36)
Picking up eq. 9 applied to panel B, and inserting eqs. 10 and 18, we get:
𝑟
2.2 Change in force with angle
𝑟=𝑅max
𝐹net B = ∫ 𝜌 𝑔 (𝑍max − 𝑟) d𝑟
tan 𝜃 If the angle 𝜃 was increased, the term 1/ sin 𝜃 in eq. 33 would decrease, and so the
(19)
𝑟=0
1 𝑟=𝑅max
= 𝜌𝑔 (𝑍max 𝑟 − 𝑟 2 ) d𝑟
tan 𝜃 ∫
(20) calculated force would decrease.
𝑟=0
1 1 1 𝑟=𝑅max
= 𝜌𝑔 𝑍 𝑟2 − 𝑟3
tan 𝜃 [ 2 3 ]𝑟=0
However, the barge would also sink in more deeply (because the immersed volume, which
(21)
1 1 1
max
is responsible for buoyancy, would have to remain constant). This would result in an
= 𝜌𝑔 𝑍max 𝑅max − 𝑅max 3 ) increase in 𝑍max , likely smaller than the decrease caused by the term 1/ sin 𝜃.
tan 𝜃 ( 2
2 2
3
(22)
1 23
= 103 × 9,81 × × 0,5 × 2 × 22 −
tan 30° ( 3)
(23)
= 2,266 ⋅ 104 N
2.3 Weight of the barge
(24)
𝐹net B = 22,66 kN (25) The weight of the barge is equal to the net force due to pressure on all panels. This can
be calculated either one of two ways:
• Using force components. The weight is equal to the force on the bottom panel,
𝐹D = 𝜌𝑔𝑍max 𝐿C 𝐿A , plus the vertical component of the force on panel C, 𝐹Cvertical =
3 4
𝐹netC cos 𝜃. The sum is:
3 Governing equations
𝐹weight = 𝐹D + 𝐹Cvertical (37)
= 𝜌𝑔𝑍max 𝐿C 𝐿A + 𝐹netC cos 𝜃 (38) 3.1 Continuity equation
= 2,558 ⋅ 106 N (39)
𝐹weight = 2,558 MN
The incompressible continuity equation (eq. 9 in the formula sheet) translates in two
(40) dimensions as:
𝜕𝑢 𝜕𝑣
+ = 0
𝜕𝑥 𝜕𝑦
• Using the immersed volume and multiplying it by the density of water and gravity (47)
to obtain the buoyancy:
𝐹weight = immersed 𝜌𝑔
In this case, we have
1 𝜕𝑢 𝜕𝑣 𝜕(2𝐴𝑥 − 𝐶) 𝜕(−2𝐴𝑦 − 𝐵𝑥)
(41)
= 𝜌𝑔 (𝐿A + 𝐿Bmax ) 𝐿C 𝑍max + = +
2 𝜕𝑥 𝜕𝑦 𝜕𝑥 𝜕𝑦
(42) (48)
1 𝑍max = 2𝐴 − 2𝐴
= 𝜌𝑔 𝐿A + 𝐿C 𝑍max
( 2 tan 𝜃 )
(49)
= 0 s−1
(43)
1 2
(50)
= 103 × 9,81 × (20 + ×6×2
2 tan 30° )
(44)
= 2,558 ⋅ 106 N (45) So the incompressible continuity equation is satisfied, and the flow is incompressible.
𝐹weight = 2,558 MN (46)
3.2 Acceleration field
The acceleration field 𝑎⃗ (in m s−2 ) is:
The mass of the barge is approximately 260 tons.
D𝑉⃗
𝑎⃗ =
D𝑡
(51)
𝜕 𝑉⃗
= + (𝑉⃗ ⋅ ∇
⃗ )𝑉⃗
𝜕𝑡
(52)
The 𝑥-component of 𝑎⃗ is
𝜕𝑢 𝜕𝑢 𝜕𝑢
𝑎𝑥 = +𝑢 +𝑣
𝜕𝑡 𝜕𝑥 𝜕𝑦
(53)
𝜕(2𝐴𝑥 − 𝐶) 𝜕(2𝐴𝑥 − 𝐶) 𝜕(2𝐴𝑥 − 𝐶)
= + (2𝐴𝑥 − 𝐶) + (−2𝐴𝑦 − 𝐵𝑥)
𝜕𝑡 𝜕𝑥 𝜕𝑦
(54)
= 0 + 2𝐴(2𝐴𝑥 − 𝐶) + 0 (55)
𝑎𝑥 = 4𝐴2 𝑥 − 2𝐴𝐶 (56)
5 6
The 𝑦-component of 𝑎⃗ is Likewise in the 𝑦-direction, we have
𝜕𝑣 𝜕𝑣 𝜕𝑣 𝜕𝑝 𝜕 2𝑣
𝑎𝑦 = +𝑢 +𝑣 𝜌𝑎𝑦 = 𝜌𝑔𝑦 − +𝜇
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑦 (𝜕𝑦)2
(57) (70)
𝜕(−2𝐴𝑦 − 𝐵𝑥) 𝜕(−2𝐴𝑦 − 𝐵𝑥) 𝜕(−2𝐴𝑦 − 𝐵𝑥) 𝜕𝑝 𝜕 2 (−2𝐴𝑦 − 𝐵𝑥)
= + (2𝐴𝑥 − 𝐶) + (−2𝐴𝑦 − 𝐵𝑥) 𝜌 (4𝐴2 𝑦 + 𝐵𝐶 ) = 𝜌𝑔𝑦 − +𝜇
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜕𝑦 (𝜕𝑦)2
(58) (71)
= 0 − 𝐵(2𝐴𝑥 − 𝐶) − 2𝐴(−2𝐴𝑦 − 𝐵𝑥) 𝜕𝑝
𝜌 (4𝐴2 𝑦 + 𝐵𝐶 ) = 𝜌𝑔𝑦 − +0
𝜕𝑦
(59)
= −2𝐴𝐵𝑥 + 𝐵𝐶 + 4𝐴2 𝑦 + 2𝐴𝐵𝑥
(72)
𝜕𝑝
= 𝜌 (−4𝐴2 𝑦 − 𝐵𝐶 + 𝑔𝑦 )
(60)
𝑎𝑦 = 4𝐴2 𝑦 + 𝐵𝐶 (61) 𝜕𝑦
(73)
So, the acceleration field is, in m s−2 : We have:
𝜕 ( 𝜕𝑥 ) 𝜕 𝜕𝑦
(
𝜕𝑝
)
4𝐴2 𝑥 − 2𝐴𝐶
𝜕𝑝
𝑎⃗ = = 0 =
( 4𝐴2 𝑦 + 𝐵𝐶 ) 𝜕𝑦 𝜕𝑥
(62) (74)
which guarantees that there exists a function to describe 𝑝.
3.3 Acceleration at one point We proceed to integrate eq. 69 with respect to 𝑦:
At point with 𝑥 = 3 and 𝑦 = 3, we have 𝑝 = 𝜌 [−2𝐴2 𝑥 2 + (2𝐴𝐶 + 𝑔𝑥 ) 𝑥 ] + 𝑓(𝑦,𝑡) (75)
12𝐴2 − 2𝐴𝐶 with 𝑓 a function of 𝑦 and 𝑡 only.
𝑎⃗point = m s−2
( 12𝐴2 + 𝐵𝐶 )
(63)
To obtain function 𝑓 , we derivate this equation 75 with respect to 𝑦, and compare with
eq. 73:
𝜕𝑝 d (𝜌 [−2𝐴2 𝑥 2 + (2𝐴𝐶 + 𝑔𝑥 ) 𝑥 ])
3.4 Function for pressure
= + 𝑓(𝑦,𝑡)
′
= 𝜌 (−4𝐴2 𝑦 − 𝐵𝐶 + 𝑔𝑦 )
𝜕𝑦 d𝑦
(76)
0 + 𝑓(𝑦,𝑡) = 𝜌 (−4𝐴2 𝑦 − 𝐵𝐶 + 𝑔𝑦 )
We use the incompressible Navier-Stokes equation (eq. 10 of the equation sheet) to express
′
the acceleration field as a function of the gravity, pressure and shear terms: (77)
D𝑉⃗ We can now integrate this equation 77 with respect to 𝑦 to find the function 𝑓 :
𝜌 = 𝜌 𝑔⃗ − ∇ ⃗ 2 𝑉⃗
⃗ 𝑝 + 𝜇∇
D𝑡
(64)
⃗ 2 𝑉⃗
⃗ 𝑝 + 𝜇∇
𝜌 𝑎⃗ = 𝜌 𝑔⃗ − ∇ (65) 𝑓(𝑦,𝑡) = 𝜌 [−2𝐴2 𝑦 2 + (−𝐵𝐶 + 𝑔𝑦 ) 𝑦 ] + 𝑖(𝑡) + 𝑝0 (78)
with 𝑖 a function of 𝑡 only,
Splitting that last equation into two components, we get, for the 𝑥-component: and 𝑝0 an arbitrary integration constant.
𝜕𝑝 𝜕 2𝑢 Inserting this equation 78 into eq. 75, we finally get the function for pressure (in Pa)
𝜌𝑎𝑥 = 𝜌𝑔𝑥 − +𝜇
𝜕𝑥 (𝜕𝑥)2
(66)
𝜕𝑝 𝜕 2 (2𝐴𝑥 − 𝐶)
corresponding to the given velocity field:
𝜌 (4𝐴2 𝑥 − 2𝐴𝐶 ) = 𝜌𝑔𝑥 − +𝜇
𝜕𝑥 (𝜕𝑥)2 𝑝 = 𝜌 [−2𝐴2 𝑥 2 + (2𝐴𝐶 + 𝑔𝑥 ) 𝑥 ] + 𝜌 [−2𝐴2 𝑦 2 + (−𝐵𝐶 + 𝑔𝑦 ) 𝑦 ] + 𝑖(𝑡) + 𝑝0
(67)
𝜕𝑝
𝜌 (4𝐴2 𝑥 − 2𝐴𝐶 ) = 𝜌𝑔𝑥 − +0
(79)
𝜕𝑥
(68) 𝑝 = 𝜌 [−2𝐴2 𝑥 2 − 2𝐴2 𝑦 2 + (2𝐴𝐶 + 𝑔𝑥 ) 𝑥 + (−𝐵𝐶 + 𝑔𝑦 ) 𝑦 ] + 𝑖(𝑡) + 𝑝0
𝜕𝑝
(80)
= 𝜌 (−4𝐴2 𝑥 + 2𝐴𝐶 + 𝑔𝑥 )
𝜕𝑥
(69)
All functions 𝑖 of time, and any initial value 𝑝0 , may be inserted in equation 80, and it
will still satisfy the incompressible Navier-Stokes equation.
7 8
• Pressure drop due to wall friction losses in each of the two branches, Δ𝑝𝑓 2 : The
4 Pipe installation with a pump-turbine
average velocity in the pipe is
𝑉av.1
𝑉av.2 =
2
4.1 Hydraulic power of pump (95)
𝑉av.2 = 2,21 m s−1 (96)
We want to calculate four pressure differences:
• Pressure drop due to wall friction losses along the horizontal pipe, Δ𝑝𝑓 1 : The The Reynolds number is
[Re]𝐷 2 =
average velocity in the pipe is
[Re]𝐷 1
̇ 2
(97)
𝑉av.1 = [Re]𝐷 2 = 2,653 ⋅ 105
𝑆
(81)
̇
(98)
= 𝐷2
𝜋 4
(82)
0,05
With the same relative roughness, the Moody diagram reads:
=
𝜋 × 0,12 𝑓2 = 0,025
2 (83)
4
𝑉av.1 = 4,42 m s−1
(99)
(84)
Finally, the wall friction losses along the pipe are calculated as:
1 2 𝐿2
The Reynolds number is
𝜌𝑉av.1 𝐷 Δ𝑝𝑓 2 = −𝑓2 𝜌𝑉av.2
[Re]𝐷 1 = 2 𝐷
(100)
𝜇 12
= −0,025 × 0,5 × 103 × 2,212 ×
(85)
103 × 4,42 × 0,12 0,12
=
(101)
10−3 Δ𝑝𝑓 2 = −6,108 ⋅ 103 Pa
(86)
[Re]𝐷 1 = 5,305 ⋅ 105
(102)
= −0,06 bar
(87)
(103)
• Pressure drop due to losses in the junction, Δ𝑝junction :
The relative roughness is
𝜖 0,25 ⋅ 10−3 1 2
= Δ𝑝junction = 𝐾𝐿 𝜌𝑉av.1
𝐷 0,12 2
(88)
𝜖
(104)
= 2,08 ⋅ 10−3 1
𝐷 = 0,2 × × 103 × 4,4212
2
(89)
(105)
Δ𝑝junction = −1,954 ⋅ 103 Pa (106)
= −0,02 bar
With those values, the Moody diagram reads:
(107)
𝑓1 = 0,0241
• Pressure drop due to hydrostatic pressure change across the pump, Δ𝑝ℎ :
(90)
Δ𝑝ℎ = 𝜌𝑔(𝐿5 + 𝐿6 − 𝐿3 )
Finally, the wall friction losses along the pipe are calculated as:
1 2 𝐿1
(108)
Δ𝑝𝑓 1 = −𝑓1 𝜌𝑉av. = 103 × 9,81 × (16 + 6 − 8)
2 𝐷
(91) (109)
15 Δ𝑝ℎ = +1,373 ⋅ 105 Pa
= −0,0241 × 0,5 × 103 × 4,422 ×
0,12
(110)
= +1,37 bar
(92)
Δ𝑝𝑓 1 = −2,944 ⋅ 104 Pa
(111)
(93)
= −0,29 bar (94)
9 10
Finally, the pump hydraulic power is obtained as:
5 Friction on a flat plate
̇
𝑊̇ pump = (Δ𝑝pump ) (112)
= ̇ (Δ𝑝ℎ − Δ𝑝𝑓 1 − Δ𝑝𝑓 2 − Δ𝑝junction )
The flow can be represented as shown below:
(113)
= 0,05 × [+1,373 ⋅ 105 − (−2,944 ⋅ 104 ) − (−6,108 ⋅ 103 ) − (−1,954 ⋅ 103 )] (114)
= 0,05 × [+1,373 + 0,2944 + 0,0611 + 0,01954] × 105 (115)
= +8,742 ⋅ 103 W (116)
𝑊̇ pump = +8,74 kW (117)
4.2 Pressure distribution
5.1 Shear force on bottom side
In order to calculate shear on the bottom face, we express the horizontal fluid velocity
below the plate (relative to the plate) in a very simple velocity field:
𝑉𝑥 = 𝑢 = 0 + 𝑘𝑦 (124)
The constant 𝑘 is found using boundary conditions: at a distance 𝐷 away from the plate,
4.3 Hydraulic power of turbine the velocity of the fluid relative to the plate is 𝑉plate :
𝑢|@ y = D = 𝑉plate = 0 + 𝑘𝐷
𝑉plate
If the flow runs from B to A, the magnitude of the pressure differences remains the same. (125)
𝑘 =
𝐷
The hydrostatic pressure difference changes sign, but the pressure losses due to friction (126)
do not:
̇
𝑊̇ turbine = (Δ𝑝turbine ) (118) So the velocity field becomes:
= ̇ (Δ𝑝ℎturbine − Δ𝑝𝑓 1 − Δ𝑝𝑓 2 − Δ𝑝junction ) 1
𝑢 = 𝑉plate 𝑦
(119)
= ̇ (−Δ𝑝ℎpump − Δ𝑝𝑓 1 − Δ𝑝𝑓 2 − Δ𝑝junction ) 𝐷
(127)
(120)
= 0,05 × [−1,373 + 0,2944 + 0,0611 + 0,01954] × 105 (121)
= −4,992 ⋅ 103 W (122)
𝑊̇ turbine = −4,99 kW (123)
So, pumping the flow from A to B costs 8,7 kW of power, but turbining the flow back
down from B to A only generates 5 kW of power.
11 12
Now, the shear on the bottom plate is found through integration, starting with equation 7 reading the value 𝜇air = 1,85 ⋅ 10−5 Pa s in figure 1, we have:
𝜌 𝑈𝑥
[Re]𝑥tr. = air tr.
from the formula sheet:
𝜇air
(138)
𝐹bottom = ∫ d𝐹 = ∬ 𝜏plate d𝑆 [Re]𝑥tr. 𝜇air
𝑥tr. =
(128)
𝜌air 𝑈
𝑆
= ∬ 𝜏plate d𝑆
(139)
𝑝air [Re]𝑥tr. 𝜇air
=
(129)
𝑥=𝐿
𝑅𝑇air 𝑈
= ∫ 𝜏plate 𝑊 d𝑥
(140)
1 ⋅ 105 × 5 ⋅ 105 × 1,85 ⋅ 10−5
=
(130)
𝑥=0
287 × (273,15 + 20) × 0,15
(141)
Expressing the shear 𝜏plate as a function of the velocity field in a Newtonian fluid, we 𝑥tr. = 73,3 m (142)
continue to get:
Since 𝑥tr. > 𝐿, the boundary layer never transits and remains completely laminar.
𝑥=𝐿
𝜕𝑢 ||
𝐹bottom = ∫ 𝜇oil 𝑊 d𝑥
𝜕𝑦 ||@𝑦=0 To find an expression for 𝜏 , we start with the given equation and implement the defi-
(131)
𝑥=0
( 𝐷 plate 𝑦 ) ||
𝑥=𝐿 𝜕 1 𝑉
= 𝜇oil 𝑊 ∫ | d𝑥
𝜕𝑦 |
nition 15 of the formula sheet, as well as the definition of the distance-based Reynolds
|@𝑦=0
(132)
𝑥=0 number [Re]𝑥 :
𝑥=𝐿
1
= 𝜇oil 𝑊 ∫ ( 𝐷 𝑉plate )@𝑦=0 d𝑥 0,664
𝑐𝑓(𝑥) = √
(133)
𝑥=0
1
(143)
𝐹bottom = 𝜇oil 𝑊 𝑉plate 𝐿
[Re]𝑥
𝐷 𝜏 laminar 0,664
= √
(134)
1
𝜌 𝑈2
wall,
𝜌air 𝑈 𝑥
(144)
2 air
Rearranging this equation to solve for 𝜇, we get:
𝜇air
𝜌air 𝑈 𝑥 2 1
−1
𝐷 𝜏 = 0,664 𝜌 𝑈2
( 𝜇air ) 2 air
𝜇oil = 𝐹bottom
(145)
𝑊 𝑉plate 𝐿 √
wall, laminar
𝜏wall, laminar = 0,332 𝜌air 𝜇air 𝑈 1,5 𝑥 − 2
(135) 1
(146)
5.2 Viscosity of oil The shear force on the top surface is the integral of the shear with respect to area:
The viscosity of the oil is obtained by plugging in values in eq. 135: 𝐹top = ∫ 𝜏wall d𝑆 (147)
4 ⋅ 10−3 We split the total area covered by the laminar boundary layer in strips of width 𝑊 and
𝜇oil = 0,47 ×
0,4 × 0,15 × 0,6
length d𝑥, with 𝑥 ranging from 0 (leading edge) to 𝐿 (trailing edge), obtaining:
(136)
𝜇oil = 5,22 ⋅ 10−2 Pa s (137)
𝑥=𝐿
𝐹top = ∫ 𝜏wall 𝑊 d𝑥 (148)
𝑥=0
𝑥=𝐿
√
5.3 Shear force on top side
= ∫ 0,332 𝜌air 𝜇air 𝑈 1,5 𝑥 − 2 𝑊 d𝑥
1
(149)
𝑥=0
√ 𝐿
= 0,332 𝜌air 𝜇air 𝑈 1,5 𝑊 ∫ 𝑥 −0,5 d𝑥
A boundary layer develops on the top side, in air. That boundary layer would transit at
[Re]𝑥 ≈ 5 ⋅ 105 . Solving for 𝑥tr. with 𝑈 (the faraway velocity) being equal to 𝑉plate , and (150)
0
√ 1
𝐿
= 0,332 𝜌air 𝜇air 𝑈 1,5 𝑊 𝑥 −0,5+1
[ −0,5 + 1 ]0
(151)
√
= 0,332 𝜌air 𝜇air 𝑈 1,5 𝑊 2 𝐿0,5 (152)
13 14
√
= 0,664 𝜌air 𝜇air 𝑈 2 𝑊 𝐿 2
3 1
(153)
6 Pickup truck with snow plow
= 0,664𝑊 (𝜌air 𝜇air 𝐿) 2 𝑈 2
1 3
(154)
𝑝air 𝜇air 𝐿 2 23
1
= 0,664𝑊 𝑈
( 𝑅𝑇air )
enters with velocity 𝑉⃗1 and leaves with velocity 𝑉⃗2 = 𝑉⃗out .
(155) We define a control volume moving together with the truck, as shown below. The snow
𝑝air 𝜇air 𝐿 2 23
1
𝐹top = 0,664𝑊 𝑉
( 𝑅𝑇air ) plate
(156)
5.4 Force exerted by the air
We simply insert values into eq. 156:
1 ⋅ 105 × 1,85 ⋅ 10−5 × 0,6
0,5
𝐹top = 0,664 × 0,4 × × 0,151,5
( 287 × (273,15 + 20) )
(157)
𝐹top = 5,6 ⋅ 10−5 N (158)
This force is is very much smaller than the force generated by the oil, so the air’s influence
in the viscosity measurement experiment is indeed very small.
6.1 Force on blade
̇ as well as for the two vectors 𝑉⃗1 and 𝑉⃗2 .
We want expressions for the mass flow 𝑚,
The mass flow is found using the inlet conditions:
𝑚̇ = 𝜌1 𝑉1 𝐴⟂1 (159)
= 𝜌𝑉𝑊
1 1 ℎsnow
30
(160)
= 400 × × 3,5 × 0,2
sweep
3,6
(161)
𝑚̇ = 2,333 ⋅ 103 kg s−1 (162)
The outlet velocity can then be obtained:
𝑚̇ = 𝜌2 𝑉2 𝐴⟂2 (163)
𝑚̇ = 𝜌2 𝑉2 𝐴out
𝑚̇
(164)
𝑉2 =
𝜌2 𝐴out
(165)
2,333 ⋅ 103
=
600 × 0,55
(166)
𝑉2 = 7,07 m s−1 (167)
15 16
The velocity vector 𝑉⃗2 is then expressed according to its two components: ⎛ +1,429 ⋅ 104 ⎞
⎜ ⎟
= ⎜ 0 ⎟N
𝑉2𝑥 = +𝑉2 sin 𝛼 ⎜ ⎟
(181)
(168)
⎝ −1,12 ⋅ 104 ⎠
= 7,07 × sin(60°) ⎛ +14,3 ⎞
⎜ ⎟
(169)
𝑉2𝑥 = +6,12 m s−1 𝐹⃗net = ⎜ 0 ⎟ kN
⎜ ⎟
(170) (182)
⎝ −11,2 ⎠
𝑉2𝑧 = +𝑉2 cos 𝛼 (171)
= 7,07 × cos(60°)
The force exerted on the blade by the snow is exactly the opposite, i.e.
⎛ −14,3 ⎞
(172)
𝑉2𝑧 = +3,54 m s−1 ⎜ ⎟
𝐹⃗snow on blade = −𝐹⃗net = ⎜ 0 ⎟ kN
(173)
⎜ ⎟
(183)
So, we have all the necessary information about the flow: ⎝ +11,2 ⎠
𝑚̇ = 2,333 ⋅ 103 kg s−1 (174) 6.2 Power required of truck
⎛ 0 ⎞
⎜ ⎟ The power 𝑊̇ truck provided by the truck to compensate for the force is
𝑉⃗1 = ⎜ 0 ⎟ m s−1
⎜ ⎟
(175)
⎝ +8,33 ⎠
𝑊̇ truck = 𝐹⃗net ⋅ 𝑉⃗truck on ground
⎛ +6,12 ⎞
(184)
⎜ ⎟ = −𝑉truck 𝐹net 𝑧
𝑉⃗2 = ⎜ 0 ⎟ m s−1
(185)
⎜ ⎟ = −8,333 × (−1,12 ⋅ 104 )
(176)
⎝ +3,54 ⎠
(186)
= 9,329 ⋅ 104 W (187)
Now, applying eq. 2 from the equation sheet, we can express the net force 𝐹⃗net on the 𝑊̇ truck = 93,3 kW (188)
snow:
𝐹⃗net = Σ [𝜌𝑉⟂ 𝐴𝑉⃗ ] + Σ [𝜌𝑉⟂ 𝐴𝑉⃗ ]
(approximately 120 horsepower)
(177)
= 𝑚̇ (𝑉⃗2 − 𝑉⃗1 )
incoming outgoing
(178) 6.3 Power change with angle
⎛ 𝑉2𝑥 − 𝑉1𝑥 ⎞
⎜ ⎟
= 𝑚̇ ⎜ 𝑉2𝑦 − 𝑉1𝑦 ⎟
The power is quantified using equation 185, in which we include eq. 179, eq. 171, and
⎜ ⎟
(179)
⎝ 𝑉2𝑧 − 𝑉1𝑧 ⎠
then eq. 160:
⎛ (+6,12) − 0 ⎞ ̇
𝑊truck = −𝑉truck 𝐹net 𝑧
⎜ ⎟
(189)
= 2,333 ⋅ 103 ⎜ 0−0 ⎟ = −𝑉truck 𝑚̇ (𝑉2𝑧 − 𝑉1𝑧 )
⎜ ⎟
(180)
⎝ (+3,54) − (+8,33)
(190)
⎠ = −𝑉truck 𝑚̇ (𝑉2 cos 𝛼 − 𝑉1𝑧 )
𝑚̇
(191)
= −𝑉truck 𝑚̇ cos 𝛼 − 𝑉truck
( 𝜌2 𝐴out )
(192)
17 18
𝜌 𝑉 𝑊sweep ℎsnow
= −𝑉truck 𝑚̇ cos 𝛼 − 𝑉truck
1 1
( 𝜌2 𝐴out )
(193)
𝜌1 𝑉truck 𝑊sweep ℎsnow
= −𝑉truck 𝜌𝑉truck ℎsnow 𝑊sweep cos 𝛼 − 𝑉truck
( 𝜌2 𝐴out )
(194)
𝜌1 𝑊sweep ℎsnow
= −𝑉truck
3
ℎsnow 𝑊sweep cos 𝛼 − 1
( 𝜌2 𝐴out )
(195)
In equation 195, decreasing 𝛼 would:
• increase the term cos 𝛼;
• decrease the effective width 𝑊sweep (unless a new, larger blade is used).
Both factors tend to decrease the magnitude of 𝑊̇ truck . Therefore, the power would
decrease, likely faster than 𝛼.
19
A8 References
These notes are based on textbooks by White [22], Çengel & al.[25], Munson & al.[29], and de Nevers [17].
[1] Ira Herbert Abbott and Albert Edward Von Doenhoff. Theory of wing
sections, including a summary of airfoil data. Courier, 1959.
[2] George Keith Batchelor. An introduction to fluid dynamics. Cambridge
University Press, 1967. isbn: 9780521663960.
[3] Geoffrey Ingram Taylor. “Film notes for low-Reynolds-number flows”.
In: National Committee for Fluid Mechanics Films. Illustrated experi-
ments in fluid mechanics: the NCFMF book of film notes. MIT Press, 1967.
isbn: 9780262640121. url: http://web.mit.edu/hml/ncfmf/07LRNF.pdf.
[4] John Leask Lumley. “Film notes for Eulerian and Lagrangian descrip-
tions in fluid mechanics”. In: National Committee for Fluid Mechanics
Films. Illustrated experiments in fluid mechanics: the NCFMF book of
film notes. MIT Press, 1969. isbn: 9780262640121. url: http://web.mit.
edu/hml/ncfmf/01ELDFM.pdf.
[5] Hendrik Tennekes and John Leask Lumley. A first course in turbulence.
MIT press, 1972. isbn: 978-0262200196.
[6] Bernard Stanford Massey. Mechanics of fluids. 5th ed. Van Nostrand
Reinhold, 1983. isbn: 0442305524.
[7] Phillip M. Gerhart and Richard J. Gross. Fundamentals of fluid dynamics.
1st ed. Addison-Wesley, 1985. isbn: 0201114100.
[8] Hendrik Tennekes. The simple science of flight. From insects to jumbo
jets. 1st ed. MIT Press, 1992. isbn: 0262201054.
[9] John David Anderson. Computational fluid dynamics. McGraw-Hill,
1995. isbn: 0071132104.
[10] Paul A. Libby. An introduction to turbulence. CRC Press, 1996. isbn:
9781560321002.
[11] Jewel B. Barlow, William H. Jr. Rae, and Alan Pope. Low-speed wind
tunnel testing. 3rd ed. Wiley & Sons, 1999. isbn: 0471557749.
[12] Jean Mathieu and Julian Scott. An introduction to turbulent flow. Cam-
bridge University Press, 2000. isbn: 0521775388.
[13] Stephen B. Pope. Turbulent flows. Cambridge University Press, 2000.
isbn: 9780521177849.
[14] Peter S. Bernard and James M. Wallace. Turbulent flow: analysis, mea-
surement, and prediction. John Wiley & Sons, 2002. isbn: 9780471332190.
[15] John Bird. Higher engineering mathematics. 4th ed. Newnes, 2004. isbn:
9780750662666.
[16] Tuncer Cebeci. Analysis of turbulent flows. 2nd ed. Elsevier, 2004. isbn:
9780080443508.
[17] Noel de Nevers. Fluid mechanics for chemical engineers. 3rd ed. McGraw-
Hill, 2004. isbn: 9781259002380.
[18] Pierre Sagaut. Large eddy simulation for incompressible flows: an in-
troduction. Ed. by Charles Meneveau. 3rd ed. Springer, 2006. isbn:
9783540263449.
292
[19] James O. Wilkes. Fluid mechanics for chemical engineers. with microflu-
idics and CFD. 2nd ed. with contributions by S. G. Birmigham, B. J.
Kirby, COMSOL (FEMLAB), and C-Y. Cheng. Prentice Hall, 2006. isbn:
9780131482128.
[20] Henk Kaarle Versteeg and Weeratunge Malalasekera. An introduction to
computational fluid dynamics: the finite volume method. 2nd ed. Pearson
Education, 2007. isbn: 9780131274983.
[21] Ascher Shapiro and the Massachusetts Institute of Technology. Na-
tional Committee for Fluid Mechanics Films. 2008. url: http://web.mit.
edu/hml/ncfmf.html (visited on 09/23/2017).
[22] Frank M. White. Fluid mechanics. 7th ed. McGraw-Hill, 2008. isbn:
9780071311212.
[23] Hendrik Tennekes. The simple science of flight. From insects to jumbo
jets. 2nd ed. MIT Press, 2009. isbn: 9780262513135.
[24] J.J Bertin and R.M. Cummings. Aerodynamics for engineers. 5th ed.
Pearson/Prentice Hall, 2010. isbn: 978-0132272681.
[25] Yunus A. Çengel and John M. Cimbala. Fluid mechanics. Fundamentals
and applications. 2nd ed. McGraw-Hill, 2010. isbn: 9780070700345.
[26] Philipp Schlatter, Mattias Chevalier, Miloš Ilak, and Dan S. Henning-
son. The structure of a turbulent boundary layer studied by numerical
simulation. 2010. arXiv: 1010.4000 [physics.flu-dyn].
[27] Philipp Schlatter and Ramis Örlü. “Assessment of direct numerical
simulation data of turbulent boundary layers”. In: Journal of Fluid
Mechanics 659 (2010), p. 116. doi: 10.1017/S0022112010003113.
[28] J. H. Lee, Y. S. Kwon, N. Hutchins, and J. P. Monty. Spatially devel-
oping turbulent boundary layer on a flat plate. 2012. arXiv: 1210.3881
[physics.flu-dyn].
[29] Bruce R. Munson, Theodore H. Okiishi, Wade W. Huebsch, and Alric P.
Rothmayer. Fluid mechanics. 7th ed. Wiley, 2013. isbn: 9781118318676.
[30] Jiyuan Tu, Guan Heng Yeoh, and Chaoqun Liu. Computational fluid
dynamics: a practical approach. 2nd ed. Butterworth-Heinemann, 2013.
isbn: 9789382291787.
[31] Peter Alan Davidson. Turbulence: an introduction for scientists and
engineers. 2nd ed. Oxford University Press, 2015. isbn: 9780198779469.
[32] Michael Leschziner. Statistical turbulence modelling for fluid dynam-
ics demystified. An introductory text for graduate engineering students.
World Scientific, 2015. isbn: 9781783266630.
[33] Zhan Su, Jinbo Wang, Patrice Klein, Andrew F Thompson, and Dimitris
Menemenlis. “Ocean submesoscales as a key component of the global
heat budget”. In: Nature communications 9.1 (2018), p. 775. doi: 10 .
1038/s41467-018-02983-w.
[34] Olivier Cleynen. Git repository: Fluid dynamics. Under CC-by-nc license.
2020. url: https : / / framagit . org / olivier / fluidmech (visited on
04/20/2020).
[35] Olivier Cleynen, Germán Santa-Maria, Mathias Magdowski, and Do-
minique Thévenin. “Peer-graded individualized student homework in
a single-instructor undergraduate engineering course”. In: Research in
Learning Technology 28 (2020). doi: 10.25304/rlt.v28.2339.
293