Fluids
Fluids
Fluids
Brian D. Storey
Olin College
© All
2018 Brian D. Storey
rights reserved
Contents
1 Introduction page 1
1.1 Scope of the text 3
1.2 What is a fluid? 4
1.3 Simple fluid properties 4
1.4 Other properties 5
1.5 What is a continuum? 7
1.6 Mathematical prerequisites 9
2 Dimensional analysis 11
2.1 Units 11
2.2 Buckingham Pi Theorem 12
2.3 Making equations dimensionless 25
2.4 Summary 29
3 The empirical approach 31
3.1 Drag revisited 32
3.2 Drag coefficients 35
3.3 Pipe flow 38
3.4 Summary 42
4 Vector calculus notation and review 43
4.1 Scalar and vector functions 43
4.2 Gradient, divergence, and curl 45
4.3 Normal vectors and flow through a surface 47
4.4 Volume integrals 48
4.5 Surface integrals 49
4.6 Surfaces and volumes in 2D 51
4.7 Divergence theorem 52
iv Contents
4.8 Summary 53
5 Diffusive mass transfer 55
5.1 Physical picture of mass diffusion 56
5.2 Continuum concentration 59
5.3 Equation in 1D 65
5.4 Role of simulation 73
5.5 Summary 74
6 Convective mass transfer 75
6.1 Convective mass transfer 75
6.2 Material point of view - be one with the fluid 76
6.3 The material derivative 77
6.4 Liebniz and Reynold’s transport theorem 81
6.5 Conservation of mass: the fluid 82
6.6 Dimensionless formulation 86
7 Conservation of momentum in a fluid 89
7.1 Transport theorem revisited 90
7.2 Linear momentum 91
7.3 F=ma 94
7.4 So what’s a tensor? 101
8 The Navier Stokes equations 103
8.1 Euler’s equation 104
8.2 Newtonian fluid & incompressible flow 106
8.3 Boundary conditions 109
8.4 Computing stress from flows 110
8.5 Comments on kinematics 111
8.6 Non-dimensionalization 117
8.7 The Reynolds number 118
8.8 Summary 119
9 Solutions to the Navier-Stokes equations 123
9.1 Flow between parallel plates - Poiseuille flow 124
9.2 Flow driven by a wall - Couette flow 129
9.3 Flow down a ramp 131
9.4 Combined Poiseuille and Couette flow 132
9.5 Slider bearing 134
9.6 Impulsively started Couette flow 138
9.7 Cylindrical Couette flow 140
Contents v
Bibliography 215
1
Introduction
The text lays the foundations for the study of fluid dynamics. The
first question a student might ask is why should we be interested in
fluid dynamics? Let me address the question through some examples
of applications where these topics are relevant.
Biology: Our bodies are mostly water. Fluids are essential to life as
we know it. The understanding of the transport of oxygen and nutrients
throughout the body by fluids (at the level of lungs, heart, and blood
vessels as well a the cellular one) requires a basic understanding of the
topics herein. Locomotion of many species from fish to birds to small
microorganisms is determined by fluid flow.
Geophysical: The atmosphere and oceans are fluids. The currents,
the winds, the gulf stream, and the jet stream carry energy and matter
around the planet. Our basic understanding of the weather, the climate,
and the environment can only begin with a basic understanding of fluid
dynamics. The Earth’s magnetic field is generated by convection of the
mantle deep within the Earth.
Energy: Most of our current energy needs are met by the combustion
of fuel; the basic processes of combustion depend upon the details of
the fluid mixing of fuel and air and and the resulting heat release.
Advances in understanding of the details of these processes has led to a
number of developments over the years which made combustion cleaner
and significantly reduced the number of pollutants emitted from cars.
Many of the renewable sources of energy such as wind, geothermal,
ocean waves and tidal seems obvious to state that fluid flow plays a
prominent role. It is fair to claim that an understanding of all energy
systems must begin with the basic understanding of topics in this text.
2 Introduction
Density which typically has the symbol, ρ, is nothing more than the
mass per unit volume. Everything we discuss in the development of the
theory is given on a per unit volume basis, thus the density will show
up throughout the course. Density of a liquid is really easy to measure,
just fill a beaker of known volume and weigh the mass of the fluid. The
ratio of mass divided by volume is density. When we write our fluid
version of Newton’s laws, density will take the role of mass when we
write F = ma.
Viscosity which typically has the symbol µ (though η is common
in chemical engineering) describes a substance’s resistance to flowing.
You probably have some intuition about this concept based on pouring
water and syrup. A simple device for quantifying viscosity comprises
two concentric cylinders with a very narrow gap between them. A fluid
of interest is placed in the gap. The inner cylinder is held fixed while the
outer one rotates at a constant angular speed, ω. The speed of the wall
at the outer cylinder is simply U = ωR, where R is the radius of the
cylinder. If we conducted this experiment for different gap sizes d and
different speeds, we would find the torque needed to spin at constant
speed would follow a law,
U
Torque ∝ µ .
d
We would find for all things equal, syrup is harder to spin than water
because the syrup has a higher viscosity. Not all fluids would follow the
law above. Only simple fluids such as water, air, syrup, and oil. Fluids
that follow a law such as that above are called Newtonian, a definition
we will be precise about later.
can hold water droplets to the ceiling tiles in your shower. Surface
tension also acts at the interface between two immiscible liquids like
oil and vinegar. If there is no free surface in the liquid then surface
tension does not play a role. For flows at a liquid interface for objects
the size of boats and people, surface tension is too small to matter
much. For objects that are measured in millimeters, surface tension
can be a dominant force.
Speed of sound is exactly what it says - it is the speed that sound
travels in the fluid. The speed of sound becomes important when flow
speeds start to approach it. The ratio of flow speed (or speed of the
object moving through the fluid) to sound speed is called the Mach
number, Ma. When the speed of the object becomes greater than the
sound speed, Ma > 1, then we have supersonic flow. Supersonic flow is
very interesting as much of our intuition about fluid flow goes out the
window. Supersonic flow is a topic for another course.
When the Mach number exceeds about 0.3, then the flow becomes
compressible meaning that the density can no longer be considered a
constant (this is a rule of thumb, not an exact criteria). It is the Mach
number that tells us when we need to start thinking differently about air
and water even though we called them both fluids. Gases are relatively
easy to compress when placed inside a closed container while liquids
are not. When we have flow, it is the Mach number that characterizes
whether the flow is compressible or not. Note that it is important to
distinguish between a compressible fluid and a compressible flow. Air is
an easily compressible substance, however most flows that we encounter
with air on a daily based are incompressible flows. It is the Mach number
that helps determine if we have a compressible flow.
Elasticity Normally we think of solids exhibiting elasticity. I pull
on a rubber band it will stretch. When I let go it will return to its
initial state. Simple fluids like water, air, syrup and simple oils do not
exhibit any significant elasticity. If I shear the fluid and then let go, the
fluid does not try to return to its original configuration. Some fluids
such as biological fluids or polymeric solutions exhibit both elasticity
and viscosity - typically called viscoelastic fluids. Silly putty is a classic
example, I can roll it in a ball and it will bounce or leave it on the table
and it will puddle up. In the case of silly putty, it is the time scale of
the flow that distinguishes between whether the flow is more viscous or
1.5 What is a continuum? 7
z
Density
took our measuring volume near the ceiling we might expect a different
answer than near the floor. This idea is shown schematically in Figure
1.1, where the local measurement of density is shown to be linear with
height, with less dense fluid rising to the top. As we shrink the mea-
suring volume we obtain a more local measure of the density. At some
point we keep shrinking the measuring volume and we keep getting the
same answer. This constant value would be what we would consider
the local, continuum value of the density at that point in space. If we
continued to shrink the measuring volume we run into trouble with the
fact that air is comprised of many molecules. Eventually the measur-
ing volume would be so small that the mass we measure would depend
upon how many molecules we managed to trap. If there are only 10
molecules on average in our volume, then the answer starts to fluctuate
depending on if we randomly had 8, 9, 10, 11, or 12 molecules in the
volume at the time of measurement. The continuum limit is thus the
intermediate asymptotic, where we take the limit such that we can talk
about local quantities (i.e. density, temperature, pressure, velocity) at
a point in space, but we don’t take the limit too small that the molec-
ular nature of matter comes into play. As long as the geometry of the
problem is large compared to the molecular scale, then the continuum
assumption is good.
1.6 Mathematical prerequisites 9
This theorem equates the flux of a vector field through a closed surface
to the volume integral of the divergence of the vector field. Above, we
use the notation that dS is a unit of surface and dV is a unit of volume.
If if the above notation seems vaguely familiar to you but you have
forgotten the details, you will have no problem. We will review vector
calculus in a later chapter. If you have never seen these symbols before,
you might have some trouble understanding some of the later chapters.
2
Dimensional analysis
2.1 Units
We often refer to fundamental and derived units. Fundamental units in
this course will be mass (M ), length (L), time (T ), and temperature
(Θ). Derived units are things like velocity (L/T ), force (M L/T 2 ), or
density (M/L3 ). The distinction above is arbitrary. Of course we could
say that force is fundamental and mass is derived. If you want to do
12 Dimensional analysis
this, this is your right. I find it easiest to stick with the distinction
above.
We could relate temperature and energy and thus remove tempera-
ture as a fundamental unit. In this course, this path would not lead to
simplification. It would require us to bring in a new parameter, Boltz-
mann’s constant. Boltzmann’s constant bridges the statistical definition
of kinetic energy of molecules to the macroscopic temperature. If we are
not dealing with statistical mechanics, then introducing Boltzmann’s
constant just leads to confusion. In this course, everything is macro-
scopic and thus treating the temperature as its own fundamental unit
is always going to be the way to go.
2.2.1 Pendulum
The first example on dimensional analysis every book uses is the pen-
dulum. I won’t be original. Imagine a point mass (m) hanging on a
massless string of length (`) in a gravitational field (g); shown schemat-
ically in Figure 2.1. There is no friction. All this is an idealization, but
it is one that we can easily realize in practice to a good degree of ap-
proximation. The question is what is the period of oscillation, t, for a
given starting angle θ and how does it depend upon the parameters in
the problem? There are five independent parameters (m, `, g, θ and
t). These parameters are made up of 3 independent dimensions (M , L,
and T ). Applying the theorem tells us that N = 5, R = 3 and thus
there are only 2 independent parameters.
There are a number different ways to formally approach the removal
of units from the problem to decide what the dimensionless parameters
are. Once you are comfortable, you can often just remove the units by
inspection. I like the table method as it is a provides a good framework
2.2 Buckingham Pi Theorem 13
θ g
ℓ
m
for formally approaching the problem. We make a table with all the
parameters and their units in brackets. We put the primary thing we
want to know, in this case the period, in the first column. For this
problem, our table starts as,
t [T ] m [M ] ` [L] g [ TL2 ] θ
Note that the starting angle θ has no units. Angles are measured in
radians which is defined as the arc length divided by the radius, thus
the radian is already dimensionless.
Now start adding lines to the table, and in each new line remove one
of the independent dimensions and one dimensional parameter from
the problem. Let’s start by removing mass from the problem. Since
the dimension M only exists in the mass, there is nothing to do but
remove it from the problem. Mass cannot be a parameter to the problem
as there is no way to cancel that fundamental unit. With each new line
to the table it is useful to add yourself a little note to the right.
t [T ] m [M ] ` [L] g [ TL2 ] θ
t [T ] ` [L] g [ TL2 ] θ remove M
Now let’s remove length from the problem by defining a new param-
eter g/` that does not have length in it. We expand our table by one
line so it reads,
14 Dimensional analysis
t [T ] m [M ] ` [L] g [ TL2 ] θ
t [T ] ` [L] g [ TL2 ] θ remove M
g 1
t [T ] ` [T2 ] θ remove L
Finally, we can remove time from the problem by adding another line
to our table as follows,
t [T ] m [M ] ` [L] g [ TL2 ] θ
t [T ] ` [L] g [ TL2 ] θ remove M
g 1
t [T ] ` [T2 ] θ remove L
t g`
p
θ remove T
p
The final result is that there are two parameters, t g/` and θ. We
can then state that the answer has the following functional form,
r s
g `
t = f (θ) or t = f (θ)
` g
At this point we don’t know what the function is, but it says the period
of any pendulum of any length in any gravitational field only depends p
upon its starting angle. We only need to measure time in units of g/`,
and all pendulums are the same. If we always started the pendulum
from the same angle, the function on the right hand side would then
just be a constant. The result tells if the length of the string is increased
by
p a factor of 4, then the period increases by a factor of 2 to keep
t g/`=constant. You can easily confirm this result experimentally.
c A2
b
A1
θ
a
2.2.3 Caveats
There is a nice paper written by Lord Rayleigh on the “Principle of
Similitude” (Rayleigh (1915)); note that similitude is another name for
dimensional analysis. He starts the paper;
I have often been impressed by the scanty attention paid even by original
workers in physics to the great principle of similitude. It happens infrequently
that results in the form of “laws” are put forward as novelties on the basis
of elaborate experiments, which might have been predicted a priori after a
minutes’ consideration.
harder for the sphere to move through the more viscous fluid all other
things being equal. It also seems reasonable to suspect that the fluids
density, ρ, might matter as well. These are the only two fluid properties
that really matter in this problem. We are starting to already see the
comments of Bridgman come in. We have to know something about the
physics of the problem to know that these are the only fluid properties
that matter. Why doesn’t the speed of sound in the fluid matter? Or its
surface tension? In later chapters as we develop the full theory of fluid
flow, we will see that these are the only two properties that emerge for
this problem. For now you will have to accept what I am suggesting.
Applying the Pi Theorem helps our problem immensely. There are
5 parameters, F , D, U , µ, and ρ. There are the 3 dimensions of M ,
L, and T . The Pi Theorem says there are only two parameters that
matter. We can proceed by constructing the table and writing all the
parameters and their units.
F [M L
T2 ] D [L] U [ TL ] ρ [ LM3 ] M
µ [ LT ]
F [M L
T2 ] D [L] U [ TL ] ρ [ LM3 ] M
µ [ LT ]
4 µ 2
F L
ρ [T2 ] D [L] U [ TL ] ρ [ LT ] Remove M
F 2 ρU 1
ρU 2 [L ] D [L] µ [L] Remove T
F ρU D
ρU 2 D 2 µ Remove L
The result says there are two parameters. One parameter is ubiqui-
tous in fluid dynamics and we will discuss it extensively in this course.
It is called the Reynolds number, Re = ρUµD . If you have the chance
to make a grouping that looks like (density·velocity·length)/viscosity,
18 Dimensional analysis
you should do so. The other parameter is the dimensionless drag force,
known as the drag coefficient. Typically, the drag coefficient is defined
by convention to be Cd = 1 ρUF2 πR2 ; which is the same as we got from
2
the table only with a few different arbitrary constants; namely a factor
of 1/2 and πR2 instead of D2 . There are good reasons that we will
come to later in the course for using the constants in the classic form
of the drag coefficient. The dimensional analysis result is powerful as
the result says there is a single master curve, namely
Cd = f (Re).
This single master curve, the function f , captures the drag behavior of
all spheres in all fluids. It turns out the details of this curve can be quite
complicated and very difficult to calculate from the basic equations of
fluid dynamics. However, we can conduct experiments. The dimensional
analysis says that if we conduct the experiment of measuring drag as a
function of speed once - for one size sphere in a wind tunnel - we now
know the result for any other sphere in any fluid. A fit to experimental
data on a logarithmic scale is shown as the solid curve in Figure 3.1.
I want to emphasize that dimensional analysis let’s us represent data
that depends on multiple parameters in a simpler way. In dimensional
terms the drag force is given as F (U, D, µ, ρ). Parameterizing this func-
tion experimentally for all four variables requires a lot of experiments.
Lets say we would want 10 experimental data points to establish a
2.2 Buckingham Pi Theorem 19
U [ TL ] ` [L] D [L] M
∆P [ LT 2] ρ [ LM3 ] M
µ [ LT ]
As before, there are some arbitrary choices to make about how to
proceed. If we try to eliminate mass, we already see that we could
divide the pressure drop by either the density or viscosity. While any
choice is just as good as the other, one possibility is worked out below,
U [ TL ] ` [L] D [L] M
∆P [ LT 2] ρ [ LM3 ] µ M
[ LT ]
2 ρ
U [ TL ] ` [L] D [L] ∆P
ρ
[ TL2 ] µ
[ LT2 ] Remove M
ρU
µ
[ L1 ] ` [L] D [L] ∆P
ρ2 U 2
Remove T
ρU D ` ∆P
µ D ρU 2
Remove L
This result is one of many possibilities. My result says there are three
parameters. The first is again the Reynolds number - it will show up
in nearly every fluid mechanics problem. The second parameter is a
geometric parameter, the length to diameter ratio. The final parameter
is the dimensionless pressure drop. Now, conducting experiments for
three parameters is a lot easier than 6. We could fit all the results on a
single graph plotting the dimensionless pressure drop versus Reynolds
number for a family of curves for different `/D. We need a sheet of
paper rather than a book to display all results.
We can actually do even better, but the comment by Bridgman again
applies. We need a little physical reasoning and some knowledge of the
problem. This step might seem a leap now, however, as we learn more
throughout this course we will make this step more easily. The pressure
drop and pipe length are not really independent. It is the parameter
∆P/` that matters most, the pressure gradient. The assumption we
make here is that when the pipe is long that any section of pipe is
basically the same. If we look at the flow here or there in the pipe, it
is just a length of pipe and the flow doesn’t change. If the character
of the flow doesn’t change as a function of length, then doubling the
length should not change anything if we double the pressure drop. The
pressure drop per unit length is the parameter that really matters.
We could confirm this assumption experimentally by measuring the
pressure as a function of length and we would find that it is essentially
linear with distance. This assumption that pressure gradient matters
really only applies to pipes that are long relative to their diameters.
If we use the pressure gradient as our parameter our result is sim-
2.2 Buckingham Pi Theorem 21
a) b)
Figure 2.4 Dimensional (a) and dimensionless (b) plots of the flow-
pressure drop data for the pipe flow problem. This data was taken
in a simple apparatus and consisted of 21 separate configurations
varying fluids (water and oil), pipe diameter, and pipe length. The
experimental dimensional data in (a) is collapsed to a single master
curve in (b). I took this data with a simple table-top device thus
the data at high Reynolds number show some significant scatter
that can be reduced with some care.
turbulent flow in any pipe with any fluid. It is also important to real-
ize that he images seen in Figure 2.5 only depend upon the Reynolds
number. If we repeated the experiment in water, air, and oil - but we
changed the size of the pipe to keep the Reynolds number the same
across the three experiments - we would make the same visualizations.
small force for large objects on the size of boats with people in them.
So our physical intuition (that we are just building) says that g is the
only new parameter.
The drag force, F , is a dimensional function, F (`, U, ρ, µ, g). The Pi
theorem has 6 parameters expressed in 3 independent dimensions. We
expect one new dimensionless parameter over what we would find for
the submarine problem. Let’s proceed with the table and see what we
get,
F [M L
T2 ] ` [L] U [ TL ] ρ [ LM3 ] M
µ [ LT ] g [ TL2 ]
F L4 µ 2
Again, the exact choices at each step above are arbitrary and I could
have come up with a number of equivalent solutions. I took a path that
leads toward the standard way the result is presented. The new di-
mensionless parameter, is called the Froude number, and is historically
defined as
U
Fr = √ .
g`
The dimensionless drag force can be expressed as,
F ρU ` U
= f , √ = f (Re, Fr)
ρU 2 `2 µ g`
This result has an important implication for our idea of using model
experiments for testing drag on boats. We can’t change the dimension-
less parameters to arbitrary values in practice. We are pretty much
stuck with g, at least until we can colonize other planets. We also can-
not change ρ by much - most liquids tend to have similar densities as
water. We could change the density by 20% by adding salt to water.
We could take an extreme case and use liquid mercury. However, we
are limited in what choices we can make. We also can easily increase
µ, but decreasing it from a value much less than that of water becomes
challenging. It is not practical to be able to set the Reynolds number
and Froude number completely independently.
The things we can vary with some degree of control are the veloc-
ity and the length. The relationship between velocity and length are
2.3 Making equations dimensionless 25
is just some constant that we pick to measure our unit of time in. There
is nothing special about the second or the minute, so we can change the
unit to whatever we feel like. Picking the ”right” scale for t0 can be a
bit of an art form that comes with experience. However, the equations
will serve as out guide us to a ”good” answer. Mathematically, t0 could
be anything.
Lets do a few examples to make this clear.
2.3.1 Pendulum
Since we started the chapter with the pendulum example, let’s repeat
this example using the equation approach. Newton’s law, F = ma, for
a simple pendulum with no friction is,
d2 θ
m` = mg sin(θ),
dt2
which is easily rewritten as
d2 θ g
2
= sin(θ).
dt `
We have already recovered the basic result of the Pi Theorem. We have
shown that the mass doesn’t matter and that the only parameter that
matters is g/`. However, the equation still has units so let’s proceed to
remove them all. The angle, θ, has no units so we leave it alone. Time,
t, has units so lets make the transformation t̃ = t/t0 (or equivalently
t = t̃t0 ).
d2 θ g
2
= sin(θ).
d(t̃ t0) `
Note how we made the derivative dimensionless. This step seems strange
to most students the first time they see it but we have to remember
that a second derivative in time has units of 1/seconds2 . Thus, the dt2
in the derivative is replaced by d(t̃ t0)2 . All we did was the substitution
t = t̃ t0 . Since t0 is a constant it can be moved in or out of the derivative
operator. This step is no different than changing the time derivative to
be measured in minutes instead of seconds.
Our equation can be rewritten as
d2 θ gt2
2
= 0 sin(θ).
dt̃ l
2.3 Making equations dimensionless 27
2.3.2 Mass-spring-damper
As an example where we have additional parameters, consider a simple
mass-spring-damper system,
d2 x dx
m = −β − kx
dt2 dt
with the initial conditions that x(t = 0) = x0 and the velocity initially
is zero.
We will make the equations dimensionless by defining x̃ = x/x0 and
t̃ = t/t0 . Here, t0 is arbitrary, but x0 is the initial position. Using
the initial position seems like a good choice to make x dimensionless
28 Dimensional analysis
2.4 Summary
Throughout this book we will continue to return to the idea of work-
ing in dimensionless terms. We will periodically appeal to both the
table and equation based approach - realizing that we can get the same
results from both approaches. We will see many examples where the
mathematical description of a problem can be stated, but it is too dif-
ficult (or impossible) to solve the problem. In these cases we can still
obtain insight to the problem by appealing to dimensional analysis.
3
The empirical approach
In the last chapter we found that by using physical insight and know-
ing the important parameters in a problem we could use dimensional
analysis to dramatically reduce the number of parameters needed to
characterize a problem. We saw through the example of drag on a
sphere and flow in a pipe that the complexity of a problem in terms of
how many experiments need to be done could be dramatically reduced.
All the behavior is described on a single plot or with a single number.
After this current chapter, we will spend much of the rest of the
book developing the underlying theory for fluid flow. We will find that
this theory can be extremely accurate when compared to experiments
and has great predictive power. We will also find that in many cases
of practical interest even with great computational resources and so-
phisticated mathematical tools, many problems are just too difficult
to solve. Fortunately, many of these cases are pretty easy to measure
experimentally. We often find that dimensional analysis coupled with
experimental data can lead to practical solutions and accurate esti-
mates to real world problems. Many times there is no real need for a
time consuming or computationally expensive solution.
The focus of this chapter will be to show how dimensional analysis
plus experimental data can be put to practical use in an engineering
setting. We can think of this approach as purely empirical, While we will
only go through a few examples, this approach in a sense could be the
end of our practical study. We might not have a lot of physical insight or
deep understanding - but we could perform simple calculations, make
predictions, and design useful things.
While all the interesting stuff (to me, anyway) will come up in later
32 The empirical approach
chapters, this chapter may be the most useful when you just need to
solve problems. We will revisit two examples from the last chapter on
drag and pipe flow and explore them in more depth. The chapter here
is meant to be illustrative and not comprehensive. Many textbooks
on engineering fluid dynamics and have numerous details and practical
situations that have been studied. My aim here is to show this empirical
approach first, and then dig into the details a little later in the book.
which states that the drag coefficient will be equal to a constant divided
24
by the Reynolds number. The data supports the function of Cd = Re .
The factor of 24 comes from a full calculation of the laws of fluid flow
and is also validated by experiments. Constants simply cannot come
from dimensional analysis.
The result I just “derived” probably seems suspicious to you. If it
doesn’t, it should. Why am I allowed to throw out density from my list
of parameters? My derivation again relies on this comment by Bridg-
man. You have to understand the physics of the problem to successfully
apply dimensional analysis. Since this is just the start of your study,
you should have no reason to accept that density is unimportant at low
Reynolds number. Remember that I have studied fluid mechanics since
graduate school, have read about this problem in numerous books, and
have run the experiment and plotted data myself. This problem is easy
for me at this point; however, faced with a new unfamiliar problem I
have to work hard to get the right answer. With a new problem, I would
also not typically believe an answer until is was backed up by a more
thorough analysis or experimental data. It will take some time until
we study fluid motion in more detail for this example to make physical
sense to you. However, the excellent fit to the experimental data at low
Reynolds number should convince you that we are on to something.
If we continued the graph in Figure 3.1 to lower and lower Reynolds
24
number we would see the fit to Cd = Re continue to high precision.
The other limit at higher values of Reynolds number where the drag
coefficient is approximately constant can also be “derived” from dimen-
sional analysis. In this limit the fact that the fluid has viscosity matters
less than the fact that the fluid has inertia. At high speeds we have to
move the fluid, which has mass, out of the way for the ball to go by.
This explanation is actually a little too simplistic and the truth is more
subtle, but let’s just assume for now that the viscosity doesn’t matter
and see what the analysis says.
F [M L
T2 ] D [L] U [ TL ] ρ [ LM3 ]
F L4
ρ [T2 ] D [L] U [ TL ] Remove M
F 2
ρU 2 [L ] D [L] Remove T
F
ρU 2 D 2 Remove L
3.2 Drag coefficients 35
Our result now says there is one dimensionless parameter and it must
be a constant,
F
= Constant.
ρU 2 D2
We see the above equation is the definition of drag coefficient. The
analysis says that drag coefficient is constant if the viscosity of the
fluid is irrelevant. Again, I should not have convinced you that it is
acceptable to ignore the fluid’s viscosity, however the data between a
Reynolds number of 103 and 105 supports this view. If we look at the
data we see a dip in the drag coefficient that we will later explain - but
it turns out the dip occurs because the fluid has viscosity.
In dimensionless terms, there is a universal master curve of one pa-
rameter (the Reynolds number) and that curve is good for all drag
experiments with a sphere that have ever been done and all that will
be done in the future. The master curve works for all fluids and spheres
of any size. However, we demonstrate here that if we understand even
a little something about the physics of our problem, we can often find
even simpler limits that can be good over a wide range of the parameter
space.
That the behavior changes dramatically with Reynolds number can
be seen through the flow visualization images in Figure 3.2. At low
Reynolds number (here Re=1.5 or lower) the flow looks very symmetric
from left to right. In this regime the flow is dominated be viscosity. At
Re=26, we see asymmetry and a recirculating wake forming behind
the cylinder. This is the effect of inertia - the fluid approaching the
cylinder has momentum that causes the flow to tend to “overshoot” on
the backside of the cylinder. As the Reynolds number increases to 2000
and 10,000 the wake becomes very complex and has no resemblance to
the flow at low Reynolds number. It is also interesting to realize that
the flow field is only dependent on Reynolds number. The experiments
in Figure 3.2 could be repeated in water, air, or oil and as long as the
Reynolds number was matched the flow would look identical.
a) b)
c) d)
F = CµU D,
where C is constant. If the flow was dominated by inertia the force had
a form
F = CρU 2 D2 .
ber, defined as
ρU D
Re = .
µ
The Reynolds number is usually large for things that you interact with
on a daily basis - a 1 mm object traveling at 1 mm/s in water will
have a Reynolds number of 1. What counts as high or low Reynolds
number depends on the situation, but usually greater than 1000 would
be considered well in the high Reynolds number regime. The exact
number really depends on the situation.
The most common way that the drag coefficient is characterized is,
1
F = Cd ρU 2 A,
2
where A is the area projected by the object to the flow. It is important
to note that this choice of area used in the formula is arbitrary. The
area used must be reported along with the drag coefficient to avoid any
ambiguity. If the drag coefficient is measured but the area that was
used to compute Cd is not reported, then the result is useless.
For many objects, the drag coefficient would be reported as a con-
stant. In some cases it will be given as a chart as a function of Reynolds
number. If the number is given as a constant, then most likely it is as-
sumed that the Reynolds number is large. Note that if it is given as
a constant in a table, it doesn’t mean that it is truly constant with
Reynolds number as seen in the sphere example. However, looking at
Figure 3.1, note how the assumption of constant drag coefficient is
pretty good over a wide range of conditions.
With a simple search will can find a lot of resources and tables of drag
coefficients. You will notice that most are around 1 for blunt objects
(sometimes around 2, but never 10) and as low as 0.04 for relatively
streamlined shapes. We will dig into the physics of this problem again
in a later chapter. For now, if you need to figure out the drag force on
an object it is pretty simple (at least conceptually),
• Compute (or estimate) the Reynolds number.
• Search for drag coefficients on the shape of interest.
• Understand what area was used on the reported drag coefficient.
Without the area defined, the coefficient is useless.
• If the coefficient is given as a function of Reynolds number, you can
just look get the coefficient for your Reynolds number.
38 The empirical approach
1 2 `
∆P = ρU f Re,
2 D D
40 The empirical approach
Figure 3.4 Moody diagram for the friction factor in pipe flow. This
figure is a reproduction of the way the data are classically presented.
1 2 `
∆P = ρU f (Re) .
2 D
3.3 Pipe flow 41
If the velocity of the flow (or volumetric flow rate) is known, then the
solution procedure for the needed pressure to drive the flow is straight-
forward. The procedure is,
• For the given pipe size and fluid, compute the Reynolds number.
• Decide on a good estimate of the pipe roughness.
• Look up f (Re) on the diagram.
• Compute the pressure drop.
If the applied pressure is known, then the solution procedure for the
flow velocity is not as straightforward. The procedure is,
• Guess a Reynolds number (you can’t compute it since you don’t know
the velocity).
• Decide on a good estimate of the pipe roughness.
• Look up f (Re) on the Moody diagram.
• Compute the flow velocity as for the known pressure case.
• Compute the Reynolds number for this new flow velocity.
• Look up f (Re) on the Moody diagram.
• Compute the flow velocity again.
• Repeat (if needed) until you get a converged solution.
1 2 64 `
∆P = ρU .
2 Re D
or
32µ`
∆P = U.
D2
For laminar flow there is a simple linear relationship between pressure
and flow velocity.
42 The empirical approach
3.4 Summary
These two examples show how we can combine dimensional analysis
with experimental measurements such that we can make accurate pre-
dictions. There is no satisfactory ”why” embedded in any of these re-
sults. However, the results can be used to design and build real devices.
Predictions can be made that will be really accurate, since they are
rooted in experiments. I again want to emphasize that we covered only
a few examples here and other books provide many more examples of
piping losses and drag on objects. We will revisit the physics of drag in
a later chapter.
4
Vector calculus notation and review
Starting in the next chapter we will begin to formulate the laws which
govern fluid flow. So far we have relied on observations, experiments,
and dimensional analysis to guide us. As we develop our more complete
theory, we will find laws will allow us to compute quantities such as the
temperature field which can vary both in time and space. Our laws
will be ”local” in that they can tell us quantities at a point (in the
continuum sense) in space. Inherent in our derivations of these local
laws will be volumes, surfaces, and integrals. Using the language of
vector calculus will become essential to maintain our results in a form
compact enough to provide some physical insight.
I am assuming that upon reading this that you have taken a course
in vector calculus but have likely forgotten many of the finer details.
This chapter is meant to review the notation and some of the useful
theorems. If you have never come across these topics then my coverage
here will not be sufficient. Our use of vector calculus will be as the
language we use to describe the physics. We will not need to carry
out complicated surface and volume integrals, however, we will need to
come to understand an intuitive and physical level what a surface or
volume integral is.
v = ui + vj + wk
∂ ∂ ∂
∇= i+ j+ k,
∂x ∂y ∂z
∂P ∂P ∂P
∇P = i+ j+ k.
∂x ∂y ∂z
Remember, the gradient of a scalar field gives a vector. We can also take
a gradient of a vector field, but let’s leave that issue to the side for now.
The most important property of the gradient is that the resulting vector
field always points perpendicular to the contours and points ”uphill”. If
you see closely spaced contour lines on a topographical map you know
the terrain is very steep.
When we take the dot product of ”del” with a vector field we define
the divergence which gives back a scalar field. The divergence of the
velocity vector is defined as,
∂u ∂v ∂w
∇·v = + + .
∂x ∂y ∂z
The divergence is a measure of how much a vector field is, well, diverg-
ing. If you see a point in the vector field where all the velocity vectors
were flowing out of a point, this would have positive divergence. If all
the vectors were flowing into a point, then that would have negative
divergence. If some vectors were flowing in and some out, and if they
were in perfect balance there would be zero divergence. The divergence
is shown schematically in Figure 4.2. We will see that divergence is con-
nected to the idea of conservation. In a fluid flow if there is non zero
divergence of the velocity field, then it means that mass is accumulat-
ing at that point. Non-zero divergence can only occur in a compressible
flow. Remember, the divergence of a vector field gives a scalar field. It
does not make sense to take the divergence of a scalar field.
Our final operator of interest occurs when we take the cross prod-
uct of ”del” with a vector. This operation is defined as the curl. For
4.3 Normal vectors and flow through a surface 47
v v v
example, the curl of the velocity field is (known by the symbol ω),
∂w ∂v
−
∂y ∂z
ω =∇×v = ∂u ∂w .
∂z − ∂x
∂v ∂u
∂x − ∂y
The curl has to do with whether the vector field is spinning or not. We
need to be a little careful, we can have circular looking vector fields
that have zero curl. The physical interpretation of curl will be put to
the side for now and we will dig into a little deeper later in the book.
The curl of a vector results in a vector. It does not make sense to take
the curl of a scalar.
It is important to remember that del has units of 1/Length; for ex-
ample the units of ∇T are Kelvin/meter. If we work in dimensionless
terms we would very often (but not always!) decide to scale the x, y,
and z coordinate directions by the same length scale L. In this case our
dimensionless del would become,
1˜
∇= ∇.
L
Note that the same units apply to divergence, gradient and curl.
n
ρv
Figure 4.3 Schematic of the normal vector and mass flux at a sur-
face. Here the vector field and the normal vector are not aligned so
only some fraction of the total mass flux at this location is cross the
surface.
the idea that if we break the such a sum up into smaller and smaller
bricks that in the limit of lots of little pieces we have an integral,
Z
M = ρdV.
Figure 4.4 Latitude and longitude grid for the earth (from Encyclo-
pedia Britannica).
of surface area. Take for example the grid of latitude and longitude
on the globe shown in Figure 4.4. The grid patches are not equal in
area, thus we would need to weight any sort of sum by the area of each
patch. This weighting of the different area patches in a sum is what the
expression for the surface integral naturally does for us. We will use
surface integrals in many ways.
One example is the average of some scalar function on the surface.
If we wanted to compute the average surface temperature of the earth,
for example, we would express the calculation as,
Z
1
Tave = T dS,
A
the how aligned the mass flux is with normal vector, and
Z
ρv · ndS,
would give the total mass flow (in units of kg/s) through the surface.
We will very often apply the surface integral to an entire closed sur-
face which encases some volume of material. A more formal notation
for the surface integral of a close surface would be,
ZZ
ρv · ndS.
The volume integral of a scalar field is a single number and the the
volume integral of a 3D vector field is a single 3D vector.
about what this means for our 3D world, just imagine that the thing
you draw extends outward and into the page forever. The 2D object
that we draw is a cross section of an infinitely long object. Working in
2D can sometimes confuse our units, so you have to remember that we
are computing things per units depth into the page.
Even though we are using v as our vector field here, the above result
is a general result from calculus and has nothing to do with whether v
is a physical velocity.
We can derive this result by sticking to a simple 2D example using
vector field v with x and y components u and v. The components of the
vector field are functions of space; u(x, y) and v(x, y). For simplicity
we will take our region of interest to be a unit square. We will start by
evaluating the surface integral. On the left boundary v · n = −u and
on the right boundary v · n = u. The change in sign is that on the left
the normal vector and the positive x component of the velocity vector
point in opposite directions whereas on the right they point in the same
direction. The dot product means that the sign will be opposite for the
two terms. The same holds for the upper and lower surfaces, only here
it is the y component of the velocity that shows up from v · n. At this
point,
Z Z y=1 Z y=1
v · ndS = − u(x = 0, y)dy + u(x = 1, y)dy
y=0 y=0
Z x=1 Z x=1
− v(x, y = 0)dy + v(x, y = 1)dx
x=0 x=0
or by grouping terms,
Z Z y=1
v · ndS = (u(x = 1, y) − u(x = 0, y)) dy
y=0
4.8 Summary 53
Z x=1
+ (v(x, y = 1) − v(x, y = 0)) dx.
x=0
Recall that the fundamental theorem of calculus states that for any
function,
Z x=1
∂u
u(x = 1) − u(x = 0) = dx.
x=0 ∂x
Note that we did this derivation for a unit square in 2D. We could have
been a little more general and the result would have held up. This is
one of the most useful theorems in this course. The theorem relates the
total flux of a vector field through a closed surface to the divergence of
the same field inside the volume.
4.8 Summary
As stated in the beginning, I am assuming that the notation is fa-
miliar to you and you are not completely new to vector calculus. If
you are new to the topic, then this “introduction” is going to be too
brief. You probably noticed that I focused more on the physical inter-
pretation of various surface and volume integrals rather than actually
computing surface and volume integrals. This focus on the meaning will
continue through the book. We will use the language of vector calculus
to describe our fundamental conservation laws and the notation is used
throughout the derivations.
5
Diffusive mass transfer
There are two mechanisms by which we can mix a dye with a fluid;
convection and diffusion. This chapter will only be concerned with
molecular diffusion in a stationary fluid. This chapter will set the stage
for the next chapter where we will start to see what happens when the
fluid starts moving. Starting with diffusion is also useful as it gives us
a simple case where we can start to connect our physical intuition with
the formal framework of vector calculus.
56 Diffusive mass transfer
Figure 5.2 Box where the left has higher concentration than the
right. .
imaginary line (i.e. c(x + dx) − c(x)), then the net flux would propor-
tional to the concentration gradient. The relationship between flux and
concentration gradient is in fact what is observed with the random walk
and is known as Fick’s law. We will formalize Fick’s law very soon.
N = cV.
Now let’s take a case where the dye is not evenly distributed. We could
imagine breaking the liter volume up into 1000 mL size samples and
count the molecules in each sub-volume. The total number of molecules
would then be the sum of the numbers in each element. In the contin-
uum limit when we can describe the concentration as a local scalar
field, c(t, x, y, z), then we formally find the total number of molecules
60 Diffusive mass transfer
This equation is nothing more than like counting the number of people
in a room. Stand at the door and count how many people come in
and out and you know by how much the total number of people in the
room has changed. In the subject of fluid dynamics, we typically work
in rate form, which we can obtain by taking the time derivative of the
expression,
dN
= Ṅin − Ṅout . (5.2)
dt
I should comment on the use of Ṅin to denote the rate of molecules
coming into our volume. It is common notation to use a dot above a
variable to denote a rate. This dot should not be confused with the
one often used in mathematics books where df /dt ≡ f 0 (t) ≡ f˙(t).
Making the analogy of the doorman, if I count 60 people entering into
a room in one minute, then I would have and average rate of Ṅin = 1
person/second.
In analyzing this equation in a continuum sense, we will need a way
to describe the rate that particles are leaving an arbitrary volume ele-
ment. To provide more formalism, we need to define a mass flux vector,
5.2 Continuum concentration 61
j. The boldface notation denotes a vector quantity. Mass flux has mag-
nitude and the direction points in the direction, on average, that the
dye molecules are moving. We saw previously that we expect the flux
to point from more to less concentrated areas of dye. The units of flux
are number of dye molecules per unit time per unit area. It is fine to
think of the flux as related to the average velocity of the continuum
collection of the dye molecules at that point in space. Remember, for
now we consider that the carrying fluid remains stationary.
If we take the surface of an arbitrary volume, and draw a surface nor-
mal vector n to point outward, then j · n is the rate that dye molecules
cross the surface at that point, per unit area. The dot product with the
surface normal vector is important since the mass flux is a vector. If
the mass flux is pointing tangential to a surface, then no dye molecules
are crossing the surface. If the flux vector and the normal vector are
aligned, then all the movement of the dye is such that it is cross the
surface. See the schematic in Figure 5.3.
The net rate that dye molecules flow across the entire boundary of a
closed volume is given by a surface integral. If we consider our surface,
S, which encloses our entire volume of interest the net rate that dye
molecules cross that surface is
Z
Ṅin − Ṅout = − j · ndS. (5.3)
The minus sign is there because the convention is to define the normal
vector to point outwards from a closed surface. The dot product of the
normal vector and the mass flux vector automatically gets the sign for
inward and outward flux from the volume correctly.
Making the analogy with people coming and going from a room,
the surface integral would be equivalent to giving 1000 bouncers, 1000
doors around the large room to watch. Each bouncer would count the
number of people coming and going. The surface integral would be the
summary report from all the bouncers. The surface integral counts the
amount of dye coming and going per unit area all along the surface,
and the adds it all up.
62 Diffusive mass transfer
n
j
j j j
Figure 5.4 Schematic of the divergence of the mass flux. The con-
centration at a point rises, falls, or stays the same depending on the
local divergence of the mass flux vectors.
An integral can be equal to zero when the thing you are integrating is
non-zero. Think about sin(x) integrated from 0 to 2π. The function is
positive some of the time and negative the rest and in perfect balance.
However, in our problem the volume of integration is arbitrary. I could
make the integration region bigger, smaller or move it to another place.
The only way the integral can always be zero is if the thing we are
integrating is zero. Everywhere. Therefore our integral equation can be
written in differential form as,
∂c
+ ∇ · j = 0. (5.4)
∂t
The law, which is not really a fundamental law, states there is a linear
relationship between diffusive mass flux and concentration gradients.
Concentration has units of number of molecules per unit volume. The
mass flux has units of number of molecules per unit area per unit time.
The diffusivity D has units of m2 /s. We treat D as a material property
that we can look up. Typically the constant is quite small if the medium
is a liquid. For a small molecule in water D ∼ 10−9 m2 /s. For larger
molecules it can be an order of magnitude smaller. It is important to
remember that Fick’s law is a constitutive law which is not necessarily
valid for non-dilute systems. For the system of random walkers, Fick’s
law turns out to hold true.
5.3 Equation in 1D
While we derived the equation in 3D, it is easiest to solve the equation
and gain some insight by focusing on a simpler 1D case. In 1D the
partial differential equation which relates changes in concentration with
respect to time and space is,
∂c ∂2c
= D 2.
∂t ∂x
66 Diffusive mass transfer
shown in Figure 5.6. However, the solution probably makes some intu-
itive sense based in your experience. Pay attention to the curvature in
the solution. The left half always has curvature downward and the right
half is curvature upward. The curvature corresponds on the right where
the concentration always increases and the left it always decreases. The
center point is an inflection point with zero curvature and the value of
the concentration there doesn’t change. Note that the final state is uni-
form where the concentration field is flat. The equation says that steady
state (when things can no longer change in time) is reached when the
concentration field is uniform or linear. The boundary conditions re-
quire the concentration field to be flat at the boundary, thus the final
state can only be uniform concentration.
5.3.1 Pi Theorem
Without solving the equation, let’s use the Pi theorem and see what
that tells us. In dimensional terms we want to find c(x, t, D, `, c0 ). To
introduce the concentration we need a new independent dimension, the
number of molecules, N . The table for our dimensionless concentration
and the systematic removal of units is,
2
c [ LN3 ] x [L] t [T] c0 [ LN3 ] ` [L] D [ LT ]
2
c
c0
x [L] t [T] ` [L] D [ LT ] remove N
c x D
c0 `
t [T] `2
[ T1 ] remove L
c x tD
c0 ` `2
remove T
did this for a few simple problems in our first chapter on dimensional
analysis. To proceed, we introduce a non-dimensional spatial coordinate
x̃ = x/` so that the domain extends from 0 < x̃ < 1. We can also define
a non-dimensional time as t̃ = t/t0 where t0 is an arbitrary (for now)
time scale. Under this change of variables, the equation becomes,
∂c ∂2c
=D . (5.8)
∂(t̃ t0 ) ∂(`x̃)2
Since ` and t0 are constant we can pull it outside the derivatives to
obtain
∂c Dt0 ∂ 2 c
= 2 . (5.9)
∂ t̃ ` ∂ x̃2
Nothing is stopping us from arbitrarily setting t0 = `2 /D so that the
diffusion equation becomes simplified with no material or geometric
parameters,
∂c ∂2c
= . (5.10)
∂ t̃ ∂ x̃2
It should be clear that since the left and right side of the equation
are proportional to the concentration, we can easily divide the overall
concentration by c0 . Normalizing the concentration is really no different
that working with molar concentration units rather than with the actual
number of molecules.
The dimensionless equation and boundary conditions show that the
solution has to have a form
c x tD
= f (x̃, t̃) = f ,
c0 ` `2
The functional form is exactly that obtained via the Pi Theorem.
While it may not seem so, these results are very powerful and useful
even though we haven’t solved a problem yet. The units of the equation
says all diffusion problems (with this boundary condition) are the same.
We simply scale the geometry to have a length that ranges from zero
to one. The combination of the size and the diffusivity gives us the
appropriate time unit. On the scaled domain and in the proper time
units, problems of different size and material property will have the
same solution. Regardless of the size or material involved, if you solve
the diffusion equation for this problem you have solved it once and for
all. Once we have our solution we simply use the definitions above to
70 Diffusive mass transfer
While I am not providing how I arrived at this form, you can check by
inspection that it in fact satisfies the partial differential equation. The
constants A, B, and λ are unknowns until we apply the two boundary
conditions and the initial conditions.
To show how we go about determining the constants, let’s consider
5.3 Equation in 1D 71
the particular case where the two boundary conditions are insulating.
The boundary conditions were stated as,
∂c
= 0.
∂x x=0,1
While we can plot the solution to see what it looks like, let’s think
about it a little more. First, note that all the terms with even values
of n would be zero due to the sin(nπ/2) part of the equation. Now
imagine carrying out the sum for the first few odd values of n,
1 2 −π2 t 2 2 2 2
c(x, t) = + e cos(πx)− e−9π t cos(3πx)+ e−25π t cos(5πx)+...
2 π 3π 5π
Note in the above solution that for n > 1, once a little time has passed
the coefficient in front of these terms would be really small. Imagine
the solution at t = 1/π 2 ≈ 0.1. At this time e−9 << e−1 . It should be
clear that once t is no longer small then the solution is dominated by
the first cosine term,
1 2 2
c(x, t) ≈ + e−π t cos(πx)
2 π
When you look at Figure 5.6, notice that not much time passes before
you see this single cosine term in the solution.
5.4 Role of simulation 73
If this is the first time you have come across this type of solution to
a partial differential equation, the method probably seems mysterious
at this point. More detail on the origins and meaning of this solution
as well as examples for different boundary conditions can be found in
my other resources.
5.5 Summary
Mass transfer by diffusion is generally a pretty slow process. If you
consider something like sugar mixed in water, the diffusivity is D =
6 × 10−10 m2 /s. We discussed that you could estimate the time it takes
diffusion to occur over some distance as t ∼ `2 /D. Using this argu-
ment and assuming your coffee cup is about ` ∼ 10 cm then it would
take about half a year for sugar to mix within your coffee cup. If you
wanted the molecules to diffuse over a length scale of your height, you
might need to wait 200 years. Diffusion is slow at the human scale. If
you consider a single cell, which might have a size of 10 microns (1
micron is 10−6 m), then we get diffusion times scales of on the order
of a second. Diffusion is not so slow if you happen to be really small.
When we observe fluid mixing in our everyday life, it is most certainly
dominated by convection. In the next chapter we will begin to describe
conservation laws in moving fluids.
6
Convective mass transfer
a fixed amount of fluid. We then watch where this fluid blob goes as
it flows and deforms. The fluid particle is considered to be composed
of the same material for all time. This consideration is not literal from
a molecular point of view, it is surely true that the random motion
of molecules means that individual molecules are exchanged from one
fluid particle to the next.
6.3.1 Derivation in 1D
Let’s start with a simplified example. Imagine the pollutants in a par-
ticular region of the river follow a linear and steady concentration field
that only varies in the downstream coordinate x. If you walk along
the river in this location and keep taking measurements of the concen-
tration field from the river bank you find a region that concentration
increases linearly, Figure 6.1. For this example, the concentration field
at the location of change is
x
c(x) = c0 + ∆c ,
L
where ∆c is the increase in concentration over distance L. The slope of
the concentration field in this region is constant,
dc ∆c
= .
dx L
c ∆c
We can also get the same result appealing to the chain rule. The boat
follows a simple path,
xboat (t) = xo + ut
The stationary concentration field, c(x) would be seen from the boat
as a function of time by c(xo + ut). Taking the derivative with respect
to time using the chain rule yields,
dc(x(t)) dc dx dc
= =u
dt dx dt dx
particle path
t=0
y
a
r(a,t)
where x(a, t) is the first component of the vector r. While this mapping
seems confusing written as an equation, intuitively the idea is simple. If
I am on the bank of the river sampling the concentration at a particular
point, and you pass through that point on your raft, at the instant you
pass through my point, our measurements must agree.
The derivative with respect to time for a fixed fluid particle is found
6.4 Liebniz and Reynold’s transport theorem 81
where ρ(x, t) is the fluid density field. The subscript V (t) reminds us
that the volume of integration is not fixed in time and may be changing
shape as the flow deforms the material. Our volume V (t) is going with
the flow and following the material for all times. Since our volume
of integration was taken to be a material volume (i.e. it encloses the
6.5 Conservation of mass: the fluid 83
same amount of material at all times) and since mass is conserved then
dM
dt = 0. Using the Transport Theorem (Equation 6.3),
Z Z
dM d ∂ρ
= ρdV = + ∇ · (ρv) dV = 0
dt dt V (t) V (t) ∂t
Since the volume in our analysis was arbitrary and since the integral is
always equal to zero then the the integrand must be zero,
∂ρ
+ ∇ · (ρv) = 0. (6.4)
∂t
This equation must be true point-wise in the fluid and is the differential
statement of mass conservation. By rearrangement,
Dρ
+ ρ∇ · v = 0. (6.5)
Dt
This equation says that the density of a fluid particle can only change
if the velocity field has non-zero divergence. Divergence of the velocity
field is a measure of the net rate of fluid flowing into/out of a point.
If the velocity vectors are all pointing to one location, then the flow
must be accumulating mass at that point and the density should be
increasing. If the velocity vectors all point away from a point in the
flow, the density should be decreasing.
For an incompressible flow with constant density we have the simpli-
fied relation,
∇ · v = 0.
This assumption of incompressibility is one that we will use throughout
the course. For a liquid the assumption of incompressibility is a good
one. I challenge you to compress some water to any significant degree.
For a gas the assumption is questionable. It turns out that a rule of
thumb people use is that if the flow velocity is less than about 1/3
of the speed of sound in a gas, then the assumption of incompressible
flow is basically fine. For air this means we can safely treat cases with
flow velocities on the order of 100 m/s or less as an incompressible flow.
Flow around a car is incompressible, but flow around an airplane is not.
Compressible flows have many interesting features that will be beyond
the scope of our first course.
It is also important that we refer to a flow as incompressible and not
the fluid. For the flow of air, we have many examples of incompressible
flow despite the fact that air is quite compressible.
84 Convective mass transfer
The subscripts V (t) and S(t) remind us that the volume of integration
is not fixed in time and is changing shape as the flow deforms the carrier
fluid. Using the Transport Theorem (Equation 6.3) for the left side of
the equation and the divergence theorem for the right side we obtain,
Z Z
∂c
+ ∇ · (cv) dV = − ∇ · jdV
V (t) ∂t V (t)
We can group all the integrands together and since the volume in our
analysis was arbitrary the integrand must always be zero,
∂c
+ ∇ · (cv) = −∇ · j.
∂t
Substituting in Fick’s law from the previous chapter and assuming a
constant diffusivity,
∂c
+ ∇ · (cv) = D∇2 c. (6.6)
∂t
Now, a new term has appeared in the equation ∇·(cv) which represents
convection of the dye molecules by fluid motion.
Let’s think about this equation for a minute. In the case where there
is no fluid motion, v = 0, we are back to the classic diffusion equation.
In another limit, let’s assume there is no molecular diffusion. As we
discussed earlier in the chapter, molecular diffusion is very slow and so
in many cases ignoring it is not an unreasonable assumption. In that
case the equation would be,
∂c
= −∇ · (cv).
∂t
The product cv is a vector which gives the convective flux or rate that
dye is carried by the fluid per unit time, per unit area at a point. If
you recall that c has units of number of molecules per unit volume,
you can check the units of cv yourself. Now recall that the divergence
6.5 Conservation of mass: the fluid 85
∂c
+ v · ∇c + c∇ · v = D∇2 c.
∂t
Substituting in the conservation of mass from the previous section we
can replace the divergence of the velocity field as,
∂c c Dρ
+ v · ∇c − = D∇2 c,
∂t ρ Dt
Dc c Dρ
− = D∇2 c. (6.7)
Dt ρ Dt
a) b)
c) d)
We can expand the derivatives on the right side to write this expression
as
Z Z
d ∂ρ ∂b
ρbdV = b + ∇ · (ρv) + ρ + v · ∇b dV.
dt V (t) V (t) ∂t ∂t
This form is convenient since the first term in parenthesis is zero from
mass conservation, Equation 7.2. Our final result is then,
Z Z Z
d ∂b Db
ρbdV = ρ + v · ∇b dV = ρ dV, (7.3)
dt V (t) V (t) ∂t V (t) Dt
where ρ(x, t) is the density field and v(x, t) is the velocity vector field.
Note that this is a vector expression as the momentum has an x, y,
and z component. In Newton’s law we know that we need to be able
to take the time derivative of momentum. To take the time derivative
wef can use Equation 7.3, simply replacing b with v,
Z Z
dp d Dv
= ρvdV = ρ dV (7.4)
dt dt V (t) V (t) Dt
This form might make you uncomfortable since this is a vector equa-
tion and we only discussed the transport theorem with scalar functions.
If it does you could proceed with only the x component of the momen-
tum instead,
Z Z
dpx d Du
= ρudV = ρ dV.
dt dt V (t) V (t) Dt
Dv ∂v
= + v · ∇v.
Dt ∂t
is a vector,
u ∂u ∂u ∂u
∂x + v ∂y + w ∂z
∂v ∂v ∂v
v · ∇v =
u ∂x + v ∂y + w ∂z ,
u ∂w ∂w ∂w
∂x + v ∂y + w ∂z
where the first, second, and third rows are the x, y, and z components
of the vector.
The expression 7.4 allows us to compute the rate of change of mo-
mentum of a material volume. Newton’s laws requires that this vector
must equal the sum of the forces acting on the control volume. Deter-
mining the forces acting on the fluid is not a simple matter and will
take some work. Before we turn to computing forces, let’s make sure
we really understand the idea of acceleration in fluid flows.
U(t=0) U(t=dt)
dx
=Udt
x=0 x = dx
to the velocity field changing in time and the other is from the velocity
field changing in space.
We can see that the spatial acceleration must follow a form like v·∇v
in a simple 1D example. Imagine a smooth river with a constant flow
in time. The river has a constant depth but there is a region where the
banks narrow down in width. Since the cross sectional area of the river
has been reduced, the flow velocity must increase in the narrow region,
shown in Figure 7.1.
Now let’s measure the acceleration of the fluid from the perspective
of the flow and the fixed river banks. Let’s assume that I am on a raft
and you are on the river bank. We are going to measure the acceleration
that a raft feels at the same location from our two different perspectives.
Even though we are observing from different points of view, in the end
our measurements should be the same. If I go right down the center of
the river on a raft, I sense the x component of the river flow velocity.
If I want to know my acceleration, I simply measure the raft velocity,
U , at two instances, t = 0 and a short time later, t = dt;
Uraft (t = dt) − Uraft (t = 0)
Acceleration =
dt
Now you are observing from the river bank and want to know the
fluid acceleration at the same point. If you watch me go past you could
measure my acceleration the same way. However, assume there is no
94 Conservation of momentum in a fluid
tracer for you to observe in the flow and you are stuck taking point
measurements of the river’s flow velocity. The previous formula for
acceleration is still valid, however from the bank you can only measure
the river velocity as a function of position. The equivalent measurement
of the fluid acceleration would would be,
u(x = dx) − u(x = 0)
Acceleration = .
dt
The distance dx that I would have moved on the raft over time dt is
related to the flow velocity as dx = u dt. Therefore substituting for dt,
u(x = dx) − u(x = 0)
Acceleration = u .
dx
Now the fraction in this equation is the definition of the spatial deriva-
tive of the flow velocity in the limit that dx goes to zero,
∂u
Acceleration = u .
∂x
The expression above is simply the 1D version of v·∇v, the acceleration
of a fluid particle due to changes in the velocity field in space.
7.3 F = m a
In particle mechanics, we usually think of Newton’s Law expressed as
the equation F = m a. However, this is for a constant mass particle
and a more general statement would be
dp
= F,
dt
where p is the linear momentum and F is the sum of the forces acting
on the particle. So we have developed a way to describe the rate of
change of the momentum for a fluid particle. Now we need to work
on the forces. We will account for body forces acting on the whole of
the material volume, fb and surface forces fs acting on the boundary
separately. Generally then, conservation of momentum can be stated
as,
Z
Dv
ρ dV = fb + fs . (7.5)
V (t) Dt
7.3 F = m a 95
We could also consider electrostatic forces (if the fluid were charged
and subjected to an electric field) or Lorenz forces (if the fluid were
passing a current and subject to magnetic fields). These are interesting
topics for another day, however their inclusion into our theory would
not be difficult. If we know how to calculate the force per unit volume,
we only need to integrate this force density over our volume.
Figure 7.2 shows a schematic where we need to add up, or integrate, all
the little surface forces acting over the entire material volume’s surface.
The general description of the state of stress within the fluid requires
us to introduce the idea of a tensor. A tensor has a structure like a ma-
trix. In three dimensions, a tensor is a 3x3 matrix with 9 components. A
tensor is needs to describe the state of stress in a material as there are 9
things that are important. The stress at a location has magnitude and
direction; it is a vector with 3 components. However, what happens to
the material under this force depends upon which face the stress acts.
A stress with only an x component which acts on the surface with the
96 Conservation of momentum in a fluid
s
s
s dS
s
s
s
Figure 7.2 The local stress vector, s, acts everywhere on the surface,
S, of an arbitrary region. Integrating that stress over the surface area
gives the net force acting on the surface of the volume.
Txz
7.3 F = m a 97
Tyy
Tyx
Tyz
Txy
Tzy
Txx
Tzx Txz
Tzz
T(x, y, z, t).
s=n·T
The dot product projects the tensor onto the surface with a normal
vector n to get the stress vector at that point on the surface. The dot
product of two vectors gives a scalar, while the dot product of vector
and a tensor gives a vector. To remember what n · T means, it is like
98 Conservation of momentum in a fluid
By the divergence theorem, which also works for tensors just as it does
for matrices,
Z Z
fs = n · TdS = ∇ · TdV. (7.7)
will compactly describe the internal state of stress of the fluid. That’s
all. This approach is not limited to the mechanics of fluids, but works
equally well in solid mechanics. In solid mechanics, we are often only
concerned with equilibrium and thus the commonly used equilibrium
statement in solid mechanics that ∇ · T = 0.
7.3.3 Meaning of ∇ · T
Above we used the definition of the stress tensor and our vector calculus
to show that the net surface force acting on an arbitrary volume is the
divergence of the stress tensor. We can get the same result using a free
body diagram for a infinitesimal element of fluid. Consider a small cube
that is dx, dy, and dz on each side. Consider the point P in the middle
of the cube. To find the forces on the face of the cube in the x-direction,
we must consider that the stress tensor varies in space.
The value of the stress at one point is related to the value at a
nearby point through a Taylor series. Don’t groan. Students always
groan at Taylor series. Taylor series is your friend. Taylor series says
any complicated crazy function, is just a line as long as you zoom in
enough. The Taylor Series expansion of the stress tensor on the right
face of the square relative to the center is,
dx ∂Txx
Txx + .
2 ∂x
Here Txx is the component of the stress tensor at the center of the cube.
Likewise, the stress on the left hand face of the cube in the x-direction
is,
dx ∂Txx
Txx − .
2 ∂x
Tyx+ ∂Tyx dy
∂x 2
dy
y
x
Txx- ∂Txx dx Txx+ ∂Txx dx
∂x 2 ∂x 2
dx
Tyx- ∂Tyx dy
∂x 2
have
dx ∂Txx dx ∂Txx
dfsx = Txx + dydz − Txx − dydz +
2 ∂x 2 ∂x
dy ∂Tyx dy ∂Tyx
Tyx + dxdz − Tyx − dxdz +
2 ∂y 2 ∂y
dz ∂Tzx dz ∂Tzx
Tzx + dxdy − Tzx − dxdy
2 ∂z 2 ∂z
Which reduces to
∂Txx ∂Tyx ∂Tzx
dfsx = + + dxdydz.
∂x ∂y ∂z
Similarly, we would find
∂Txy ∂Tyy ∂Tzy
dfsy = + + dxdydz,
∂x ∂y ∂z
∂Txz ∂Tyz ∂Tzz
dfsz = + + dxdydz.
∂x ∂y ∂z
Using vector notation,
df s = (∇ · T)dV.
Thus the total force is Z
fs = ∇ · TdV.
and we get the same result as the previous section. The net surface
7.4 So what’s a tensor? 101
Since all the integrals are volume integrals over any arbitrary volume
in the fluid, the equation holds at every point,
Dv
ρ = ρg + ∇ · T. (7.9)
Dt
While this looks like a beautifully simple equation, there is a seri-
ous problem. There are too many unknowns. The equation is actually
three equations for the 3 unknown components of the velocity vector.
However, there are 9 components of the stress tensor. If we considered
conservation of angular momentum (we will skip the derivation) we
would find that the stress tensor must be symmetric. Thus there are
only 6 unknown components of the tensor. This is a better situation
but still 6 equations short.
What is needed to close the problem is the same type of relationship
as Fourier’s law provided when we studied heat conduction. We need a
law that relates the unknown forces to the deformation of the material.
This is called a constitutive law and cannot be consider more than an
empirical relationship for the material of interest. The expression in
terms of the stress tensor is general and does not make any assumption
other than that of continuum. It works for fluids, solids, and other
crazy materials. The choice of constitutive law determines whether we
are working with fluids, solids, or something in between.
Dρ
= −ρ∇ · v,
Dt
Dv
ρ = ρg + ∇ · T.
Dt
These equations say nothing because we have not defined the stress ten-
sor, T. Defining the stress tensor requires knowledge of the material.
Its definition is essentially empirical or comes from a molecular model
of the material. The stress tensor is not a fundamental law of nature.
The situation is just like the case of heat conduction where our funda-
mental conservation law was written in terms of the heat flux vector
and Fourier’s law was needed to close the problem. The relationship
which defines the stress tensor is our constitutive law.
There are some constraints on what form the tensor must take. For
example, we stated in the previous chapter (without proof) that it is
symmetric due to conservation of angular momentum. It also can only
take limited forms if we want its definition to remain constant under
a change of coordinate systems (we always want this) or if we want it
to be isotropic (we often, but not always want this). We’ll save these
topics for a more advanced course on continuum mechanics and not
discuss them further.
104 The Navier Stokes equations
0 0 −P
∂P ∂P ∂P
=− , , = −∇P
∂x ∂y ∂z
Substituting this expression into our conservation law would yield con-
servation of momentum to be,
Dv
ρ = −∇P + ρg.
Dt
Now we have 4 equations (1 mass and 3 momentum) and 5 unknowns,
the density, velocity vector, and the pressure. This is a bit better but
still a problem. The final equation to close the problem must come from
thermodynamics, which relates temperature, pressure, and density. To
completely define the problem we would need to introduce conservation
of energy to our formulation, which we consider energy later.
However, if the flow is incompressible, we do not need to consider
energy or thermodynamics. For incompressible flow, the density is con-
stant and taken as a material property. For an incompressible flow,
Euler’s equations reduce to,
∇ · v = 0, (8.1)
Dv
ρ = −∇P + ρg. (8.2)
Dt
8.1 Euler’s equation 105
The terms with the velocity gradients is a little more complicated and
will be the subject of much discussion later in this chapter. Since ve-
locity is a vector, it’s gradient is a tensor. The velocity gradient tensor
is written out as,
∂u ∂v ∂w
∂x ∂x ∂x
∂u ∂v ∂w
∇v =
∂y ∂y ∂y
∂u ∂v ∂w
∂z ∂z ∂z
∇ · T = −∇P + µ∇2 v.
Equations 8.5 and 8.6 represent four equations for four unknowns; the
three components of velocity and the pressure. These equations are
known as the incompressible Navier Stokes equations and comprise the
mathematical foundation for describing fluid flow.
Just to be clear on the notation, if we expand the momentum equa-
tion in component form for a Cartesian coordinate system we obtain,
2
∂ u ∂2u ∂2u
∂u ∂u ∂u ∂u ∂P
ρ +u +v +w = ρgx − +µ + +
∂t ∂x ∂y ∂z ∂x ∂x2 ∂y 2 ∂z 2
curvature in the velocity field. While the full Navier Stokes equations
have other more complicated terms, we can think that viscosity will
always act to smooth the velocity field and pull the velocity to some
uniform or simple state - just like the heat equation.
The ratio of µ/ρ appears often that it is given its own symbol,
ν = µ/ρ. This parameter is called the kinematic viscosity and has units
of m2 /s. The material property, µ, is properly called the dynamic vis-
cosity. The kinematic viscosity plays a role like the thermal diffusivity,
α, or the diffusion coefficient D. In thermal problems, α, sets the rate
that heat propagated over some distance; the kinematic viscosity plays
a similar role. The kinematic viscosity sets the rate that momentum is
transferred from one sliding fluid layer to another.
u
n = [1 ]
0
y
x
The result says that the x component of the stress is given by the
horizontal velocity gradient, just as we described when defining a New-
tonian fluid. This is the force that would cause a fluids viscosity to drag
a solid along with the flow. The y component of the stress is given by
the pressure.
U(XA) U(XB)
t=0
XA(0) XB(0)
t
x
The velocity gradient, ∂u∂x , provides the linear strain rate that a ma-
terial line elongates or shrinks. This makes sense. Imagine you and a
friend are in hot air balloons up above the earth. Since the length scale
of the earth is so big we can take dx at the human scale to be infinites-
imal. If the distance between you and your friend is growing, it means
there are velocity gradients in the flow. Your friend, even though close
by, is subjected to a different fluid velocity than you, thus you separate
from each other. The rate that you separate is a measure of the flows
velocity gradient.
dt UA
A
α
dt VA t=dt
A
β
o B
t=0 dt VB
dt Vo
o B
y dt Uo
dt UB
x
Using these relations and a little bit of geometry, you can find the
tangent of the angles β and α in terms of the velocities at points A, B,
and O as,
(vB − v0 )dt
tan(β) = ,
dx + (uB − u0 )dt
(uA − u0 )dt
tan(α) = .
dx + (vA − v0 )dt
You can then use the Taylor Series to relate the velocities at point
A and B to that of point O, i.e. uB = u0 + ∂u∂x dx. Substituting the
8.5 Comments on kinematics 115
Taylor Series results and taking the limit, these geometric relationships
become rate equations for the rate of change of the angles,
dβ ∂v dα ∂u
= ; = .
dt ∂x dt ∂y
Note that in the limit of small angles which occur over small times,
tan(α) ≈ α. Thus the sum of these two rates,
dβ dα ∂v ∂u
+ = + ,
dt dt ∂x ∂y
is the rate the angle between the two lines OA and OB close or open
up. This is the shear strain rate.
This relationship makes sense. Hold your hands together in front of
you, pointing upwards. Sliding your hands up and down relative to each
other creates shear. The right hand has a different vertical velocity than
∂v
the left hand. Therefore ∂x is not zero. Holding your hands horizontal
creates velocity gradient as ∂u/∂y. The shearing motion of your hands
would act to deform a material held between them.
8.5.3 Rotation
Using the results from the prior section we can also ask what is the
difference of the rate of change of the two angles. This difference is
twice the rate of solid body rotation of the line elements at that point,
dβ dα ∂v ∂u
− = − .
dt dt ∂x ∂y
In two dimensions, you may recognize this quantity is the same as the
curl of the velocity, which is referred to in the fluid mechanics world
as vorticity, ω. The curl is denoted by ω = ∇ × v. In component form,
the vorticity vector is
∂w ∂v
∂y − ∂z
ω = ∇ × v = ∂u ∂w
∂z − ∂x
.
∂v ∂u
∂x − ∂y
The curl is the vector cross product of the gradient operator ∇ and the
velocity vector v. In two dimensions where the velocity is constrained
to the x-y plane, the vorticity only has a component normal to this
116 The Navier Stokes equations
8.5.4 Generalized to 3D
In three dimensions, the results of this section are summarized as fol-
lows. The strain rate tensor is defined as
1
∇v + ∇vT .
S= (8.8)
2
In component form this tensor is
2 ∂u
∂x
∂u
∂y + ∂x
∂v ∂u
∂z + ∂x
∂w
1
S = ∂u ∂y + ∂x
∂v ∂v
2 ∂y ∂v ∂w
∂z + ∂y
2 ∂u ∂w ∂v ∂w
∂z + ∂x ∂z + ∂y 2 ∂w
∂z
The strain rate tensor is symmetric and provides the rate of strain
of fluid elements in the different directions. The diagonal components
of the tensor are the linear strain rates in the three dimensions. The
off-diagonal components are the shear strain rates in the three planes;
x − y, x − z, and y − z.
To rotation tensor is defined as
0 −ωz ωy
1
R = ωz 0 ωx (8.9)
2
−ωy −ωx 0
where ω are three components of the vorticity vector, ω = ∇ × v.
The off diagonal elements of the rotation tensor provide the amount
of rotation in the three planes x − y, x − z, and y − z. The tensor is
anti-symmetric. It is fairly easy to see that the velocity gradient is the
sum of the strain rate and the rotation tensors,
∇v = S + R.
Velocity gradients lead to both strain of material elements and rotation
of those elements. The strain rate tensor provides the rate of strain of
the material. The rotation tensor provides the amount of rotation.
8.6 Non-dimensionalization 117
Using the above tensors, the Newtonian constitutive law for an in-
compressible flow states that
T = P I + 2µS.
The forces exerted on the fluid are proportional the strain rate, but not
the rotation of the fluid.
Please note that this discussion of kinematics is quite brief and more
detailed derivations can be found in the references. The main point is
to demonstrate that the form of the Newtonian stress tensor is con-
nected to the kinematics of rate of deformation of the material. The
stress is related to the velocity gradients. However, the use of the sym-
metric strain rate tensor subtracts our solid body rotation which does
not generate stress. In a Newtonian fluid only the rate of deformation
generates stress.
8.6 Non-dimensionalization
Consider the Navier-Stokes equations for incompressible constant den-
sity flow,
∂v
ρ + v · ∇v = −∇P + µ∇2 v,
∂t
∇ · v = 0.
Let’s say the geometry of the problem provides some characteristic
length, L, and some characteristic flow velocity, U0 . The equations can
be made dimensionless following the same procedure as in previous
chapters. Let’s define ṽ = v/U0 , [x̃, ỹ, z̃] = [x, y, z]/L, t̃ = tU0 /L,
and P̃ = P/P0 . Our time scale is picked so that it takes one time
unit for a fluid particle to travel the length L; i.e. our dimensionless
time is t̃ = tU/L. When making equations dimensionless, it is always
important to recall that derivatives have units. The derivative, ∂/∂x has
units of 1/Length, for example. Therefore, when we make our operator
∇ dimensionless, we make the substitution ∇ ˜ = L∇. The dimensionless
˜
version of the Laplacian is given as ∇ = L2 ∇2 . Making this change of
2
is, lets consider a few hypothetical cases. For water, the density is ρ =
1000 kg/m3 and the viscosity is approximately µ = 0.001 Ns/m2 . So
lets say a person 2m tall who can swim at 1 m/s would have a Reynolds
number of 2 × 106 . A small organism that is 100 µm and can swim one
body length per second would have a Reynolds number of 0.01. For
you to match the Reynolds number of this organism you would need to
swim at about the same speed (100 µm/s) in molasses!
In a pipe flow, a standard rule of thumb is that the flow will transition
from laminar to turbulent at a Reynolds number of 2300. Here the
Reynolds number is defined on the diameter of the pipe. To get a sense
of where a Reynolds number of 2300 is - water flowing in a 1/4 inch pipe
(6.35 mm) at a velocity of 1 ft/s (0.3 m/s) we would be at a Reynolds
number of about 2000. The transition number is only a rule of thumb. If
you try this experimentally you will find the transition number changes
based on how carefully you set up the experiment. If you just hook up
some tubing and push water through it without worrying about keeping
everything smooth, straight, and well conditioned, you will probably
observe a transition at a much lower Reynolds number, probably under
1000. If you are extremely careful and take great pains to do a perfect
experiment the flow will transition at a much high Reynolds number. In
2011, the best theoretical determination of a true transition Reynolds
number to sustained turbulence was determined to be Re = 2040. While
a seemingly simple problem, the details are quite complex.
Reynolds original 1885 experimental apparatus survived at the Uni-
versity of Manchester, where he was a professor. A century later, his
exact experiments were recreated with his apparatus. The experiment
showed a transition Reynolds number much lower than he reported.
The difference was found to be due to cars in modern day Manchester,
the small amount of shaking from the streets perturbed the simple but
delicate experiments.
8.8 Summary
The starting point for much of our discussion in future chapters are the
Navier-Stokes equations for incompressible constant density flow,
∂v
ρ + v · ∇v = ρg − ∇P + µ∇2 v,
∂t
120 The Navier Stokes equations
∇ · v = 0.
Mathematicians, or people who have very mathematical minds, are often led
astray when “studying” physics because they lose sight of the physics. They
say: “Look, these differential equations - the Navier-Stokes equations - are
all there is to fluid dynamics; it is admitted by the physicists and engineers
that there is nothing which is not contained in the equations. The equations
are complicated, but after all they are only mathematical equations and if
I understand them mathematically inside out, I will understand the physics
inside out”. Only it doesn’t work that way. ... They fail because the actual
physical situations in the real world are so complicated that it is necessary to
8.8 Summary 121
x
L
Figure 9.1 Schematic for Poiseuille flow between two parallel plates.
solve some simple problems. In all the examples that follow we will
study flows where there is no acceleration and force due to viscosity
will balance some driving force.
+ ∂v = 0.
S∂u
∂x
S ∂y
S
Note that we keep the axial pressure gradient as the pressure drives the
flow and must drop continuously as we move down the channel.
From conservation of mass, ∂v/∂y = 0; an equation which states that
the vertical velocity is a constant in the y direction. Since v = 0 at the
wall, v = 0 everywhere. We can now cancel out all terms that have the
vertical velocity, v in them;
!
2 2
∂u ∂u ∂u ∂P S∂ u ∂ u
ρ S +u
@ +v =−
S +µ +
2
∂t
S
S ∂x
@@ ∂y SS ∂x ∂xSS ∂y 2
!
2 2
∂v ∂v ∂v ∂P ∂
S v S∂ v
ρ A +u
@ +v =−
S − ρg + µ +
2 S2
∂t
AA ∂x
@ ∂y
@ SS ∂y ∂xSS
∂y S
It makes sense that the left side of the Navier Stokes equations is zero.
The left side represents the acceleration of a fluid particle. There is
no acceleration in this system. There is neither acceleration due to
unsteadiness (change in velocity with respect to time) nor from the
perspective of a fluid particle. The channel is like a calm river whose
width doesn’t change; if you were on a raft you would translate at
constant velocity. Neither the observer on the raft or on the banks
would see any acceleration in this flow.
Thus, the momentum equations are simplified significantly to
∂P ∂2u
= µ 2. (9.1)
∂x ∂y
∂P
= −ρg. (9.2)
∂y
Integration of the second equation yields,
P = P (x, y = 0) − ρgy.
where P (x, y = 0) is the pressure along the lower wall, which is a func-
tion of x only. When we take the pressure gradient in the x direction,
the hydrostatic component (the term with ρg) does nothing to modify
126 Solutions to the Navier-Stokes equations
the axial flow. Since the flow was assumed invariant in x, the only con-
sistent solution would be that the axial pressure gradient, dP/dx, is a
constant. Since the left side of Equation 9.1 is a constant the equation
can be easily integrated twice with respect to y,
dP 1 2
u(y) = y + C1 y + C2 .
dx 2µ
We are left with two constants of integration which are determined by
the no-slip boundary conditions at both walls. For the lower wall we set
u(y = 0) = 0. Substituting in this boundary condition states that the
constant C2 must be zero. The upper wall condition of u(y = H) = 0
allows us to determine C1 . The final result is,
dP 1
u(y) = − y(H − y).
dx 2µ
The velocity profile follows a simple parabola which is zero at y = 0
and y = H. Recall from our discussion on diffusion that the second
derivative in space is equivalent to the curvature. Therefore, the gov-
erning equation states that the curvature of the x component of the
velocity field is proportional to a constant (the pressure gradient). The
parabola is the only function which has constant curvature, thus we
could obtain the qualitative shape of the velocity profile without even
solving the equation. RH
The total flow, per unit width is Q = 0 u(y)dy. Integrating our
velocity profile gives the total flow as
dP H 3
Q=−
dx 12µ
Usually we use the fact that dP/dx is the same as the total applied
pressure difference divided by the total length of channel. We need to
be careful with the signs. Convention would usually define a positive
pressure difference, ∆P , as going from high pressure on the inlet to low
pressure on the outlet. With this convention, dP/dx = −∆P/L. Using
the total pressure drop notation gives,
12µL
∆P = Q .
H3
The expression says that pressure and flow rate are linearly related
- double the pressure and you double the flow. This expression is a
hydraulic version of Ohm’s law (V = iR) where pressure difference
9.1 Flow between parallel plates - Poiseuille flow 127
acts like voltage and volumetric flow rate acts like current. The number
12µL/H 3 is the hydraulic resistance.
There are a few things we should check with our resistance formula.
First, we can check the units to make sure they agree with what we
expect and that we did not make an error. The other check is that
the trends all go in the right way. It makes physical sense that the
resistance increases with viscosity, channel length, and with a decrease
in the channel height. It is always good to look at a result and see if
you believe the basic facts about that result before proceeding.
Note that the mean flow velocity is ū = Q/H,
∆P H 2
ū =
12Lµ
and the maximum flow velocity (along the centerline) is found by eval-
uating the velocity profile at y = H/2,
∆P H 2
H 3
u y= = = ū.
2 8Lµ 2
Since there is no acceleration in this flow we can check our result
with a simple force balance If we draw a box (control volume) the net
force (per unit depth into the page) exerted on the box due to pressure
is
FP = ∆P H.
This force must be balanced by the shear stress exerted at the two
walls. Recall the method of finding the stress on a surface discussed
in the chapter on the Navier-Stokes equations. The shear stress acting
tangential to the wall is given by n·T where T is the total stress tensor.
The normal vector for the lower wall is n = [0 1]. The stress tensor at
the wall for this simplified flow is
" #
T
−P µ ∂u ∂y
T = −P I + µ ∇v + ∇v = .
µ ∂u
∂y −P
u(y) H
x
U
Figure 9.2 Schematic for Couette flow between two parallel plates.
example and cross out all the time derivatives and x derivatives. We can
then apply conservation of mass to tell us there is no vertical velocity,
v. Crossing out all these terms again yields the x and y momentum
equations as in Equations 9.1 and 9.2,
∂P ∂2u
= µ 2.
∂x ∂y
∂P
= 0.
∂y
However, in this case the x momentum equation simplifies further as
there is no applied pressure gradient. In the example of the concentric
cylinders there could be no pressure gradient around the closed loop
of the fluid gap, otherwise the pressure would be discontinuous as we
went around the circle. For Couette flow, the x momentum equation
simplifies to,
∂2u
0= .
∂y 2
The curvature of the velocity field is zero and thus we expect a lin-
ear velocity profile. Integrating this expression twice with respect to y
yields,
u(y) = C1 y + C2 .
∂P
= −ρgcos(θ).
∂y
Integrating the second equation across the thickness of the film and
using the boundary condition that the pressure at the surface of the
film is that of the air around, P∞ , we obtain the pressure everywhere.
P = P∞ + ρgcos(θ)(H − y)
132 Solutions to the Navier-Stokes equations
The pressure does not depend on x, thus the x momentum equation is,
∂2u
−ρgsin(θ) = µ .
∂y 2
Integrating twice yields (since the left side is constant)
ρgsin(θ) 2
u(y) = − y + C1 y + C2 .
2µ
The two constants of integration are found from the boundary con-
dition. At the surface of the solid ramp, the no-slip condition holds,
u(y = 0) = 0. Applying this boundary condition shows that C2 = 0.
Therefore,
ρgsin(θ) 2
u(y) = − y + C1 y.
2µ
The second boundary condition is that there is no shear stress at the
free surface of the liquid. Nothing is there (other than a little air) to
exert a force on the upper surface of the free fluid film. Thus we need
to set ∂u/∂y = 0 at the upper surface y = H. Applying this condition
yields,
ρgsin(θ) y
u= y H− .
µ 2
You should always confirm at the end that the solution satisfies the
equation and the boundary conditions.
Also note that the solution here is half a parabola - or essentially one
half of the flow between two solid plates. The driving force in Poiseuille
flow is pressure whereas here it is gravity. Otherwise, the solutions look
very similar.
dP 1 2
u(y) = y + C1 y + C2 .
dx 2µ
u(y = 0) = U = C2 .
dP 1 2 U dP 1
u(y = H) = 0 = H + C1 H + U → C1 = − − H.
dx 2µ H dx 2µ
dP 1 H −y
u(y) = − y(H − y) + U .
dx 2µ H
If you look at the velocity profile and total flow rate derived for
Poiseuille and Couette flow independently, you will see the above result
is just the superposition of the two solutions. We could have guessed
this result from the start since under our assumptions the equations are
linear. When you have linear equations you can compose a solution to
a problem as the sum of other solutions. Superposition is a useful tech-
nique for solving more complex flow problems as we will demonstrate
in a few of the coming examples.
134 Solutions to the Navier-Stokes equations
High Pressure
L1 L2
Figure 9.4 Schematic for sliding block moving to the left over a
stationary surface. In the reference frame of the block the lower
wall appears to move to the right at constant velocity. While the
gap size is exaggerated here, the solution assumes the gap is very
thin relative to the length of the block.
Figure 9.4. If the pressure is high at the step then the flow rate will be
enhanced in region 2 and decreased in region 1.
The velocity profile in each region can be thought of as the superposi-
tion of Couette and Poiseuille flow. Using the results derived previously,
the total flow rate in region 1 will be
dP H13
U H1
Q1 = − + ,
dx 1 12µ
2
and similarly for region 2. Overall conservation of mass would state
that Q1 = Q2 ,
dP H13 dP H23
U H1 U H2
− + =− + .
dx 1 12µ
2 dx 2 12µ
2
The expression gives us the relationship between the two pressure gra-
dients in regions 1 and 2. To close the problem, we need additional
information. We need to know what is the overall applied pressure
across the entire fluid gap. Let’s assume the upper slider is open to the
fluid through which it moves. Therefore, the pressure at the two ends
would be the same and there is no overall applied pressure. Since only
pressure differences matter, we can take the far left and right ends of
the slider to be zero pressure. If P is the pressure at the step then the
pressure gradient in region 1 would be P/L1 and the pressure gradient
in region two is −P/L2 . The difference in sign is because the pressure
rises from 0 to P in region 1 and falls from P to 0 in region 2. We now
have enough information to solve for the pressure at the step, P ,
1 H13
1 6µU
P + 3 = (H1 − H2 ).
L2 L1 H2 H23
Notice that when H1 = H2 the pressure is zero and we are back to
Couette flow. Also notice that the pressure under the slider is positive
when H1 > H2 and negative when the situation is reversed.
In order to understand the result a little easier, lets explore the case
where L1 = L2 = L/2 and H2 = H1 /2. Substituting in these parame-
ters yields a pressure at the step of,
4µU L
P = .
3H 2
The pressure grows as the gap gets smaller. The positive pressure in
the gap (when H1 > H2 ) provides a lift force on the sliding object.
136 Solutions to the Navier-Stokes equations
Since the pressure grows as the layer is squeezed, the fluid can prevent
the two solids from coming into contact. The basic effect is used in
hydraulic bearings where the viscous fluid forces replaces those of solid-
solid contact.
Recall from the Poiseuille flow example that the stress acting on the
solid wall is,
" #
µ ∂u
∂y
s=n·T= .
−P
The total normal load that could be supported by the pressure in the
lubrication layer is the integral of the pressure (the y component of
the stress) under the slider. Since the pressure gradient is a constant,
the pressure varies linearly with x and the total normal force per unit
width would be,
4µU L2
FN = .
3H 2
The force is directed upward due to the high pressure in the gap and
thus there is a lift force exerted on the block.
The tangential force is that required to drag the block through the
fluid. The tangential force is found by evaluating the shear stress τ =
µ∂u/∂y at the wall and then integrating over the length. The shear
stress in region one is
∂u PH Uµ
τ1 = µ =− − ,
∂y y=0
L H
Substituting for P gives,
4µU Uµ 7µU
τ1 = − − =− .
3H H 3H
Likewise for region 2,
∂u P 2U µ µU
τ2 = µ = − =− ,
∂y y=0 2L H 3H
The total tangential force per unit width is then (τ1 L1 + τ2 L2 ), or
8µU L
FT = − .
3H
A variation of this problem is a classic one where instead of the step
change in the height, a smooth upper surface (with no step in height)
9.5 Slider bearing 137
slides at an angle relative to the lower one. This problem was solved
by Reynolds (yes, the same one) shortly after his work on pipe flow
and the transition from laminar to turbulent flow. The basic behavior
Reynolds found is comparable to the case here, it just requires a little
more analysis. When the upper surface angles upward (same as the
incoming section is thicker than the trailing one) a high pressure is
found in the gap which provides a lift force. This lift force can keep the
upper block from contacting the lower surface. As the gap gets thinner,
the pressure and thus the lift force increases so there is a stabilizing
tendency. You can observe this effect by tossing playing cards across a
table. Incline the card upward and give it a good flick and it will slide
a long distance. You probably know this if you play a lot of poker (or
other card games). If you try this experiment, you can also try to punch
holes in the card such that the pressure can’t build up in the gap. The
cards with holes will come to a quick halt when you try to slide them.
A shaft rotating through a bushing is an important component of one
of our most useful inventions; the wheel. It has been long known that
lubricating the gap between the shaft and bushing with oil or other
viscous fluid can dramatically reduce friction. The problem we worked
in this section is the beginning of the study of this classic journal bear-
ing problem. If there is load on the shaft, there will be a tendency for
the shaft to move and contact the bushing. However, when the shaft
138 Solutions to the Navier-Stokes equations
∂u ∂2u
ρ = µ 2.
∂t ∂y
The reduction for the Navier-Stokes is precisely as we had for steady
Couette flow, only now we have retained the transient term.
Dividing by the density and recalling the µ/ρ is the kinematic viscos-
ity ν, we have something that looks suspiciously like the 1D diffusion
equation,
∂u ∂2u
= ν 2.
∂t ∂y
The kinematic viscosity plays the same role in this equation as diffu-
sivity in the diffusion equation. We expect that the behavior would be
the same as applying an instantaneous change in temperature to one
side of a block of material while holding the other side cold. In this flow
case, we are talking about diffusion of momentum instead of diffusion
of thermal energy. The kinematic viscosity ν has the same units as the
thermal diffusivity.
Let’s make this equation dimensionless. Setting a scale for velocity
and length seems obvious; ũ = u/U , ỹ = y/H. However, what should
the time scale be? Using an arbitrary scale t̃ = t/t0 would yield,
∂ ũ νt0 ∂ 2 ũ
= 2 2.
∂ t̃ H ∂ ỹ
Therefore the solution at steady state will be ũ(ỹ) = 1 − ỹ. The system
will approach this equilibrium state with a time scale on the order of
H 2 /ν. The dimensionless formulation of this problem is exactly as we
found for similar problems in diffusion.
140 Solutions to the Navier-Stokes equations
θ
r
Ro Ri
Figure 9.6 Schematic for Couette flow between two concentric cylin-
ders. The outer cylinder with radius Ro rotates with angular velocity
Ω.
stated where we cancel unsteady terms and ones that disappear due to
the axisymmetric assumption. The equations with discarded terms are
for the r momentum,
v2
∂ur ∂ur vθ ∂u
r ∂P
+ − θ =−
ρ + ur +
∂t ∂r r ∂θ r ∂r
!
1 ∂ 2
∂ 1 ∂rur u
r 2 ∂u
θ
µ + 2 2 − 2 ,
∂r r ∂r r ∂θ
r ∂θ
θ momentum,
∂vθ ∂vθ vθ ∂v
θ
ur vθ 1 ∂P
ρ + ur + + = − +
∂t ∂r r ∂θ r r ∂θ
!
1 ∂ 2
∂ 1 ∂rvθ v
θ 2 ∂ur
µ + 2 2 + 2 ,
∂r r ∂r r ∂θ
r ∂θ
1 ∂rur 1 ∂vθ
+ = 0.
r ∂r r ∂θ
Integration of conservation of mass tells us the rur =Constant. Since
the radial velocity at the surface of the cylinder is zero, the radial veloc-
ity is zero everywhere. This further simplifies the momentum equations
to
v2
∂ur ∂ur vθ ∂ur ∂P
+ ur + − θ = −
ρ +
∂t ∂r r ∂θ r ∂r
!
1 ∂ 2
∂ 1 ∂ru r u
r 2 ∂u
θ
µ + 2 2 − 2 ,
∂r r ∂r
r ∂θ
r ∂θ
θ momentum,
∂vθ ∂v
θ vθ ∂v
θ
ur vθ 1 ∂P
ρ + ur + + = − +
∂t ∂r r ∂θ r r ∂θ
!
1 ∂ 2
∂ 1 ∂rvθ v
θ 2 ∂u
r
µ + 2 2 + 2 ,
∂r r ∂r r ∂θ
r ∂θ
142 Solutions to the Navier-Stokes equations
r
θ z U( r )
Since there is only one component of the velocity we drop the subscript
on the uz and just use u to denote the axial velocity for simplicity.
The final reduced momentum equation is analogous to Equation 9.1 in
cylindrical coordinates,
dP 1 ∂ ∂u
=µ r . (9.3)
dz r ∂r ∂r
Integrating twice gives
1 dP 2
u(r) = r + C1 ln(r) + C2
4µ dz
Since we don’t want a singularity at r = 0, C1 = 0. Applying the no-slip
boundary condition that u(r = R) = 0 gives,
dP 1 2
u(r) = (r − R2 ).
dz 4µ
The total flow rate, Q, is found by
Z R Z R
dP 1
Q= 2πru(r)dr = 2π r(r2 − R2 )dr,
0 dz 4µ 0
Note that we need to remember our factors of r and π when integrating
over the surface of the pipe inlet in cylindrical coordinates. Performing
the integral we obtain
π∆P R4
Q= ,
8µL
or written as the common pressure-flow relationship,
128µL
∆P = Q
πD4
Notice the term 128µL/πD4 is the hydraulic resistance which relates
pressure and flow. The pipe resistance is a strong function of the di-
ameter. One effect is the area, there is less volumetric flow if the area
9.9 Comments on the stability of solutions 145
is reduced (for the same velocity). The other effect is that viscosity is
stronger when the diameter is small.
Taking the relationship for pressure and flow in a pipe and recasting
in terms of the velocity, we could rearrange the equation to obtain,
uµL
∆P = 32 .
D2
Here u would be the average velocity in the pipe, i.e. Q/(πD2 /4). Re-
member from an earlier chapter that pressure drop is pipe flow is char-
acterized as
1 L
∆P = ρu2 f (Re).
2 D
Equating the last two expression yield,
uµL 1 L 64µ 64
32 2
= ρu2 f (Re) → f = = .
D 2 D ρuD Re
the analytical solution for laminar flow friction factor that we saw on
the Moody diagram in Chapter 3. As we have described before, this
solution is only observed up to around a Reynolds number of 2300 and
then stability is lost and flow become turbulent.
Figure 9.8 Image of the unstable vortices that wrap around the
cylinder in a Taylor-Couette device. On the right, the phase diagram
of the different types of flow observed if one can independently rotate
the inner and out cylinders in the device. Image from Album of Fluid
Motion. Phase diagram from Anderdeck, Liu, and Swinney, J. Fluid
Mech 1986.
controls the speed of both the inner and outer cylinder the flow is ex-
tremely complex and many different regimes of qualitative behavior
are seen in the phase diagram of Figure 9.8. The lower boundary of the
phase diagram from simple sheared Couette flow to the initial instabil-
ity is well predicted by linear stability analysis. The threshold and flow
patterns that one observes in experiment closely match those that are
predicted by linear analysis. The rest of the phase diagram beyond this
initial instability is much more complicated to predict. As you exceed
the threshold non-linear effects come into play and the patterns become
ever more complicated and beautiful. While the topic of linear stability
analysis is beyond what we will do in this class, it is not that difficult.
It is a topic found in a number of textbooks.
Flow in a pipe is a case where the linear stability analysis works
remarkably bad. In this case, the analysis predicts the flow is stable
at all Reynolds numbers, though we know in practice that is not true.
The mechanism for transition to turbulence in simple pipe flow is a
problem which is still worked on extensively to this day. While a lot is
understood about the laminar to turbulent transition in this problem,
148 Solutions to the Navier-Stokes equations
it is still an area of research 125 years after Reynolds first made his
observations.
There are many complications with the Navier-Stokes equations but
the point I want to emphasize is simple. It is not sufficient to simply
solve the Navier Stokes equations. We always need to ask whether those
solutions are stable. Predicting stability is sometimes straightforward
and sometimes not. Just be aware that asking the question of stability
is a big deal in fluid mechanics and the complications of stability lead a
number of interesting and beautiful phenomenon. In the end, however,
any analysis or simulations you conduct must be brought before the
ultimate judge - experiments.
etry with very little effort. A novice user can quickly generate a good
simulation.
Due to the non-linear nature of the Navier-Stokes, we are not yet at
a state of “plug and play” for all problems. Laminar single phase flows
are at that point today. If we have a 2D or 3D flow where the Reynolds
number is low enough that the non-linear terms are not too “strong”,
then a modern CFD package can usually do a good job without the user
getting too involved in the details. This does not mean that laminar
flow is always easy, that there aren’t sometimes difficulties, that you
should run to the computer and shy away from analysis or that you
are always going to get a physical answer. But practically speaking, if
you are interested in a single phase laminar flow you could probably
simulate everything you need without too much trouble or expertise.
All the examples we worked in this section could be set up and run
very quickly using a commercial CFD package. However, as we dis-
cussed earlier in the diffusion chapter, when you just do a simulation
you can miss simple scaling laws, and simple formulas that are good
for basic understanding of trends and also very useful in design prob-
lems. Simulation is another tool that can work side by side with good
analysis, but is rarely the substitute.
We will discuss turbulent flows in a later chapter, but simulating
turbulent flows is a much trickier situation and one where you should
know what you are doing. There are fundamental limitations that make
high Reynold’s number flows challenging that we will address later. For
now, it is worth noting that if you use a CFD package and get good
results, turn up the Reynold’s number and it is guaranteed to stop
working.
10
Inviscid flow, Euler’s equation and Bernoulli
means that we take the limit of the small parameter (viscosity) going
to zero and we fundamentally change the problem at hand.
In the last chapter on Navier-Stokes solutions we studied examples
where the balance of forces was between viscosity and some driving
force such as applied pressure. In all the problems in the last chapter,
the fluid had no acceleration. When using Euler’s equation we are ig-
noring viscosity and the balance of forces is between acceleration and
pressure (or gravity). A good strategy for cases of high Reynolds num-
ber flows where fluid acceleration is important is to start a problem
by assuming that Euler’s equations work. We could then check what
the predictions and trends are versus what experiments and experience
say to decide if the approximation was reasonable. We will begin to
understand the breakdown of Euler’s equations better after the next
chapter when we discuss boundary layers.
ρv · ∇v = −∇P.
make sense from the perspective of the fluid particle. If you are flowing
through the device and are accelerating upon going into the throat, then
there has to be some net force acting on you pushing you from behind.
This net force is the high pressure behind you and lower pressure in
front of you.
Note that at this point we really haven’t done anything, just rewrit-
ten the same equation in a different form using some vector calculus
identities.
Let’s take the whole equation and take the dot product with the
velocity vector itself,
1 2
−v · ρv × ω = −∇ ρv + P + ρgz .
2
Notice the term on the left side. The curl of the velocity and the vor-
ticity vectors will always be perpendicular to those two vector. Thus
the dot product with the velocity vector itself will be zero. This fact
can also be easily proven by carrying out all the terms in component
form. Since the left side is always zero, we obtain
1 2
0=v·∇ ρv + P + ρgz .
2
This fact means that the quantity in parenthesis does not change in the
direction of the velocity vector. In a fluid flow we define a streamline as
a line that follows the velocity vectors. In a steady flow a streamline
will correspond to the path a blob of injected dye would follow. Thus
in a steady flow,
1 2
ρv + P + ρgz = a constant along a streamline. (10.3)
2
This equation is known as Bernoulli’s equation. It is based on some
restricted assumptions, namely 1) incompressible, 2) steady, and 3)
inviscid flow. Despite these restrictions it is a powerful equation because
of its simplicity.
Bernoulli’s equation has a simple interpretation. The kinetic energy
per unit volume is 12 ρv2 , and the potential energy is ρgz. These expres-
sions should look familiar from particle mechanics only here we use the
mass density rather than the total mass. Pressure exerts a force per
unit area, thus a change in pressure between two locations indicates
that the pressure is doing work on the fluid. This work is related to the
familiar expression that work is equivalent to the integral of force over
distance. Bernoulli’s equation is saying that the work done by pressure
is equal to the change in energy (kinetic plus potential).
While simple and powerful, Bernoulli’s equation can also be mislead-
ing. There is a tendency to want to use it under conditions that are
156 Inviscid flow, Euler’s equation and Bernoulli
not appropriate. We must always ask if the assumptions are met when
using Bernoulli, and ultimately determine whether this simple equation
provides predictions which match reality. The most common mistake is
using Bernoulli’s equation when viscosity is not negligible.
Vθ
more general, but for simplicity lets consider that the streamlines at
some point in the flow are curved as perfect circles, such that we can
describe the flow in cylindrical coordinates. The nice thing about our
vector calculus approach is that we simply look up the different opera-
tors in different coordinate systems, and we are good to go. Looking up
the operators for cylindrical coordinates where the flow is in the r − θ
plane and gravity only points in the z direction the Euler equations in
steady flow become,
vθ2
∂ur vθ ∂ur ∂P
ρ ur + − =−
∂r r ∂θ r ∂r
∂vθ vθ ∂vθ 1 ∂P
ρ ur + =−
∂r r ∂θ r ∂θ
H(r)
z
z=0 r
Figure 10.3 Schematic for the example of computing the free surface
shape of a cylindrical tank of water in solid body rotation.
rotation and gravity are aligned such that gravity does not appear in
the radial momentum equation. Using our expression from the previous
section, the radial component of the momentum equation is
vθ2 ∂P ∂P
ρ = → ρΩ2 r = .
r ∂r ∂r
Since everything on the left side is a constant we can integrate this
equation with respect to r (holding z = Z constant) from the origin to
obtain the pressure distribution as a function of r.
1 2 2
P (r, Z) − P (r = 0, Z) = ρΩ r
2
The pressure is higher at the wall of the container than the center. The
vertical momentum equation is simple since there is no vertical flow,
namely, at a constant radial location (r = R),
P (z, R) − P (z = 0, R) = −ρgz.
P (r, z = 0) = ρgH(r) − P (z = H, r)
r
r
1
2
P
Center
pressure
Wall pressure
10.4 Vorticity
This section is going to rely on a fair amount of vector calculus and
identities. If you do not remember all these identities, then don’t get
too worried. Try to follow the derivation, but focus on understanding
the consequences. First let’s restate that vorticity is a vector defined as
the curl of the velocity,
ω = ∇ × v. (10.5)
lantic hurricane which has wind vectors drawn on it. You will notice
the vectors show the winds going in a circle in the counterclockwise
direction. To estimate the circulation, draw an arbitrary circle which
roughly follows the winds. To estimate the magnitude of the circulation
simply estimate the average wind speed on your circle, and multiply
that speed by the circumference. To compute the circulation more ac-
curately you could break your circle into little pieces to account for the
fact that the wind speeds will vary around the circle. Adding up these
little pieces is the approximation of the integral in the formal definition.
Now let us derive an equation for the vorticity. Lets take the curl of
Euler’s momentum equation in constant density, incompressible flow,
∂v
ρ∇ × + v · ∇v = −∇(P + ρgz) . (10.7)
∂t
Now we need to recall our vector calculus identities. You might re-
member that the curl of the gradient of something is zero. If you don’t
believe this, you could carry out the operation for each component to
find out. There are also identities that allow us to expand the ”v dot
grad v term”.
∂ω
+ v · ∇ω − ω · ∇v = 0 (10.8)
∂t
which is the same as
Dω
= ω · ∇v. (10.9)
Dt
What is interesting is that the pressure and gravity have disappeared.
Vorticity is a measure of solid body rotation and pressure and grav-
ity act through the fluid particles center of mass in constant density
flows, thus they can not change the rotation. If there are density gradi-
ents in the fluid, then these can couple with gravity to create vorticity.
This mechanism of vorticity production is common in oceanic and at-
mospheric flows. Now lets turn to trying to interpret these equations
starting with a simple example of 2D flow, remembering the restriction
that we have assume inviscid flow.
10.4.1 2D flows
Lets restrict our analysis to 2D flows to make things a little easier.
First of all in a 2D flow, velocity only has components in u and v and
10.4 Vorticity 163
Γ=0
Figure 10.5 Drag a spoon slowly through a bowl of water. You will
see two vortices form behind the behind of opposite sign but equal
strength. The two vortices will cancel each other such that the total
circulation around the outer dashed contour is zero.
face of the fluid a little flick. You should see two vortices generated.
They will be of opposite spin and propagate themselves for a bit before
decaying away. Now, imagine you draw your loop to calculate the cir-
culation around the spoon and far away from it. Since everything is at
rest intially, the circulation is zero. Now you flick the spoon and create
your vortices. So even though viscosity has done something inside the
loop (we could not create the vortices without it), the circulation the-
orem still applies far away from the loop. So the circulation is still zero
even though there is vorticity inside the loop. The circulation theorem
states that the circulation is zero and remains so. This means that the
positive and negative vorticity you generated must cancel each other
out via the definition of circulation as the vorticity integrated over the
area (via Stokes’s theorem). In this experiment you get two vortices of
equal strength and equal size, such that the total integrated vorticity
is zero. The theorem tells you nothing of the vortices decay, just that
the total circulation should always be zero.
You may have noticed that the equal sized but oppositely signed
vortices propagate themselves in a straight line. This behavior can be
understood by using an ideal vortex which is a solution to Euler’s equa-
tion. The ideal vortex has a velocity field given as
Γ
vθ (r) =
2πr
This velocity field is provided in cylindrical coordinates, thus the flow
is in the θ direction but only depends on r. This velocity field can be
plugged into Euler’s equation and would be found to be an acceptable
solution. It is easy to show that the circulation of this vortex is
Z 2π Z 2π Z 2π
Γ Γ
vθ (r)rdθ = rdθ = dθ = Γ.
0 0 2πr 0 2π
The circulation is a constant regardless of the radius of the circle we
choose to perform our calculation along. The vorticity, which only has
a component in the z direction is,
1 ∂rvθ 1 ∂ Γ
ωz = = =0
r ∂r r ∂r 2π
zero everywhere, except at the origin where it is undefined.
Now consider two vortices of equal circulation but opposite sign, sep-
arated by some distance, Figure 10.6. One vortex will induce a velocity
10.4 Vorticity 165
Image vortices
Solid
wall
field that will push the other forward. Since the vortices just move with
the fluid, that is what Dω/Dt = 0 says, then you can think about the
velocity of one vortex simply acting to move the other. Since the vor-
tices are the same strength they push each other in a straight line. If
you created two vortices of equal strength but the same direction of the
spin, the two vortices would orbit each other in a circle.
Now repeat your experiment with the spoon near the wall of your
bowl. You should be able to see that the vortices propagate themselves
to the wall, and then spread apart as they approach the wall. This
can be understood by imagining a set of image vortices on the other
side of the wall. These image vortices don’t really exist, but you should
be able to convince yourself with symmetry arguments that the flow
generated by the image vortices will always cancel the real vortices such
that there is no velocity penetrating the wall. The flow on the left side
of Figure 10.6 where the real vortices exist is the same whether there
was the wall or the set of symmetric image vortices. Figuring out the
qualitative flow is easier by visualizing the image vortices. Now as the
four vortices approach the wall, you can see that the image vortices
push each other outward if you consider the action of each vortex on
the other.
166 Inviscid flow, Euler’s equation and Bernoulli
10.4.2 3D flows
In 3D, the above analysis just gets a little more complicated. First of all,
the circulation theorem remains, so without proof we will simply state
that in 3D the circulation is constant around a material loop. However,
simple applications of the circulation theorem (at least in this course)
are usually simple 2D approximations to give us some qualitative un-
derstanding of the flow.
The vorticity equation in 3D is,
Dω
= ω · ∇v.
Dt
and has a non-zero term on the right hand side. Vorticity for a material
point is not constant in 3D as there is an extra term in the equation.
This extra term accounts for things which cannot occur in a 2D flow.
One effect is vortex stretching. If a vortex is stretched out by a flow,
it intensifies, just like an ice skater speeding up as they pull their arms
in. If a vortex is squashed, the vorticity decreases. The stretching of
the vortex can be see in any bathtub drain. The vortex gets intensified
as it stretches down the drain. Another easy experiment to see vortex
stretching is to take a soda bottle and fill it with water. Turn it upside
down over the sink and as it is glugging, give it a strong swirl by hand.
You will set up a vortex that will be intensified as it is stretched and
pulled down out of the bottle. There are also effects that occur due
to vortex tilting. Tilting a vortex line can change a particles rotation.
This effect would be analogous to the physics demo where the instructor
takes a spinning bicycle wheel and tilts it while sitting on a stool which
is free to rotate.
In vortex dynamics we refer to vortex lines and tubes. A vortex line is
one which follows the vorticity vector, like a stream line. It would be the
axis of a tornado. A vortex tube would be a collection of vortex lines and
the vorticity vector is everywhere parallel to the surface of the vortex
tube. In 3D there are three laws of vortex motion that were derived by
Helmholtz in 1858. These laws are good for the approximations we have
been dealing with - inviscid and constant density. Helmholtz’s laws are
(Saffman (1992)).
Due to the definition of vorticity as the curl of the velocity, it true that
∇ · ω = 0 since the divergence of the curl of any vector is zero (you
should quickly see if you can prove this fact to yourself). Since vorticity
is divergence free, then a vortex tube must have constant strength along
its length. While this last statement might not sound obvious, the proof
is precisely the same as showing that the total volumetric flow rate
through a pipe of varying cross section area must be the same at every
cross section if the flow is incompressible. If a flow is incompressible,
∇ · v = 0. The walls of the pipe are like the walls of the vortex tube
- since by definition there is no vorticity normal to the surface of the
vortex
R tube. At every cross section of Rthe tube the total flux of vorticity,
ω·n, is a constant just like in a pipe v·n is a constant. If the strength
of the vortex tube is constant everywhere along the length, then vortex
tubes must either close on themsleves (i.e. a smoke ring), go to infinity,
or end on solid boundaries. An example of vortex tubes are the long
white contrails seen behind airplanes on clear days when the weather
conditions are right. The two long white streaks are intense vortex
tubes which extend far across the sky. If you observe these streaks
closely you will notice that some distance away from the plane, the
individual tubes start developing a wavy character. These waves are
due to a fluid instability known as the Crow instability. Notice that
as the amplitude of the waviness increases the two tubes will interact,
cross, reconnect, and form a series of vortex loops. These circles are now
closed vortex tubes. How well you can make these observations depends
on the weather as the level of atmospheric turbulence can disrupt the
vortex visualization.
v = ∇φ,
potential flow problems for you. Potential flow finds a few niche appli-
cations, can many times (especially in aerospace applications) provide
a good qualitative picture of the flow and the theory is extremely im-
portant in the history of the development of the field of fluid dynamics.
However, the practical utility of potential flow is limited.
a) b)
Force/width
CL = 1 2
,
2 ρU `
-Γ Γ
Figure 10.8 Schematic of the starting vortex for an airfoil. The cir-
culation calculated around the outer oval must be zero, so the circu-
lation around the two inner loops must cancel each other out. The
airfoil must have circulation of the opposite sign as the starting vor-
tex. If the airfoil is suddenly stopped a vortex of the opposite sign as
the starting vortex is shed in order to maintain no net circulation.
a) b)
The analysis of fluid flow via Euler’s equations seems relatively nice.
While we presented only a few highlights in this book you should have
gotten the impression that a number of problems could be readily solved
mathematically or computationally using Euler’s equations for incom-
pressible flow. So while we know viscosity is always with us, it is a small
force compared to the others in the problem in a number of applica-
tions. At the scale of humans, everything is high Reynolds number.
However, it has been observed for a long time that Euler’s equations
do an unusually bad job at making useful predictions. Euler’s equation
predicts no drag force, for example. It was not until the early 1900s
when Prandtl introduced the idea of the boundary layer, that people
began to appreciate why Euler’s equations were so poor and how to
rectify our understanding (Anderson (2005)).
If we have a high Reynolds number flow around a streamlined ob-
ject such as an airfoil, what is observed is that the fluid velocity just
away from the surface of the airfoil very closely matches the solution to
Euler’s equations. This part of the flow is not influenced by viscosity.
As we approach the surface, the flow velocity goes to zero because the
no-slip condition is obeyed. However, the size of this boundary layer is
extraordinarily thin. Flow over an airplane wing could have boundary
layers measured in the units of millimeters. Why should this thin region
play such a major role in determining flow? The answer is the topic of
this chapter.
We need to keep in mind all the forces at play in our problem. When
we looked at solutions to the Navier-Stokes equations we found we
could readily solve problems where there was no fluid acceleration - the
174 Boundary layers
balance was between viscosity and some driving force such as applied
pressure. When we discussed Bernoulli’s and Euler’s equation, we con-
sider the balance between pressure and acceleration. In this chapter we
will find that in many cases there is a complex interplay of fluid ac-
celeration, viscosity and pressure. While our analytical tools get more
complex as the flows do, all is not hopeless and there is a lot we can
do with simple problems.
∂u ∂2u
= ν 2.
∂t ∂y
Recall that the kinematic viscosity plays the same role in this equation
as thermal diffusivity in the heat equation. We expect that the behavior
would be the same as applying an instantaneous change in temperature
to a semi-infinite solid.
We now want to make this equation dimensionless as we did with
Couette flow. We can set the velocity scale to ũ = u/U . However, what
should the time scale or spatial scale be? There is no natural geometric
length to scale length by. Our only choice is δ, the thickness of the fluid
region which is in motion. However this is an unknown scale. Using an
11.1 Impulsively started plate 175
U∞ U∞
Boundary
δ
Layer
The boundary layer thickness can only depend upon the plate Reynolds
number.
We could estimate the dependence by a simple √ argument using the
solution to the impulsively started plate, δ ∼ νt. In the boundary
layer problem, there is no time. At a given downstream location, the
“time” is the time it takes fluid to reach that point, i.e. t ∼ L/U∞ .
√
r
νL
δ ∼ νt ∼ .
U∞
In terms of the Reynolds number this equation can be rewritten as
r r
δ ν 1
∼ ∼ .
L U∞ L Re
Since the flow progresses downstream, we the argument above works
for any x location. Thus we might expect,
r r
δ ν 1
∼ ∼ ,
x U∞ x Rex
where we use the Reynolds number based on the x location. This scaling
suggests that the boundary layer thickness grows as the square root
of the distance down the plate. To get a sense of scale, lets take air
ν = 10−6 flowing at U∞ = 1 m/s over a 1 m length plate. The boundary
layer would be 1 mm. If the velocity were increased to 100 m/s (an
airplane taking off), the boundary layer thickness would be 100 microns.
The shear stress on the plate due to viscosity can also be estimated.
For a Newtonian fluid, you can look back at the previous chapter and
178 Boundary layers
see that the shear stress (tangential to the surface) is τ = µ∂u/∂y. The
stress is given by the velocity gradient at the surface. The shear stress
would be approximately,
∂u U∞ U∞ µU∞ p
τ (x) = µ ∼µ ∼µ ∼ Rex
∂y δ xδ/x x
Making an analogy with drag, we would follow dimensional analysis to
define the coefficient of friction as,
τ (x)
Cf = 2 /2
.
ρU∞
Using the estimate we have already obtain for τ we have an estimate
for the scaling of the coefficient of friction,
r
µU∞ p 2µ p 1
Cf ∼ 2
Rex ∼ Rex ∼ .
xρU∞ /2 xρU∞ Rex
This simple estimate predicts that the shear stress should be highest
near the leading edge of the plate and continually decrease with down-
stream distance. The Rtotal drag force, F , for a plate of length L, is then
calculated from F = τ (x)dx. The total drag coefficient is thus given
as,
F 1
CD = 2 /2
∼√ .
LρU∞ ReL
From such simple arguments we have no way to compute what the
pre-factors in these estimates might be. However, we find that we can
get the basic scaling and qualitative behavior for the laminar boundary
layer.
U∞ U∞
δ
calculate everything you want to know about the boundary layer. The
thickness, defined as the distance when the velocity field is 99 % of the
free stream is,
δ 5
=√ .
x Rex
The factor of 5 comes from the solution, but the form was acquired
through dimensional arguments. The shear stress vector at a point in
a 2D flow is given by n · T. We can evaluate the stress at the point to
be in dimensionless terms,
τ (x) 0.664
Cf = 2 /2
=√ ,
ρU∞ Rex
where Cf is known as the coefficient of friction. The total drag force
(per unit width) is thus given as,
F 1.328
CD = 2 /2
=√ .
LρU∞ ReL
The numerical values come from the calculated numerical solution. The
really important point is that the basic scaling of the answer comes from
consideration of dimensional analysis and some physical arguments.
182 Boundary layers
tion to the boundary layer thickness. Empirical fits to the data given
boundary layer thickness as,
δ 0.37
= ,
x Re1/5
x
a) b)
a) b)
Note that the total force has a contribution from pressure and viscous
stresses - distinguishing between these two forces will be important to
our understanding soon. From dimensional analysis, we know that the
drag force is expressed as the drag coefficient which is only a function
of the Reynolds number
F
1 2 2
= Cd (Re).
2 ρU0 πR
Knowing that there is some functional relationship for the drag coef-
11.5 Observations of drag on a sphere 187
the surface due to the velocity gradient. The shear stress at the surface
depends upon the viscosity of the fluid. On a sphere, the stress vector
at a point is given by the velocity gradient at the surface and the local
pressure. In Euler’s equation the pressure on the surface of the sphere
is symmetric, thus the net force due to pressure is zero. If we watch the
experimental streamlines at high Reynolds number where the boundary
layer around the sphere is still laminar but strong separation sets in, we
would see the streamlines move around the sphere, but do not recover
on the other side. If we measured the pressure on the surface we find
strong asymmetry when there is flow separation. Thus pressure at the
forward stagnation point is greater than the ambient by ρU 2 /2. On the
back side of the sphere, the wake is relatively quiet. If we look at the
streamlines at the top or bottom of the sphere, they would be straight
and thus the pressure would be the ambient. Thus the pressure in the
wake should be something close to the ambient pressure. Therefore, very
crudely, the force acting on the front half of the sphere due to pressure
is 21 ρU 2 πr2 . The force in the back half is zero. Thus the net force due
to pressure drag is on the order of 12 ρU 2 πr2 . Thus the drag coefficient
is about 1. Pressure drag dominates over friction at the surface in this
regime.
viscous drag, the net effect is that the turbulent boundary layer has
lower overall drag.
The role of boundary turbulence in reducing drag is the reason for
dimples on golf balls. The irregular surface induces turbulence to occur
at lower Reynolds number than a smooth sphere. The turbulent bound-
ary layer delays separation and the golf ball goes further. Imperfections
on the surface of a sphere and its effect on tripping the boundary layer
into a turbulent state can help explain why curveballs curve in baseball.
12
Turbulence
lenge the fastest supercomputers and we can quickly exceed the limits
of the worlds fastest computers. For many problems, the solutions span
too many orders of magnitude in time and space to be resolved. Fur-
ther, the outlook is such that for many important problems waiting for
computers to get faster doesn’t seem like a good prospect for the near
future, even if Moore’s law continues to hold.
The problem is that scaling arguments show that the total amount
of computational effort to solve a 3D turbulent flow scales very poorly
with Reynolds number. The total amount of computational effort scales
as
total computing power ∼ Re3 .
where L is the length scale of the largest eddies and the start of the
cascade. For an atmospheric flow we can easily reach Reynolds numbers
of 1012 or higher; there are indeed many orders of magnitude of length
over which the turbulence cascade exists.
We can describe the flow via a spatial Fourier series u(x) = Ak eikx ,
P
where
Z
Total kinetic energy = E(k)dk.
cide that we only care about the big swirls so we decide to make our
grid spacing 1 mm. Certainly this will suffice? Now we have 300 grid
points in each direction; still a lot but not outside of what a modern
computer and software could handle in 3D. However, 1 mm is 20 times
greater than the length scale where viscosity can remove energy from
the system. What happens is that energy will be put in at the large
scale, that energy will cascade to smaller scales, and when the energy
in the simulation reaches 1 mm (the grid spacing) it will just sit there.
There is no mechanism to remove energy from the simulation. Energy
will build up and the simulation will die a sad numerical death.
There are fixes to this problem by adding models which act as numer-
ical sponges to soak up all the grid scale energy from the simulation.
Methods called Large Eddy Simulations (LES) have this feature of be-
ing able to resolve our stirred bucket by modeling, but not resolving,
the smallest scale motions. The LES models are often called turbulence
closure models. There are many different models that have advantages
and disadvantages in different situations. There is no single model that
always works the best and it requires training and experience to do LES
of turbulent flows well. You can easily see that this turbulence closure
problem will be with us for a while. For an atmospheric flow which
effect the climate, the Reynolds number can be 1015 the large scales
structures are the sizes of continents. We are going to be continue to
need closure models in weather and climate modeling in my lifetime.
Most modern CFD packages have a number of different turbulence
closure models built in. The variety and character of these models is
too extensive to cover here. In the software, it can be just as easy as
clicking from one model to the next. However the different models have
different limitations and cases where they work better or worse. Just
be warned that you should be skeptical of the results unless you have
had more training and experience than you are getting from this brief
chapter.
Many people think that CFD is the answer to everything fluids. The
impulse for many people when confronted with a fluids problem is to
run to CFD. CFD can be powerful, but it is a tool that can be easily
misused. If you are ever working somewhere and your boss wants to do
CFD of a turbulent flow, be cautious going down this road unless you
have learned more about fluid dynamics than this course. It is very easy
to get a commercial code to ”solve” but give you an incorrect answer.
12.3 Numerical simulation of turbulence (CFD) 197
problems (including the ones I’ll demonstrate) that tend to gloss over
the assumptions that are made as though you just believe they are
true. Like dimensional analysis, the procedure of control volume analy-
sis is straightforward, it is the proper use of it that takes some physical
insight that only comes with experience. The control volume is an excel-
lent tool, but keep in mind that there are also problems where doesn’t
really help us much.
Control volume analysis is an excellent tool for getting order of mag-
nitude estimates or simple scaling laws. Often in engineering situations
we care about the approximate magnitude of the total force or the pres-
sure, for example, and a simple control volume analysis can be more
useful that a complex simulation. Often we can get a useful and rea-
sonable answer with just a little effort. Control volume analysis is also
a quick way to check other more detailed calculations or simulations.
If you do a full analysis of the Navier-Stokes equations and solve for
the full velocity field, but that field does not satisfy the overall control
volume balance then you have done something wrong.
Many textbooks on fluid dynamics put the control volume analy-
sis very early in the development. A good reason for doing so is that
the analysis is much easier than analysis with the Navier-Stokes equa-
tions. However, when confronted with a real world problem that we
don’t find in a textbook executing a useful control volume analysis re-
quires some physical intuition. Similar to dimensional analysis, control
volume analysis is usually most powerful when coupled with good phys-
ical understanding and intuition about the important effects. Similar to
dimensional analysis, it can be straightforward to learn the technique
of control volume analysis, but using it wisely comes with experience.
Z
Dv
ρ = ρg + ∇ · T dV
Dt
13.1 Control volume formulation 201
We can now move the integral on the left hand side to a different form
using the Reynolds Transport Theorem,
Z Z Z
Dv ∂(ρv)
ρ dV = dV + (ρv)v · ndS.
Dt ∂t
Using the divergence theorem we can also convert the stress tensor
term on the right side of the original equation back to integration over
a surface,
Z Z Z Z
∂(ρv)
dV + (ρv)v · ndS = ρgdV + n · TdS.
∂t
Substituting our constitutive law for a Newtonian fluid in incompress-
ible flow (T = −P I + 2µS), we have,
Z Z Z Z Z
∂(ρv)
dV + (ρv)v · ndS = ρgdV − P ndS + 2µ n · SdS.
∂t
Since we are now interested in a control volume fixed in space, the
derivative on the left hand side can be moved outside the integral, so
Z Z Z Z Z
d
ρvdV + (ρv)v·ndS = ρgdV − P ndS+2µ n·SdS. (13.1)
dt
The above expression holds for the fluid only. A more useful form in-
cludes an external force acting on an object in the flow. In this case,
you can draw a control volume around whatever you like, as long as
you remember to include the force which is holding the object in place.
Z Z Z Z Z
d
ρvdV + (ρv)v·ndS = ρgdV − P ndS +2µ n·SdS +Fext .
dt
(13.2)
Following the same ideas, the integral form of conservation of mass for
a fixed control volume is
Z Z
d
ρdV + ρv · ndS = 0. (13.3)
dt
The integral form of the momentum and mass conservation equations
for a fixed region can be very useful in a number of cases. The primary
use is in making estimations/calculations of net forces acting on an
object in a flow.
You should first note that conservation of momentum is a vector
equation - 3 components for x, y and z. On the left side of Equation
13.1 we have the rate of change of momentum inside the control volume
202 Control volume analysis
(first term) and the net flux of momentum coming into/out of the
control volume (2nd term). In the second term on the left side, note
that the momentum ρv is a vector quantity, while v · n is the rate
that the momentum crosses the surface. On the right side, the first
term is the net body force which acts on the whole volume. The second
term is the force due to pressure and the third term is the force due to
viscosity - both of these exert forces only on the surface of our volume.
What is nice about this expression is there are many cases where we
are only interested in the net force on an object, and in those cases we
can often use the integral form of the equation without worrying about
the details of the flow.
R
Getting the sign right of the term (ρv)v · ndS can be a little tricky.
You have be be careful that the velocity vector has a sign, as well as
the term v · n. Since we take normal vectors to point out of the volume,
then v · n is positive for outflow and negative for inflow.
13.2 Examples
The only way we can make sense of the complex looking equation is to
work through a few examples. The basic approach I usually take with
these problems is as follows. I start by drawing a sketch and picking
a control volume that seems convenient. A convenient control volume
often (but not always) has surfaces where the velocity inlets and outlets
are normal to the surface. I like to write out the entire control volume
equation and then systematically cross out terms that I have good
reason to ignore (say unsteady effects in a steady flow). I then would see
where that takes me. If I can get an interesting result without making
many assumptions, then great. If I don’t get much then I might start
making assumptions that I am less confident are true (say ignoring
viscosity in a case where it is not clear I can). I like to try and keep
track of my assumptions so when I get to the end of a problem, it is
easy for me to review what I had to do to get there. One of the most
common assumption you will tend to make in these problems is that
the velocity field is uniform over some region, an assumption we often
make because we have little choice.
13.2 Examples 203
A2 n
Fx
U A1 θ
n
Fy
Figure 13.1 Control volume for finding the force a water jet exerts
on an inclined block. The control volume is shown as the dashed
line.
evaluate the surface integral without knowing the velocity field of the
inlet and outlet jet. Let’s assume for now that the jet enters and exits
with a uniform velocity field, U1 and U2 respectively. On the inlet the
normal vector and the velocity are not aligned so the dot product is
negative and on the exit, the velocity and the normal vector are aligned
so the dot product is positive. Since the velocity at the inlet and exit
is assumed constant, then the surface integral is easy to do,
Z
ρv · ndS = −ρU1 A1 + ρU2 A2 = 0.
This statement simply says that the total mass flow rate in equals
the total mass flow out. Note that when we write the surface integral
above, it means that we have to evaluate ρv · n at every point along the
control volume surface and integrate it up. In this example it is only
the two little patches where fluid comes in and goes out that has any
fluid velocity.
Now let’s consider the momentum equation. Let’s start by writing
out the full integral equation and cross out the unsteady terms and
gravity which we previously stated we would neglect,
Z Z Z Z Z
Hd H Z
− P ndS +2µ n·SdS +Fext .
H
ρvdV + (ρv)v·ndS = ZρgdV
dt HH
H
Z
Z
Notice that the surface integral terms for pressure and viscosity are
evaluated only on the surface of the control volume. So while viscous
forces and pressure may do something inside the control volume on the
surface we have drawn it would be reasonable to assume that they have
no contribution. Since the control volume is open to atmosphere every-
where, the pressure around the surface of the control volume would be
atmospheric. If we integrate a constant pressure around a closed loop,
then the integral of P n around the loop results in zero net force. While
viscosity certainly acts along the surface of the block, along the sur-
face we have drawn the control volume boundary viscosity would not
seem to play a role since we are assuming uniform flow outward. Recall
that viscous stresses only occur when we have velocity gradients, so
a uniform free jet with uniform flow has no viscous stresses. You are
starting to see that in order for us to make progress we have to start
to use some physical insight. By making these physical arguments to
neglect the pressure and viscous terms on the control volume boundary,
13.2 Examples 205
Note that the dot product of v · n = −U1 since the velocity vector and
normal vector are not aligned. Since the velocity at this surface is all
in the x-direction, then v = U1 . On the outlet at surface 2,
Z Z
(ρv)v · ndS = (ρU2 cos(θ))(U2 )dS = ρU22 A2 cos(θ).
2 1
Note that the dot product of v · n = U2 since the velocity vector and
normal vector are aligned. Since the velocity at this surface is inclined
at angle θ the x-component of the velocity is U2 cos(θ). Putting this
together gives us the external force as,
ρ U22 A2 cos(θ) − U12 A1 = Fx .
Note that the dot product of v · n = U2 since the velocity vector and
normal vector are aligned. The dot product term is the same in the x
and y momentum equations. Since the velocity at this surface is inclined
the y-component is v = U2 sin(θ). The total y component of the force
is,
ρU22 A2 sin(θ) = Fy .
When we make our assumption that the area of the jet does not change,
ρU12 A1 sin(θ) = Fy .
Thus the final result for the case where the area of the jet does not
change is
cos(θ) − 1
Fext = ρU12 A1
sin(θ)
13.2 Examples 207
While this result assumes the constant velocity profile and the fact that
the jet area remains the same from inlet to exit, it seems reasonable
to think that scaling, order of magnitude and trends might be quite
reasonable.
v(x)
U0
Fx H u(y)
U0 v(x) U0
L
Figure 13.2 Control volume for finding the drag on an object from
an experimentally measured wake.
a very viscous fluid coming from our hose, we would expect there to
be a more dramatic change in the increased area of the jet. Using the
control volume approach we can only compute the force on the plate if
we had a measurement of the exit area.
If the fluid had very low viscosity (high Reynolds number), we could
apply Benoulli’s equation. If we applied Bernoulli’s equation from the
inlet to the exit, we would see that just like in the last example the
velocity should not change. Since Bernoulli’s equation is only valid for
inviscid flow, our conclusion is consistent with the fact that at high
Reynolds number, we expect the inlet and outlet areas of the jet to be
approximately equal.
The order of the terms are left, right, and top/bottom (hence the factor
of 2). Note that on the inlet (left boundary) the velocity and normal
vector point in opposite directions hence the negative sign on that term.
Conservation of mass simply states that since there is a velocity deficit
in the wake of the object, there must be some flow out the top and
bottom surface.
Conservation of momentum is,
Z Z Z Z Z
Hd H Z
−HH XX
X · SdS +Fext .
ρvdV
H + (ρv)v·ndS = Z HH nX
ZρgdV P ndS
+ 2µ
dt HH
H Z
XX
We have crossed out terms since 1) the flow is steady, 2) we will ig-
nore the effect of gravity (or assume perpendicular to the page), 3) the
object is in an external flow field and thus the pressure is atmospheric
everywhere, 4) that along the boundary of control volume which is far
from the object there are no significant viscous stresses. Under our as-
sumptions, the net force on the object is given by the imbalance in the
momentum coming in and out of the control volume.
Z
(ρv)v · ndS = Fext .
Now, lets evaluate our terms for each surface for the x component of
the momentum,
Z H Z H Z L
2 2
− ρU0 dy + ρu(y) dy + 2 ρU0 v(x)dx = Fx .
0 0 0
The first term is the inlet. Since the flow and the normal vector are in
opposite directions, v · n = −U0 and the x component of the velocity
is U0 . The exit integral is computed similarly, only the sign is positive
since the velocity and flow direction are aligned. The top and bottom
terms should be treated with care. For the top and bottom, the outflow
is v · n = v(x). The x component of velocity is U0 since the surface is
far from the object.
210 Control volume analysis
A2
U1 A1
U2
Figure 13.3 Control volume for finding the net force required to
hold the nozzle on a pipe.
13.2.4 Nozzle
Consider the tapered nozzle as shown in Figure 13.3. Lets consider a
steady flow rate and we want to estimate the force required to hold the
nozzle in place at the flange. We draw the control volume as shown.
We will assume steady flow, ignore gravity, assume the inlet and outlet
flow fields are uniform, and assume a high Reynolds number flow.
At this point, we can probably do the conservation of mass pretty
easily,
Z Z
HdH
ρdV + ρv · ndS = −ρU1 A1 + ρU2 A2 = 0.
H
dt H
H
13.2 Examples 211
Here we have neglected the viscous terms, not because we believe that
viscosity doesn’t matter inside the nozzle but that viscosity probably
doesn’t add much to on the surfaces of the control volume that we have
drawn. Since the flow is contained inside the nozzle, we must be careful
to use different pressure. Now, evaluating the x component of this force
and being careful with the signs,
In this example the momentum in and out of the control volume is sim-
ilar to previous examples. What is new is how we handle the pressure.
There are a few signs to keep track of. There is a negative in front of
the pressure term. We also must account for the direction of the normal
vector. Also, all of pressure at location 1 is acting on the area of the
inlet, while the pressure at point 2 is acting only smaller exit area. The
atmosphere is pushing on the right side of the control volume on the
area A1 − A2 . Since the pressure just outside the nozzle is in the open
air, it is reasonable to set the pressure at the exit to be P2 ≈ P∞ . In
this case our equation becomes,
Now we see in this form that the force to hold the nozzle is the difference
in momentum coming in and out of the volume and the pressure at the
nozzles inlet. We can use conservation of mass (U1 A1 = U2 A2 )to write
as
A1
ρU12 A1 − 1 − (P1 − P∞ )A1 = Fx .
A2
Note the sign of the terms. The net change in momentum alone would
lead to a force in the positive direction - meaning the nozzle is com-
pressed on the tube. Since the flow shoots fast out of the nozzle there
is a thrust like a rocket. The pressure term leads a negative force and
wants to pop the nozzle off. We can’t solve for the total force without
more information.
One assumption we could make is assume that the flow is high
212 Control volume analysis
Reynolds number and apply Bernoulli’s equation from the inlet to the
outlet. Bernoulli would yield,
1 1
P1 + ρU12 = P∞ + ρU22
2 2
or
1 2 1 2
P1 − P∞ = ρU − ρU1 .
2 2 2
Using conservation of mass gives,
A21
1
P1 − P∞ = ρU12 −1
2 A22
Using the result from Bernoulli and going back to our momentum
balance would give the nozzle force as
2
1 2 A1
Fx = − ρU1 A1 −1
2 A2
When A1 > A2 then the force acting on the nozzle is negative - meaning
without a bolt holding the nozzle on, there is a tendency of the nozzle
to pop off. The final result is only valid for high Reynolds number where
Bernoulli’s equation is reasonable.
water intakes into the sprinkler, the question is which way does the
sprinkler spin? It seems the control volume analysis may be useful for
this problem because we just want to know which way it spins, not how
fast.
To analyze the problem, let’s look at one half of the sprinkler, imagine
that we hold the sprinkler fixed and calculate the force, F , needed to
hold the normal sprinkler in place. Let’s assume the water comes out
as a jet of uniform velocity u and area A, the flow is steady, the density
is constant, and viscosity is not important. Under these assumptions,
we have,
Z Z
(ρv)v · ndS = P ndS + F
The only tricky part of control volume analysis is getting the sign cor-
rect and since we are trying to get the direction correct, we need to treat
our signs with care. There are two signs to keep straight on the left side
of the equation, v · n and v. At the exit as drawn,R both v · n and v are
positive and thus we can estimate the term as (ρv)v · ndS ≈ ρAu2 .
Normally, we would assume that the pressure at the exit is the same
as the pressure in the environment. The force is positive and estimated
to have a magnitude of F = ρAu2 and be positive as drawn in the
schematic of figure 13.4. If we remove the force, the sprinkler spins.
Now consider the inverse sprinkler. The same integral expression
must hold true. Momentum enters the control volume at the same rate
as the sprinkler in forward motion. Since v · n and v are both nega-
tive, their product is positive; the momentum across the surface is the
same in magnitude and sign as the sprinkler in normal operation. If this
214 Control volume analysis
were all, the sprinkler should spin the same direction whether the fluid
comes in or out of the sprinkler. If one does the experiment (you can
try yourself with a bendable drinking straw) you find that the sprinkler
does not spin. If you did the experiment carefully you would actually
find the sprinkler spins very slowly in the reverse direction.
The reason the control volume approach does not work is that we
have not gotten the exit pressure right. In the sprinkler operating in
“normal” mode the pressure at the exit must be at least a little higher
at the exit than the environment since pressure gradients would be re-
sponsible for “pushing” the flow. When the sprinkler is operating in
reverse mode, the pressure at the exit must be lower than the environ-
ment for flow to go in that direction. It turns out that if viscosity is
not important then the low pressure at the sprinkler exit (which wants
to make the sprinkler spin counterclockwise) exactly cancels the mo-
mentum coming in (which wants to make the sprinkler spin clockwise).
I cannot derive the experimental result using control volumes. Control
volume analysis gives you no way to calculate the pressure in this prob-
lem. While a number of people have written about this problem I have
yet to find a solution that I find truly satisfying, nor have I been able
to work one out myself.
Bibliography
Anderson, J.D. 2005. Lundwig Prandtl’s boundary layer. Physics Today, 58,
42–48.
Bird, R.B., Stewart, W.E., and Lightfoot, E.N. 1960. Transport Phenomena.
Wiley.
Bridgman, P.W. 1922. Dimensional Analysis. Yale University Press.
Kundu, P.K., and Cohen, I.M. 2004. Fluid Mechanics. Elsevier Academic
Press.
Panton, R.L. 1996. Incompressible Flow. Wiley.
Rayleigh. 1915. The principle of similitude. Nature, 95, 66–68.
Reynolds, O. 1883. An experimental investigation of the circumstances which
determine whether the motion of water shall be direct or sinuous and
the law of resistance in parallel channels. Phil. Trans. R. Soc., 174,
935–982.
Saffman, P.G. 1992. Vortex Dynamics. Cambridge University Press.
Schey, H.M. 1997. div grad curl and all that. W.W. Norton and Co.
VanDyke, M.V. 1982. An Album of Fluid Motion. Parabolic Press.