You are on page 1of 52

Lecture notes

Introductory uid mechanics


Simon J.A. Malham
Simon J.A. Malham (18th September 2012)
Maxwell Institute for Mathematical Sciences
and School of Mathematical and Computer Sciences
Heriot-Watt University, Edinburgh EH14 4AS, UK
Tel.: +44-131-4513200
Fax: +44-131-4513249
E-mail: S.J.Malham@ma.hw.ac.uk
2 Simon J.A. Malham
1 Introduction
The derivation of the equations of motion for an ideal uid by Euler in 1755, and
then for a viscous uid by Navier (1822) and Stokes (1845) were a tour-de-force of
18th and 19th century mathematics. These equations have been used to describe and
explain so many physical phenomena around us in nature, that currently billions of
dollars of research grants in mathematics, science and engineering now revolve around
them. They can be used to model the coupled atmospheric and ocean ow used by
the meteorological oce for weather prediction down to any application in chemical
engineering you can think of, say to development of the thrusters on NASAs Apollo
programme rockets. The incompressible NavierStokes equations are given by
u
t
+u u =
2
u p +f,
u = 0,
where u = u(x, t) is a three dimensional incompressible uid velocity (indicated by
the last equation), p = p(x, t) is the pressure and f is an external force eld. The
frictional force due to stickiness of a uid is represented by the term
2
u. An ideal
uid corresponds to the case = 0, when the equations above are known as the Euler
equations for a homogeneous incompressible ideal uid. We will derive the Navier
Stokes equations and in the process learn about the subtleties of uid mechanics and
along the way see lots of interesting applications.
2 Fluid ow
2.1 Flow
A material exhibits ow if shear forces, however small, lead to a deformation which is
unboundedwe could use this as denition of a uid. A solid has a xed shape, or at
least a strong limitation on its deformation when force is applied to it. With the cate-
gory of uids, we include liquids and gases. The main distinguishing feature between
these two uids is the notion of compressibility. Gases are usually compressibleas we
know from everyday aerosols and air canisters. Liquids are generally incompressiblea
feature essential to all modern car braking mechanisms.
Fluids can be further subcatergorized. There are ideal or inviscid uids. In such
uids, the only internal force present is pressure which acts so that uid ows from
a region of high pressure to one of low pressure. The equations for an ideal uid
have been applied to wing and aircraft design (as a limit of high Reynolds number
ow). However uids can exhibit internal frictional forces which model a stickiness
property of the uid which involves energy losssuch uids are known as viscous
uids. Some uids/material known as non-Newtonian or complex uids exhibit even
stranger behaviour, their reaction to deformation may depend on: (i) past history
(earlier deformations), for example some paints; (ii) temperature, for example some
polymers or glass; (iii) the size of the deformation, for example some plastics or silly
putty.
Introductory uid mechanics 3
2.2 Continuum hypothesis
For any real uid there are three natural length scales:
1. L
molecular
, the molecular scale characterized by the mean free path distance of
molecules between collisions;
2. L
uid
, the medium scale of a uid parcel, the uid droplet in the pipe or ocean
ow;
3. L
macro
, the macro-scale which is the scale of the uid geometry, the scale of the
container the uid is in, whether a beaker or an ocean.
And, of course we have the asymptotic inequalities:
L
molecular
L
uid
L
macro
.
We will assume that the properties of an elementary volume/parcel of uid, however
small, are the same as for the uid as a wholei.e. we suppose that the properties of
the uid at scale L
uid
propagate all the way down and through the molecular scale
L
molecular
. This is the continuum assumption. For everyday uid mechanics engineer-
ing, this assumption is extremely accurate (Chorin and Marsden [3, p. 2]).
2.3 Conservation principles
Our derivation of the basic equations underlying the dynamics of uids is based on
three basic principles:
1. Conservation of mass, mass is neither created or destroyed;
2. Newtons 2nd law/balance of momentum, for a parcel of uid the rate of change of
momentum equals the force applied to it;
3. Conservation of energy, energy is neither created nor destroyed.
In turn these principles generate the:
1. Continuity equation which governs how the density of the uid evolves locally and
thus indicates compressibility properties of the uid;
2. NavierStokes equations of motion for a uid which indicates how the uid moves
around from regions of high pressure to those of low pressure and under the eects
of viscosity;
3. Equation of state which indicates the mechanism of energy exchange within the
uid.
3 Trajectories and streamlines
Suppose that our uid is contained with a region/domain D R
d
where d = 2 or
3, and x = (x, y, z) D is a position/point in D. Imagine a small uid particle or
a speck of dust moving in a uid ow eld prescribed by the velocity eld u(x, t) =
(u, v, w). Suppose the position of the particle at time t is recorded by the variables
_
x(t), y(t), z(t)
_
. The velocity of the particle at time t at position
_
x(t), y(t), z(t)
_
is
x(t) = u
_
x(t), y(t), z(t), t
_
,
y(t) = v
_
x(t), y(t), z(t), t
_
,
z(t) = w
_
x(t), y(t), z(t), t
_
.
4 Simon J.A. Malham
In shorter vector notation this is
d
dt
x(t) = u(x(t), t).
The trajectory or particle path of a uid particle is the curve traced out by the particle
as time progresses. It is the solution to the dierential equation above (with suitable
initial conditions).
Suppose now for a given uid ow u(x, t) we x time t. A streamline is an integral
curve of u(x, t) for t xed, i.e. it is a curve x = x(s) parameterized by the variable s,
that satises the system of equations
d
ds
x(s) = u(x(s), t),
with t held constant. If the velocity eld u is time-independent, i.e. u = u(x) only,
or equivalently
t
u = 0, then trajectories and streamlines coincide. Flows for which

t
u = 0 are said to be stationary.
Example. Suppose a velocity eld u(x, t) = (u, v, w) is given for t > 1 by
u =
x
1 +t
, v =
y
1 +
1
2
t
and w = z.
To nd the particle paths or trajectories, we must solve the system of equations
dx
dt
= u,
dy
dt
= v and
dz
dt
= w,
and then eliminate the time variable t between them. Hence for the particle paths we
have
dx
dt
=
x
1 +t
,
dy
dt
=
y
1 +
1
2
t
and
dz
dt
= z.
Using the method of separation of variables and integrating in time from t
0
to t, in
each of the three equations, we get
ln
_
x
x
0
_
= ln
_
1 +t
1 +t
0
_
, ln
_
y
y
0
_
= 2 ln
_
1 +
1
2
t
1 +
1
2
t
0
_
and ln
_
z
z
0
_
= t t
0
,
where we have assumed that at time t
0
the particle is at position (x
0
, y
0
, z
0
). Expo-
nentiating the rst two equations and solving the last one for t, we get
x
x
0
=
1 +t
1 +t
0
,
y
y
0
=
(1 +
1
2
t)
2
(1 +
1
2
t
0
)
2
and t = t
0
+ ln(z/z
0
).
We can use the last equation to eliminate t so the particle path/trajectory through
(x
0
, y
0
, z
0
) is the curve in three dimensional space given by
x = x
0

_
1 +t
0
+ ln(z/z
0
)
_
(1 +t
0
)
, and y = y
0

_
1 +
1
2
t
0
+
1
2
ln(z/z
0
)
_
2
(1 +
1
2
t
0
)
2
.
To nd the streamlines, we x time t. We must then solve the system of equations
dx
ds
= u,
dy
ds
= v and
dz
ds
= w,
Introductory uid mechanics 5
with t xed, and then eliminate s between them. Hence for streamlines we have
dx
ds
=
x
1 +t
,
dy
ds
=
y
1 +
1
2
t
and
dz
ds
= z.
Assuming that we are interested in the streamline that passes through the point
(x
0
, y
0
, z
0
), we again use the method of separation of variables and integrate with
respect to s from s
0
to s, for each of the three equations. This gives
ln
_
x
x
0
_
=
s s
0
1 +t
, ln
_
y
y
0
_
=
s s
0
1 +
1
2
t
and ln
_
z
z
0
_
= s s
0
.
Using the last equation, we can substitute for s s
0
into the rst equations. If we then
multiply the rst equation by 1 + t and the second by 1 +
1
2
t, and use the usual log
law ln a
b
= b ln a, then exponentiation reveals that
_
x
x
0
_
1+t
=
_
y
y
0
_
1+
1
2
t
=
z
z
0
,
which are the equations for the streamline through (x
0
, y
0
, z
0
).
4 Conservation of mass
4.1 Continuity equation
Recall, we suppose our uid is contained with a region/domain D R
d
(here we will
assume d = 3, but everything we say is true for the collapsed two dimensional case
d = 2). Hence x = (x, y, z) D is a position/point in D. At each time t we will suppose
that the uid has a well dened mass density (x, t) at the point x. Further, each uid
particle traces out a well dened path in the uid, and its motion along that path is
governed by the velocity eld u(x, t) at position x at time t. Consider an arbitrary
subregion D. The total mass of uid contained inside the region at time t is
_

(x, t) dV.
where dV is the volume element in R
d
. Let us now consider the rate of change of mass
inside . By the principle of conservation of mass, the rate of increase of the mass in
is given by the mass of uid entering/leaving the boundary of per unit time.
To compute the total mass of uid entering/leaving the boundary per unit time,
we consider a small area patch dS on the boundary of , which has unit outward
normal n. The total mass of uid owing out of through the area patch dS per unit
time is
mass density uid volume leaving per unit time = (x, t) u(x, t) n(x) dS,
where x is at the center of the area patch dS on . Note that to estimate the uid
volume leaving per unit time we have decomposed the uid velocity at x , time t,
into velocity components normal (u n) and tangent to the surface at that point.
The velocity component tangent to the surface pushes uid across the surfaceno uid
6 Simon J.A. Malham

D
Fig. 1 The uid of mass density (x, t) swirls around inside the container D, while is an
imaginary subregion.
dS
n u
u.n
Fig. 2 The total mass of uid moving through the patch dS on the surface per unit time,
is given by the mass density (x, t) times the volume of the cylinder shown which is u ndS.
enters or leaves via this component. Hence we only retain the normal component
see Fig. 2.
Returning to the principle of conservation of mass, this is now equivalent to the
integral form of the law of conservation of mass:
d
dt
_

(x, t) dV =
_

u ndS.
The divergence theorem and that the rate of change of the total mass inside equals
the total rate of change of mass density inside imply, respectively,
_

(u) dV =
_

(u) ndS and


d
dt
_

dV =
_

t
dV.
Using these two relations, the law of conservation of mass is equivalent to
_

t
+ (u) dV = 0.
Now we use that is arbitrary to deduce the dierential form of the law of conservation
of mass or continuity equation that applies pointwise:

t
+ (u) = 0.
This is the rst of our three conservation laws.
Introductory uid mechanics 7
4.2 Incompressible ow
Having established the continuity equation we can now dene a subclass of ows which
are incompressible. The classic examples are water, and the brake uid in your car whose
incompressibility properties are vital to the eective transmission of pedal pressure to
brakepad pressure.
Denition 1 (Incompressibility) A uid with the property u = 0 is incompressible.
The continuity equation and the identity, ( u) = u + u, imply

t
+u + u = 0.
Hence since > 0, a ow is incompressible if and only if

t
+u = 0.
If the uid is homogeneous so that is constant in space, then the ow is incompressible
if and only if is constant in time.
4.3 Stream functions
A stream function exists for a given ow u = (u, v, w) if the velocity eld u is solenoidal,
i.e. u = 0, and we have an additional symmetry that allows us to eliminate one
coordinate. For example, a two dimensional incompressible uid ow u = u(x, y, t) is
solenoidal since u = 0, and has the symmetry that it is uniform with respect to z.
For such a ow we see that
u = 0
u
x
+
v
y
= 0.
This equation is satised if and only if there exists a function (x, y, t) such that

y
= u(x, y, t) and

x
= v(x, y, t).
The function is called Lagranges stream function. A stream function is always only
dened up to any arbitrary additive constant. Further note that for t xed, streamlines
are given by constant contour lines of (note u = 0 everywhere).
Note that if we use plane polar coordinates so u = u(r, , t) and the velocity
components are u = (u
r
, u

) then
u = 0
1
r

r
(r u
r
) +
1
r
u

= 0.
This is satised if and only if there exists a function (r, , t) such that
1
r

= u
r
(r, , t) and

r
= u

(r, , t).
Example Suppose that in Cartesian coordinates we have the two dimensional ow
u = (u, v) given by
(u, v) = (k x, k y),
8 Simon J.A. Malham
for some constant k. Note that u = 0 so there exists a stream function satisfying

y
= k x and

x
= k y.
Consider the rst partial dierential equation. Integrating with respect to y we get
= k xy +C(x)
where C(x) is an arbitrary function of x. However we know that must simultaneously
satisfy the second partial dierential equation above. Hence we substitute this last
relation into the second partial dierential equation above to get

x
= k y k y +C

(x) = k y.
We deduce C

(x) = 0 and therefore C is an arbitrary constant. Since a stream function


is only dened up to an arbitrary constant we take C = 0 for simplicity and the stream
function is given by
= k xy.
Now suppose we used plane polar coordinates instead. The corresponding ow
u = (u
r
, u

) is given by
(u
r
, u

) = (k r cos 2, k r sin 2).


First note that u = 0 using the polar coordinate form for u indicated above.
Hence there exists a stream function = (r, ) satisfying
1
r

= k r cos 2 and

r
= k r sin 2.
As above, consider the rst partial dierential equation shown, and integrate with
respect to to get
=
1
2
k r
2
sin 2 +C(r).
Substituting this into the second equation above reveals that C

(r) = 0 so that C is a
constant. We can for convenience set C = 0 so that
=
1
2
k r
2
sin 2.
Comparing this form with its Cartesian equivalent above, reveals they are the same.
5 Transport theorem
Recall our image of a small uid particle moving in a uid ow eld prescribed by the
velocity eld u(x, t). The velocity of the particle at time t at position x(t) is
d
dt
x(t) = u(x(t), t).
As the particle moves in the velocity eld u(x, t), say from position x(t) to a nearby
position an instant in time later, two dynamical contributions change: (i) a small instant
in time has elapsed and the velocity eld u(x, t), which depends on time, will have
changed a little; (ii) the position of the particle has changed in that short time as it
Introductory uid mechanics 9
moved slightly, and the velocity eld u(x, t), which depends on position, will be slightly
dierent at the new position.
Let us compute the acceleration of the particle to explicitly observe these two
contributions. By using the chain rule we see that
d
2
dt
2
x(t) =
d
dt
u
_
x(t), t
_
=
u
x
dx
dt
+
u
y
dy
dt
+
u
z
dz
dt
+
u
t
=
_
dx
dt

x
+
dy
dt

y
+
dz
dt

z
_
u +
u
t
= u u +
u
t
.
Indeed for any function F(x, y, z, t), scalar or vector valued, the chain rule implies
d
dt
F
_
x(t), y(t), z(t), t
_
=
F
t
+u F.
Denition 2 (Material derivative) If the velocity eld components are
u = (u, v, w) and u u

x
+v

y
+w

z
,
then we dene the material derivative following the uid to be
D
Dt
:=

t
+u .
Suppose that the region within which the uid is moving is D. Suppose is a
subregion of D identied at time t = 0. As the uid ow evolves the uid particles that
originally made up will subsequently ll out a volume
t
at time t. We think of
t
as the volume moving with the uid.
Theorem 1 (Transport theorem) For any function F and density function satisfying
the continuity equation, we have
d
dt
_

t
F dV =
_

DF
Dt
dV.
We will use the transport theorem to deduce Cauchys equation of motion from the
primitive integral form of the balance of momentum; see Section 6.
Proof There are four steps; see Chorin and Marsden [3, pp. 611].
Step 1: Fluid ow map. For a xed position x D we denote by (x, t) = (, , )
the position of the particle at time t, which at time t = 0 was at x. We use
t
to denote
the map x (x, t), i.e.
t
is the map that advances each particle at position x at
time t = 0 to its position at time t later; it is the uid ow-map. Hence, for example

t
() =
t
. We assume
t
is suciently smooth and invertible for all our subsequent
manipulations.
Step 2: Change of variables. For any two functions and F we can perform the
change of variables from (, t) to (x, t)with J(x, t) the Jacobian for this transforma-
tion given by denition as J(x, t) := det
_
(x, t)
_
. Here the gradient operator is with
10 Simon J.A. Malham
respect to the x coordinates, i.e. =
x
. Note for
t
we integrate over volume ele-
ments dV = dV (), i.e. with respect to the coordinates, whereas for we integrate
over volume elements dV = dV (x), i.e. with respect to the xed coordinates x. Hence
by direct computation
d
dt
_

t
F dV =
d
dt
_

t
( F)(, t) dV ()
=
d
dt
_

( F)((x, t), t) J(x, t) dV (x)


=
_

d
dt
_
( F)((x, t), t) J(x, t)
_
dV
=
_

d
dt
( F)((x, t), t) J(x, t) + ( F)((x, t), t)
d
dt
J(x, t) rdV
=
_

_
D
Dt
( F)
_
((x, t), t) J(x, t) + ( F)((x, t), t)
d
dt
J(x, t) dV.
Step 3: Evolution of the Jacobian. We establish the following result for the Jacobian:
d
dt
J(x, t) =
_
u((x, t), t)
_
J(x, t).
We know that a particle at position (x, t) =
_
(x, t), (x, t), (x, t)
_
, which started at
x at time t = 0, evolves according to
d
dt
(x, t) = u
_
(x, t), t
_
.
Taking the gradient with respect to x of this relation, and swapping over the gradient
and d/dt operations on the left, we see that
d
dt
(x, t) = u
_
(x, t), t
_
.
Using the chain rule we have

x
u
_
(x, t), t
_
=
_

u
_
(x, t), t
_
_

x
(x, t)
_
.
Combining the last two relations we see that
d
dt
= (

u) .
Abels Theorem then tells us that J = det evolves according to
d
dt
det =
_
Tr(

u)
_
det ,
where Tr denotes the trace operator on matricesthe trace of a matrix is the sum of
its diagonal elements. Since Tr(

u) u we have established the required result.


Step 4: Conservation of mass. We see that we thus have
d
dt
_

t
F dV =
_

_
D
Dt
( F) + ( F)
_
u
_
_
((x, t), t) J(x, t) dV
=
_

t
_
D
Dt
( F) +
_
u
_
F
_
dV
=
_

DF
Dt
dV,
where in the last step we have used the conservation of mass equation.
Introductory uid mechanics 11
Corollary 1 (Equivalent incompressibility statements) The following statements are
equivalent, for any subregion of the uid, the:
1. Fluid is incompressible;
2. Jacobian J 1;
3. Volume of
t
is constant in time.
Proof Using the result in Step 3 of the proof of the transport theorem, we see that
d
dt
vol(
t
) =
d
dt
_

t
dV ()
=
d
dt
_

J(x, t) dV (x)
=
_

_
u((x, t), t)
_
J(x, t) dV (x)
=
_

t
_
u(, t)
_
dV ().
Further, noting that by denition J(x, 0) = 1, establishes the result.
6 Balance of momentum
6.1 Rate of strain tensor
Consider a uid ow in a region D R
3
. Suppose x and x +h are two nearby points
in the interior of D. How is the ow, or more precisely the velocity eld, at x related
to that at x +h? From a mathematical perspective, by Taylor expansion we have
u(x +h) = u(x) +
_
u(x)
_
h +O(h
2
),
where (u) h is simply matrix multiplication of the 3 3 matrix u by the column
vector h. Recall that u is given by
u =
_
_
u/x u/y u/z
v/x v/y v/z
w/x w/y w/z
_
_
.
In the context of uid ow it is known as the rate of strain tensor. This is because,
locally, it measures that rate at which neighbouring uid particles are being pulled
apart (it helps to recall that the velocity eld u records the rate of change of particle
position with respect to time).
Again from a mathematical perspective, we can decompose u as follows. We can
always write
u =
1
2
_
(u) + (u)
T
_
+
1
2
_
(u) (u)
T
_
.
We set
D :=
1
2
_
(u) + (u)
T
_
,
R :=
1
2
_
(u) (u)
T
_
.
12 Simon J.A. Malham
Note that D = D(x) is a 3 3 symmetric matrix, while R = R(x) is the 3 3
skew-symmetric matrix given by
R =
_
_
0 u/y v/x u/z w/x
v/x u/y 0 v/z w/y
w/x u/z w/y v/z 0
_
_
.
Note that if we set

1
=
w
y

v
z
,
2
=
u
z

w
x
and
3
=
v
x

u
y
,
then R is more simply expressed as
R =
1
2
_
_
0
3

2

3
0
1

2

1
0
_
_
.
Further by direct computation we see that
Rh =
1
2
h,
where = (x) is the vector with three components
1
,
2
and
3
. At this point, we
have thus established the following.
Theorem 2 If x and x +h are two nearby points in the interior of D, then
u(x +h) = u(x) +D(x) h +
1
2
(x) h +O(h
2
).
The symmetric matrix D is the deformation tensor. Since it is symmetric, there is
an orthonormal basis e
1
, e
2
, e
3
in which D is diagonal, i.e. if X = [e
1
, e
2
, e
3
] then
X
1
DX =
_
_
d
1
0 0
0 d
2
0
0 0 d
3
_
_
.
The vector is the vorticity eld of the ow. An equivalent denition for it is
= u.
It encodes the magnitude of, and direction of the axis about which, the uid rotates,
locally.
Now consider the motion of a uid particle labelled by x +h where x is xed and
h is small (for example suppose that only a short time has elapsed). Then the position
of the particle is given by
d
dt
(x +h) = u(x +h)

dh
dt
= u(x +h)

dh
dt
u(x) +D(x) h +
1
2
(x) h.
Let us consider in turn each of the eects on the right shown:
Introductory uid mechanics 13
1. The term u(x) is simply uniform translational velocity (the particle being pushed
by the ambient ow surrounding it).
2. Now consider the second term D(x) h. If we ignore the other terms then, approx-
imately, we have
dh
dt
= D(x) h.
Making a local change of coordinates so that h = X

h we get
d
dt
_
_

h
1

h
2

h
3
_
_
=
_
_
d
1
0 0
0 d
2
0
0 0 d
3
_
_
_
_

h
1

h
2

h
3
_
_
.
We see that we have pure expansion or contraction (depending on whether d
i
is positive or negative, respectively) in each of the characteristic directions

h
i
,
i = 1, 2, 3. Indeed the small linearized volume element

h
1

h
2

h
3
satises
d
dt
(

h
1

h
2

h
3
) = (d
1
+d
2
+d
3
)(

h
1

h
2

h
3
).
Note that d
1
+d
2
+d
3
= Tr(D) = u.
3. Let us now examine the eect of the third term
1
2
h. Ignoring the other two
terms we have
dh
dt
=
1
2
(x) h.
Direct computation shows that
h(t) = (t, (x))h(0),
where (t, (x)) is the matrix that represents the rotation through an angle t about
the axis (x). Note also that
_
(x) h
_
= 0.
6.2 Internal uid forces
Let us consider the forces that act on a small parcel of uid in a uid ow. There are
two types:
1. external or body forces, these may be due to gravity or external electromagnetic
elds. They exert a force per unit volume on the continuum.
2. surface or stress forces, these are forces, molecular in origin, that are applied by
the neighbouring uid across the surface of the uid parcel.
The surface or stress forces are normal stresses due to pressure dierentials, and shear
stresses which are the result of molecular diusion. We explain shear stresses as follows.
Imagine two neighbouring parcels of uid P and P

as shown in Fig. 3, with a mutual


contact surface is S as shown. Suppose both parcels of uid are moving parallel to S
and to each other, but the speed of P, say u, is much faster than that of P

, say u

.
In the kinetic theory of matter molecules jiggle about and take random walks; they
diuse into their surrounding locale and impart their kinetc energy to molecules they
pass by. Hence the faster molecules in P will diuse across S and impart momentum
to the molecules in P

. Similarly, slower molecules from P

will diuse across s to slow


the uid in P down. In regions of the ow where the velocity eld changes rapidly over
small length scales, this eect is importantsee Chorin and Marsden [3, p. 31].
14 Simon J.A. Malham
P
P
u
u
S
Fig. 3 Two neighbouring parcels of uid P and P

. Suppose S is the surface of mutual contact


between them. Their respective velocities are u and u

and in the same direction and parallel


to S, but with |u| |u

|. The faster molecules in P will diuse across the surface S and


impart momentum to P

.
dS
n
dF
x
(1)
(2)
Fig. 4 The force dF on side (2) by side (1) of dS is given by (n) dS.
We now proceed more formally. The force per unit area exerted across a surface
(imaginary in the uid) is called the stress. Let dS be a small imaginary surface in the
uid centered on the point xsee Fig. 4. The force dF on side (2) by side (1) of dS
in the uid/material is given by
dF = (n) dS.
Here is the stress at the point x. It is a function of the normal direction n to the
surface dS, in fact it is given by:
(n) = (x) n.
Note = [
ij
] is a 3 3 matrix known as the stress tensor. The diagonal components
of
ij
, with i = j, generate normal stresses, while the o-diagonal components, with
i = j, generate tangential or shear stresses. Indeed let us decompose the stress tensor
= (x) as follows (here I is the 3 3 identity matrix):
= p I + .
Here the scalar quantity p = p(x) > 0 is dened to be
p :=
1
3
(
11
+
22
+
33
)
and represents the uid pressure. The remaining part of the stress tensor = (x) is
known as the deviatoric stress tensor. In this decomposition, the term p I generates
the normal stresses, since if this were the only term present,
= p I (n) = p n.
The deviatoric stress tensor on the other hand, generates the shear stresses.
We assume that the deviatoric stress tensor is a function of the rate of strain
tensor u. We shall make three assumptions about the deviatoric stress tensor and
its dependence on the velocity gradients u. These are that it is:
Introductory uid mechanics 15
1. Linear: each component of is linearly related to the rate of strain tensor u.
2. Isotropic: if U is an orthogonal matrix, then

_
U u U
1
_
U (u) U
1
.
Equivalently we might say that it is invariant under rigid body rotations.
3. Symmetric; i.e.
ij
=
ji
. This can be deduced as a result of balance of angular
momentum.
Hence each component of the deviatoric stress tensor is a linear function of each
of the components of the velocity gradients u. This means that there is a total of
81 constants of proportionality. We will use the assumptions above to systematically
reduce this to 2 constants.
When the uid performs rigid body rotation, there should be no diusion of momen-
tum (the whole mass of uid is behaving like a solid body). Recall our decomposition
of the rate of strain tensor, u = D + R, where D is the deformation tensor and R
generates rotation. Thus only depends on the symmetric part of u, i.e. it is a linear
function of the deformation tensor D. Further, since is symmetric, we can restrict our
attention to linear functions from symmetric matrices to symmetric matrices. We now
lean heavily on the isotropy assumption 2; see Gurtin [7, Section 37] for more details.
First, we have the transfer theorem. Let Sym
2
(R
3
) denote the set of 33 symmetric
matrices.
Theorem 3 (Transfer theorem) Let be an endomorphism on Sym
2
(R
3
). Then if
is isotropic, the matrices D Sym
2
(R
3
) and (D) Sym
2
(R
3
) are simultaneously
diagonalizable.
Proof Let e be an eigenvector of D and let U be the orthogonal matrix denoting
reection in the plane perpendicular to e, so that Ue = e, while any vector per-
pendicular to e is invariant under U. The eigenstructure of D is invariant to such
a transformation so that UDU
1
= D. Thus, since = (D) is isotropic, we have
U U
1
= (UDU
1
) = (D) and thus U = U. Any such commuting matrices
share eigenvectors since U e = Ue = e. Thus e is also an eigenvector of the
reection transformation U corresponding to the same eigenvalue 1. Thus e is pro-
portional to e and so e is an eigenvector of . Since e was any eigenvector of D, the
statement of the theorem follows.
Second, for any matrix A R
33
with eigenvalues
1
,
2
,
3
, the three scalar functions
I
1
(A) := Tr A, I
2
(A) :=
1
2
_
(TrA)
2
Tr(A
2
)
_
and I
2
(A) := det A,
are isotropic. This can be checked by direct computation. Indeed these three functions
are the elementary symmetric functions of the eigenvalues of A:
I
1
(A) =
1
+
2
+
3
, I
2
(A) =
1

2
+
2

3
+
2

3
and I
2
(A) =
1

3
.
We have the following representation theorem for isotropic functions.
Theorem 4 (Representation theorem) An endomorphism on Sym
2
(R
3
) is isotropic
if and only if it has the form
(D) =
0
I +
1
D +
2
D
2
,
for every D Sym
2
(R
3
), where
0
,
1
and
2
are scalar functions that depend only
on the isotropic invariants I
1
(D), I
2
(D) and I
3
(D).
16 Simon J.A. Malham
Proof Scalar functions = (D) are isotropic if and only if they are functions of the
isotropic invariants of D only. The if part of this statement follows trivially as the
isotropic invariants are isotropic. The only if statement is established if, assuming
is isotropic, we are able to show that
I
i
(D) = I
i
(D

) for i = 1, 2, 3 = (D) = (D

).
Since the map between the eigenvalues of D and its isotropic invariants is bijective,
if I
i
(D) = I
i
(D

) for i = 1, 2, 3, then D and D

have the same eigenvalues. Since


the isospectral action UDU
1
of orthogonal matrices U on symmetric matrices D is
transitive, there exists an orthogonal matrix U such that D

= UDU
1
. Since is
isotropic, (UDU
1
) = (D), i.e. (D

) = (D).
Now let us consider the symmetric matrix valued function . The if statement of
the theorem follows by direct computation and the result we just established for scalar
isotropic functions. The only if statement is proved as follows. Assume has three
distinct eigenvalues (we leave the other possibilities as an exercise). Using the transfer
theorem and the Spectral Theorem (see for example Meyer [15, p. 517]) we have
(D) =
3

i=1

i
E
i
where
1
,
2
and
3
are the eigenvalues of and the projection matrices E
1
, E
2
and
E
3
have the properties E
i
E
j
= O when i = j and E
1
+E
2
+E
3
= I. Since we have
span{I, D, D
2
} = span{E
1
, E
2
, E
3
},
there exist scalars
0
,
1
and
2
depending on D such that
(D) =
0
I +
1
D +
2
D
2
.
We now have to show that
0
,
1
and
2
are isotropic. This follows by direct compu-
tation, combining this last representation with the property that is isotropic.
Remark 1 Note that neither the transfer theorem nor the representation theorem re-
quire that the endomorphism is linear.
Third, now suppose that is a linear function of D. Thus for any symmetric matrix
D it must have the form
(D) = I + 2D,
where the scalars and depend on the isotropic invariants of D. By the Spectral
Theorem we have
D =
3

i=1
d
i
E
i
,
where d
1
, d
2
and d
3
are the eigenvalues of D and E
1
, E
2
and E
3
are the correspond-
ing projection matricesin particular each E
i
is symmetric with an eigenvalue 1 and
double eigenvalue 0. Since is linear we have
(D) =
3

i=1
d
i
(E
i
)
=
3

i=1
d
i
_
I + 2E
i
_
.
Introductory uid mechanics 17
where for each i = 1, 2, 3 the only non-zero isotropic invariant is I
1
(E
i
) = 1 so that
and are simply constant scalars. Using that E
1
+E
2
+E
3
= I we have
= (d
1
+d
2
+d
3
)I + 2D.
Recall that d
1
+d
2
+d
3
= u. Thus we have
= ( u)I + 2D.
If we set = +
2
3
this last relation becomes
= 2
_
D
1
3
( u)I
_
+( u)I,
where and are the rst and second coecients of viscosity, respectively.
Remark 2 Note that if u = 0, then the linear relation between and D is homog-
neous, and we have the key property of what is known as a Newtonian uid: the stress
is proportional to the rate of strain.
6.3 NavierStokes equations
Consider again an arbitrary imaginary subregion of D identied at time t = 0, as in
Fig. 1. As the uid ow evolves to some time t > 0, let
t
denote the volume of the
uid occupied by the particles that originally made up . The total force exerted on
the uid inside
t
through the stresses exerted across its boundary
t
is given by
_

t
(pI + ) ndS
_

t
(p + ) dV,
where (for convenience here we set (x
1
, x
2
, x
3
) (x, y, z) and (u
1
, u
2
, u
3
) (u, v, w))
[ ]
i
=
3

j=1

ij
x
j
= [( u)]
i
+ 2
3

j=1
D
ij
x
j
= [( u)]
i
+
3

j=1

x
j
_
u
i
x
j

u
j
x
i
_
= [( u)]
i
+
3

j=1

2
u
i
x
2
j


2
u
j
x
i
x
j
= ( +)[( u)]
i
+
2
u
i
.
If f is a body force (external force) per unit mass, which can depend on position and
time, then the body force on the uid inside
t
is
_

t
f dV.
18 Simon J.A. Malham
Thus on any parcel of uid
t
, the total force acting on it is
_

t
p + + f dV.
Hence using Newtons 2nd law (force = mass acceleration) we have
d
dt
_

t
udV =
_

t
p + + f dV.
Now we use the transport theorem with F u and that and thus
t
are arbitrary.
We see that for at each x D and t 0, we can deduce the following relation
Cauchys equation of motionthe dierential form of the balance of momentum:

Du
Dt
= p + + f.
Combining this with the form for we deduced above, we arrive at

Du
Dt
= p + ( +)( u) +u + f,
where =
2
is the Laplacian operator. These are the NavierStokes equations. If we
assume we are in three dimensional space so d = 3, then together with the continuity
equation we have four equations, but ve unknownsnamely u, p and . Thus for a
compressible uid ow, we cannot specify the uid motion completely without specify-
ing one more condition/relation. (We could use the principle of conservation of energy
to establish as additional relation known as the equation of state; in simple scenarios
this takes the form of relationship between the pressure p and density of the uid.)
For an incompressible homogeneous ow for which the density =
0
is constant,
we get a complete set of equations known as the NavierStokes equations for an in-
compressible ow:
u
t
+u u = u p +f,
u = 0,
where = /
0
is the coecient of kinematic viscosity. Note that the pressure eld
here is the rescaled pressure by a factor 1/
0
: since
0
is constant (p)/
0
(p/
0
),
and we re-label the term p/
0
to be p. Note that we have a closed system of equations:
we have four equations in four unknowns, u and p.
For any motion of an ideal uid we only include normal stresses and completely
ignore any shear stresses. Hence instead of the the NavierStokes equation above we
get the Euler equations of motion for an ideal uid (derived by Euler in 1755) given
by (take = = 0):
u
t
+u u =
1

p +f,
The fact that there are no tangential forces in an ideal uid has some important
consequences, quoting from Chorin and Marsden [3, p. 5]:
...there is no way for rotation to start in a uid, nor, if there is any at the
beginning, to stop... ... even here we can detect trouble for ideal uids because
of the abundance of rotation in real uids (near the oars of a rowboat, in
tornadoes, etc. ).
We discuss the Euler equations in more detail in Section 13.2.
Introductory uid mechanics 19
6.4 Boundary conditions
Now that we have the partially dierential equations that determine how uid ows
evolve, we complement them with the boundary and initial conditions. The initial
condition is the velocity prole u = u(x, 0) at time t = 0. It is the state in which
the ow starts. To have a well-posed evolutionary partial dierential system for the
evolution of the uid ow, we also need to specify how the ow behaves near boundaries.
Here a boundary could be a rigid boundary, for example the walls of the container the
uid is conned to or the surface of an obstacle in the uid ow. Another example of a
boundary is the free surface between two immiscible uidssuch as between seawater
and air on the ocean surface. Here we will focus on rigid boundaries.
For an ideal uid ow, i.e. one evolving according to the Euler equations, we simply
need to specify that there is no net ow normal to the boundarythe uid does not
cross the boundary but can move tangentially to it. Mathematically this is means that
we specify that u n = 0 everywhere on the rigid boundary.
For viscous ow, i.e. evolving according to the NavierStokes equations, we need
to specify additional boundary conditions. This is due to the inclusion of the extra
term u which increases the number of spatial derivatives in the governing evolution
equations from one to two. Mathematically, we specify that
u = 0
everywhere on the rigid boundary, i.e. in addition to the condition that there must be no
net normal ow at the boundary, we also specify there is no tangential ow there. The
uid velocity is simply zero at a rigid boundary; it is sometimes called no-slip boundary
conditions. Experimentally this is observed as well, to a very high degree of precision;
see Chorin and Marsden [3, p. 34]. (Dye can be introduced into a ow near a boundary
and how the ow behaves near it observed and measured very accurately.) Further,
recall that in a viscous uid ow we are incorporating the eect of molecular diusion
between neighbouring uid parcelssee Fig. 3. The rigid non-moving boundary should
impart a zero tangential ow condition to the uid particles right up against it. The no-
slip boundary condition is crucially repsresents the mechanism for vorticity production
in nature that can be observed everywhere. Just look at the ow of a river close to the
river bank.
Remark 3 At a material boundary (or free surface) between two immiscible uids, we
would specify that there is no jump in the velocity across the surface boundary. This
is true if there is no surface tension or at least if it is negligiblefor example at the
seawater-air boundary of the ocean. However at the surface of melting wax at the top of
a candle, there is surface tension, and there is a jump in the stress n at the boundary
surface. Surface tension is also responsible for the phenomenon of being able to oat a
needle on the surface of a bowl of water as well as many other interesting eects such
as the shape of water drops.
6.5 Evolution of vorticity
Recall from our discussion in Section 6.1, that the vorticity eld of a ow with velocity
eld u is dened as
:= u.
20 Simon J.A. Malham
It encodes the magnitude of, and direction of the axis about which, the uid rotates,
locally. Note that u can be computed as follows
u = det
_
_
i j k
/x /y /z
u v w
_
_
=
_
_
w/y v/z
u/z w/x
v/x u/y
_
_
.
Using the NavierStokes equations for a homogeneous incompressible uid, we can in
fact derive a closed system of equations governing the evolution of vorticity = u
as follows. Using the identity u u =
1
2

_
|u|
2
_
u ( u) we see that we can
equivalently represent the NavierStokes equations in the form
u
t
+
1
2

_
|u|
2
_
u = u p +f.
If we take the curl of this equation and use the identity
(u ) = u( ) ( u) + ( )u (u ),
noting that u = 0 and = (u) 0, we nd that we get

t
+u = + u +f.
Note that we can recover the velocity eld u from the vorticity by using the identity
(u) = ( u) u. This implies
u = ,
and closes the system of partial dierential equations for and u. However, we can
also simply observe that
u =
_

_
1
().
If the body force is conservative so that f = for some potential , then f 0.
Remark 4 We can replace the vortex stretching term u in the evolution equation
for the vorticity by D, where D is the 3 3 deformation matrix, since
u = (u) = D +R = D,
as direct computation reveals that R 0.
7 Simple example ows
We roughly follow an illustrative sequence of examples given in Majda and Bertozzi [13,
pp. 815]. The rst few are example ows of a class of exact solutions to both the Euler
and NavierStokes equations.
Introductory uid mechanics 21
Lemma 1 (Majda and Bertozzi, p. 8) Let D = D(t) R
3
be a real symmetric matrix
such that Tr(D) = 0 (respresenting the deformation matrix). Suppose that the vorticity
= (t) solves the ordinary dierential system
d
dt
= D(t)
for some initial data (0) =
0
R
3
. If the three components of vorticity are thus
= (
1
,
2
,
3
), set
R :=
1
2
_
_
0
3

2

3
0
1

2

1
0
_
_
.
Then we have that
u(x, t) =
1
2
(t) x +D(t) x,
p(x, t) =
1
2
_
dD
dt
+D
2
(t) +R
2
(t)
_
x x,
are exact solutions to the incompressible Euler and NavierStokes equations.
Remark 5 Since the pressure is a quadratic function of the spatial coordinates x, these
solutions only have meaningful interpretations locally. Further note that the velocity
solution eld u only depends linearly on the spatial coordinates x; this explains why
once we established these are exact solutions of the Euler equations, they are also
solutions of the NavierStokes equations.
Proof Recall that u is the rate of strain tensor. It can be decomposed into a direct
sum of its symmetric and skew-symmetric parts which are the 3 3 matrices
D :=
1
2
_
(u) + (u)
T
_
,
R :=
1
2
_
(u) (u)
T
_
.
We can determine how u evolves by taking the gradient of the homogeneous (no
body force) NavierStokes equations so that

t
(u) +u (u) + (u)
2
= (u) p.
Note here (u)
2
= (u)(u) is simply matrix multiplication. By direct computation
(u)
2
= (D +R)
2
= (D
2
+R
2
) + (DR +RD),
where the rst term on the right is symmetric and the second is skew-symmetric. Hence
we can decompose the evolution of u into the coupled evolution of its symmetric and
skew-symmetric parts
D
t
+u D +D
2
+R
2
= D p,
R
t
+u R +DR +RD = R.
22 Simon J.A. Malham
Directly computing the evolution for the three components of = (
1
,
2
,
3
) from
the second system of equations we would arrive at the following equation for vorticity,

t
+u = +D,
which we derived more directly in Section 6.5.
Thusfar we have not utilized the ansatz (form) for the velocity or pressure we
assume in the statement of the theorem. Assuming u(x, t) =
1
2
(t) x + D(t) x,
for a given deformation matrix D = D(t), then u = (t), independent of x,
and substituting this into the evolution equation for = u above we obtain the
following system of ordinary dierential equations governing the evolution of = (t):
d
dt
= D(t).
Now the symmetric part governing the evolution of D = D(t), which is independent of
x, reduces to the system of dierential equations
dD
dt
+D
2
+R
2
= p.
Note that R = R(t) only as well, since = (t), and thus p must be a function of
t only. Hence p = p(x, t) can only quadratically depend on x. Indeed after integrating
we must have p(x, t) =
1
2
(dD/dt +D
2
+R
2
) x x.
Example (jet ow) Suppose the initial vorticity
0
= 0 and D = diag{d
1
, d
2
, d
3
}
is a constant diagonal matrix where d
1
+ d
2
+ d
3
= 0 so that Tr(D) = 0. Then from
Lemma 1, we see that the ow is irrotational, i.e. (t) = 0 for all t 0. Hence the
velocity eld u is given by
u(x, t) = D(t)x =
_
_
d
1
x
d
2
y
d
3
z
_
_
.
The particle path for a particle at (x
0
, y
0
, z
0
) at time t = 0 is given by: x(t) = e
d
1
t
x
0
,
y(t) = e
d
2
t
y
0
and z(t) = e
d
3
t
z
0
. If d
1
< 0 and d
2
< 0, then d
3
> 0 and we see the ow
resembles two jets streaming in opposite directions away from the z = 0 plane.
Example (strain ow) Suppose the initial vorticity
0
= 0 and D = diag{d
1
, d
2
, 0}
is a constant diagonal matrix such that d
1
+d
2
= 0. Then as in the last example, the
ow is irrotational with (t) = 0 for all t 0 and
u(x, t) =
_
_
d
1
x
d
2
y
0
_
_
.
The particle path for a particle at (x
0
, y
0
, z
0
) at time t = 0 is given by: x(t) = e
d
1
t
x
0
,
y(t) = e
d
2
t
y
0
and z(t) = z
0
. Since d
2
= d
1
, the ow forms a strain ow as shown in
Fig. 5neighbouring particles are pushed together in one direction while being pulled
apart in the other orthogonal direction.
Introductory uid mechanics 23
Fig. 5 Strain ow example.
Example (vortex) Suppose the initial vorticity
0
= (0, 0,
0
) and D = O. Then
from Lemma 1 the velocity eld u is given by
u(x, t) =
1
2
x =
_
_

1
2

0
y
1
2

0
x
0
_
_
.
The particle path for a particle at (x
0
, y
0
, z
0
) at time t = 0 is given by: x(t) =
cos(
1
2

0
t)x
0
sin(
1
2

0
t)y
0
, y(t) = sin(
1
2

0
t)x
0
+cos(
1
2

0
t)y
0
and z(t) = z
0
. These are
circular trajectories, and indeed the ow resembles a solid body rotation; see Fig. 6.
Fig. 6 When a uid ow is a rigid body rotation, the uid particles ow on circular streamlines.
The uid particles on paths further from the origin or axis of rotation, circulate faster at just
the right speed that they remain alongside their neighbours on the paths just inside them.
Example (jet ow with swirl) Now suppose the initial vorticity
0
= (0, 0,
0
) and
D = diag{d
1
, d
2
, d
3
} is a constant diagonal matrix where d
1
+d
2
+d
3
= 0. Then from
Lemma 1, we see that the only non-zero component of vorticity is the third component,
say = (t), where
(t) =
0
e
d
3
t
.
The velocity eld u is given by
u(x, t) =
_
_
d
1
x
1
2
(t)y
d
2
y +
1
2
(t)x
d
3
z
_
_
.
The particle path for a particle at (x
0
, y
0
, z
0
) at t = 0 can be described as follows.
We see that z(t) = z
0
e
d
3
t
while x = x(t) and y = y(t) satisfy the coupled system of
ordinary dierential equations
d
dt
_
x
y
_
=
_
d
1

1
2
(t)
1
2
(t) d
2
__
x
y
_
.
24 Simon J.A. Malham
If we assume d
1
< 0 and d
2
< 0 then the particles spiral around the z-axis with
decreasing radius and increasing angular velocity
1
2
(t). The ow thus resembles a
rotating jet ow; see Fig. 7.
x
y
z
Fig. 7 Jet ow with swirl example. Fluid particles rotate around and move closer to the z-axis
whilst moving further from the z = 0 plane.
Example (shear-layer ows) We derive a simple class of solutions that retain the
three underlying mechanisms of NavierStokes ows: convection, vortex stretching and
diusion. Recall that the vorticity evolves according to the partial dierential system

t
+u = +D,
with u = . The material derivative term /t +u convects vorticity
along particle paths, while the term is responsible for the diusion of vorticity
and Du represents vortex stretchingthe vorticity increases/decreases when aligns
along eigenvectors of D corresponding to positive/negative eigenvalues of D.
We seek an exact solution to the incompressible NavierStokes equations of the
following form (the rst two velocity components represent a strain ow)
u(x, t) =
_
_
x
y
w(x, t)
_
_
where is a constant, with p(x, t) =
1
2

_
x
2
+ y
2
_
. This represents a solution to the
NavierStokes equations if we can determine the solution w = w(x, t) to the linear
diusion equation
w
t
x
w
x
=

2
w
x
2
,
with w(x, 0) = w
0
(x). Computing the vorticity directly we get
(x, t) =
_
_
0

_
w/x
_
(x, t)
0
_
_
.
If we dierentiate the equation above for the velocity eld component w with respect
to x, then if := w/x, we get

t
x

x
= +

x
2
,
Introductory uid mechanics 25
with (x, 0) =
0
(x) = (w
0
/x)(x). For this simpler ow we can see simpler sig-
natures of the three eects we want to isolate: there is the convecting velocity x;
vortex stretching from the term and diusion in the term
2
/x
2
. Note that is
in the general case, the velocity eld w can be recovered from the vorticity eld by
w(x, t) =
_
x

(, t) d.
Let us consider a special case: the viscous shear-layer solution where = 0. In this
case we see that the partial dierential equation above for reduces to the simple heat
equation with solution
(x, t) =
_
R
G(x , t)
0
() d,
where G is the Gaussian heat kernel
G(, t) :=
1

4t
e

2
/4t
.
Indeed the velocity eld w is given by
w(x, t) =
_
R
G(x , t) w
0
() d,
so that both the vorticity and velocity w elds diuse as time evolves; see Fig. 8.
It is possible to write down the exact solution for the general case in terms of
the Gaussian heat kernel, indeed, a very nice exposition can be found in Majda and
Bertozzi [13, p. 18].
x
x
w(x,0)
w(x,t)
Fig. 8 Viscous shear ow example. The eect of diusion on the velocity eld w = w(x, t) is
to smooth out variations in the eld as time progresses.
Example (channel shear ow) Consider the two-dimensional ow given by u =
1 y
2
and v = 0 for 1 y 1 and all x R (which is an exact solution of the
incompressible NavierStokes equations). For this ow the vorticity is given by
u =
_
v
x

u
y
_
k = 2y k.
26 Simon J.A. Malham
See the shape of the ow in Fig. 9. The ow is stationary near the channel walls (no-slip
boundary conditions are satised there) and the ow rate a maximum in the middle
of the channel. The gradient of the horizontal velocity u with respect to y is non-zero
and thus the vorticity is non-zero (the vertical velocity component is zero).
x
y
y=1
y=+1
Fig. 9 Shear ow in a two-dimensional horizontal channel.
Example (sink or bath drain) As the water (of uniform density ) ows out through
a hole at the bottom of a bath the residual rotation is conned to a core of radius a,
so that the water particles may be taken to move on horizontal circles with
u

=
_
r, r a,
a
2
r
, r > a.
As we have all observed when water runs out of a bath or sink, the free surface of the
water directly over the drain hole has a depression in itsee Fig. 10. The question is,
what is the form/shape of this free surface depression?
r
z p
a
0
Fig. 10 Water draining from a bath.
We know that the pressure at the free surface is uniform, it is atmospheric pres-
sure, say P
0
. We need the Euler equations for a homogeneous incompressible uid in
Introductory uid mechanics 27
cylindrical coordinates (r, , z) with the velocity eld u = (u
r
, u

, u
z
). These are
u
r
t
+ (u )u
r

u
2

r
=
1

p
r
+f
r
,
u

t
+ (u )u

+
u
r
u

r
=
1
r
p

+f

,
u
z
t
+ (u )u
z
=
1

p
z
+f
z
,
where p = p(r, , z, t) is the pressure, is the uniform constant density and f =
(f
r
, f

, f
z
) is the body force per unit mass. Here we also have
u = u
r

r
+
u

+u
z

z
.
Further the incompressibility condition u = 0 is given in cylindrical coordinates by
1
r
(ru
r
)
r
+
1
r
u

+
u
z
z
= 0.
Now we look at the setting we are presented with for this problem. Note the ow
is steady and u
r
= u
z
= 0, f
r
= f

= 0. The force due to gravity implies f


z
= g.
The whole problem is also symmetric with respect to , so that all partial derivatives
with respect to should be zero. Combining all these facts reduces Eulers equations
above to

u
2

r
=
1

p
r
, 0 =
1
r
p

and 0 =
1

p
z
g.
The incompressibility condition is satised trivially. The second equation above tells
us the pressure p is independent of , as we might have already suspected. Hence we
assume p = p(r, z) and focus on the rst and third equation above.
Assume r a. Using that u

= r in the rst equation we see that


p
r
=
2
r p(r, z) =
1
2

2
r
2
+C(z),
where C(z) is an arbitrary function of z. If we then substitute this into the third
equation above we see that
1

p
z
= g C

(z) = g,
and hence C(z) = gz +C
0
where C
0
is an arbitrary constant. Thus we now deduce
that the pressure function is given by
p(r, z) =
1
2

2
r
2
gz +C
0
.
At the free surface of the water, the pressure is constant atmospheric pressure P
0
and
so if we substitute this into this expression for the pressure we see that
P
0
=
1
2

2
r
2
gz +C
0
z = (
2
/2g) r
2
(C
0
P
0
)/g.
Hence the depression in the free surface for r a is a parabolic surface of revolution.
Note that pressure is only ever globally dened up to an additive constant so we are
at liberty to take C
0
= 0 or C
0
= P
0
if we like.
28 Simon J.A. Malham
For r > a a completely analogous argument using u

= a
2
/r shows that
p(r, z) =

2
a
4
2 r
2
gz +K
0
,
where K
0
is an arbitrary constant. Since the pressure must be continuous at r = a, we
substitute r = a into the expression for the pressure here for r > a and the expression
for the pressure for r a, and equate the two. This gives

1
2

2
a
2
gz +K
0
=
1
2

2
a
2
gz K
0
=
2
a
2
.
Hence the pressure for r > a is given by
p(r, z) =

2
a
4
2 r
2
gz +
2
a
2
.
Using that the pressure at the free surface is p(r, z) = P
0
, we see that for r > a the
free surface is given by
z =

2
a
4
g r
2
+

2
a
2
g
.
8 Kelvins circulation theorem, vortex lines and tubes
We turn our attention to important concepts centred on vorticity in a ow.
Denition 3 (Circulation) Let C be a simple closed contour in the uid at time t = 0.
Suppose that C is carried along by the ow to the closed contour C
t
at time t, i.e.
C
t
=
t
(C). The circulation around C
t
is dened to be the line integral
K =
_
C
t
u dx.
Using Stokes Theorem an equivalent denition for the circulation is
K =
_
C
t
u dx =
_
S
(u) ndS =
_
S
ndS
where S is any surface with perimeter C
t
; see Fig. 12. In other words the circulation is
equivalent to the ux of vorticity through the surface with perimeter C
t
.
Theorem 5 (Kelvins circulation theorem (1869)) For ideal, incompressible ow with-
out external forces, the circulation K for any closed contour C
t
is constant in time.
Proof Using a variant of the Transport Theorem for closed loops of uid particles, and
the Euler equations, we see that
d
dt
_
C
t
u dx =
_
C
t
Du
Dt
dx =
_
C
t
p dx = 0
since C
t
is closed.
Introductory uid mechanics 29
Corollary 2 The ux of vorticity across a surface moving with the uid is constant in
time.
Denition 4 (Vortex lines) These are the lines that are everywhere parallel to the
local vorticity , i.e. with t xed they solve
d
ds
x(s) = (x(s), t).
These are the trajectories for the eld for t xed.
Denition 5 (Vortex tube) This is the surface formed by the vortex lines through the
points of a simple closed curve C; see Fig. 12. We can dene the strength of the vortex
tube to be
_
S
ndS
_
C
t
u dx.
Remark 6 This is a good denition because it is independent of the precise cross-
sectional area S, and the precise circuit C around the vortex tube taken (because
0); see Fig. 12. Vorticity is larger where the cross-sectional area is smaller and
vice-versa. Further, for an ideal uid, vortex tubes move with the uid and the strength
of the vortex tube is constant in time as it does so (Helmholtzs theorem; 1858); see
Chorin and Marsden [3, p. 26].
C
t
S
S
S
2
1
0
Fig. 11 Stokes theorem tells us that the circulation around the closed contour C equals the
ux of vorticity through any surface whose perimeter is C. For example here the ux of vorticity
through S
0
, S
1
and S
2
is the same.
C
S
Fig. 12 The strength of the vortex tube is given by the circulation around any curve C that
encircles the tube once.
30 Simon J.A. Malham
9 Bernoullis Theorem
Theorem 6 (Bernoullis Theorem) Suppose we have an ideal homogeneous incompress-
ible stationary ow with a conservative body force f = , where is the potential
function. Then the quantity
H :=
1
2
|u|
2
+
p

+
is constant along streamlines.
Proof We need the following identity that can be found in Appendix A:
1
2

_
|u|
2
_
= u u +u (u).
Since the ow is stationary, Eulers equation of motion for an ideal uid imply
u u =
_
p

_
.
Using the identity above we see that
1
2

_
|u|
2
_
u (u) =
_
p


_
1
2
|u|
2
+
p

+
_
= u (u)
H = u (u),
using the denition for H given in the theorem. Now let x(s) be a streamline that
satises x

(s) = u
_
x(s)
_
. By the fundamental theorem of calculus, for any s
1
and s
2
,
H
_
x(s
2
)
_
H
_
x(s
1
)
_
=
_
s
2
s
1
dH
_
x(s)
_
=
_
s
2
s
1
H x

(s) ds
=
_
s
2
s
1
_
u (u)
_
u
_
x(s)
_
ds
= 0,
where we used that (u a) u 0 for any vector a (since u a is orthogonal to u).
Since s
1
and s
2
are arbitrary we deduce that H does change along streamlines.
Remark 7 Note that H has the units of an energy density. Since is constant here,
we can interpret Bernoullis Theorem as saying that energy density is constant along
streamlines.
Example (Torricelli 1643). Consider the problem of an oil drum full of water that
has a small hole punctured into it near the bottom. The problem is to determine the
velocity of the uid jetting out of the hole at the bottom and how that varies with the
amount of water left in the tankthe setup is shown in Fig 13. We shall assume the
hole has a small cross-sectional area . Suppose that the cross-sectional area of the
drum, and therefore of the free surface (water surface) at z = 0, is A. We naturally
assume A . Since the rate at which the amount of water is dropping inside the
Introductory uid mechanics 31
h
z=0
z=h
P
U
P = air pressure
0
0
Typical streamline
Fig. 13 Torricelli problem: the pressure at the top surface and outside the puncture hole is
atmospheric pressure P
0
. Suppose the height of water above the puncture is h. The goal is to
determine how the velocity of water U out of the puncture hole varies with h.
drum must equal the rate at which water is leaving the drum through the punctured
hole, we have
_

dh
dt
_
A = U
_

dh
dt
_
=
_

A
_
U.
We observe that A , i.e. /A 1, and hence we can deduce
1
U
2
_
dh
dt
_
2
=
_

A
_
2
1.
Since the ow is quasi-stationary, incompressible as its water, and there is conserva-
tive body force due to gravity, we apply Bernoullis Theorem for one of the typical
streamlines shown in Fig. 13. This implies that the quantity H is the same at the free
surface and at the puncture hole outlet, hence
1
2
_
dh
dt
_
2
+
P
0

=
1
2
U
2
+
P
0

gh.
Thus cancelling the P
0
/ terms then we can deduce that
gh =
1
2
U
2

1
2
_
dh
dt
_
2
=
1
2
U
2
_
1
1
U
2
_
dh
dt
_
2
_
=
1
2
U
2
_
1
_

A
_
2
_

1
2
U
2
for /A 1. Thus in the asymptotic limit gh =
1
2
U
2
so that
U =
_
2gh.
32 Simon J.A. Malham
Remark 8 Note the pressure inside the container at the puncture hole level is P
0
+gh.
The dierence between this and the atmospheric pressure P
0
outside, accelerates the
water through the puncture hole.
Example (Channel ow: Froude number). Consider the problem of a steady ow
of water in a channel over a gently underlating bedsee Fig 14. We assume that the
x
P
0
U
H
u
h(x)
y(x)
Fig. 14 Channel ow problem: a steady ow of water, uniform in cross-section, ows over a
gently undulating bed of height y = y(x) as shown. The depth of the ow is given by h = h(x).
Upstream the ow is characterized by ow velocity U and depth H.
ow is shallow and uniform in cross-section. Upstream the ow is characterized by ow
velocity U and depth H. The ow then impinges on a gently undulating bed of height
y = y(x) as shown in Fig 14, where x measures distance downstream. The depth of the
ow is given by h = h(x) whilst the uid velocity at that point is u = u(x), which is
uniform over the depth throughout. Re-iterating slightly, our assumptions are thus,

dy
dx

1 (bed gently undulating)


and

dh
dx

1 (small variation in depth).


The continuity equation (incompressibility here) implies that for all x,
uh = UH.
Then Eulers equations for a steady ow imply Bernoullis theorem which we apply
to the surface streamline, for which the pressure is constant and equal to atmospheric
pressure P
0
, hence we have for all x:
1
2
U
2
+gH =
1
2
u
2
+g(y +h).
Substituting for u = u(x) from the incompressibility condition above, and rearranging,
Bernoullis theorem implies that for all x we have the constraint
y =
U
2
2g
+H h
(UH)
2
2gh
2
.
We can think of this as a parametric equation relating the uid depth h = h(x) to the
undulation height h = h(x) where the parameter x runs from x = far upstream
Introductory uid mechanics 33
to x = +far downstream. We plot this relation, y as a function of h, in Fig 15. Note
that y has a unique global maximum y
0
coinciding with the local maximum and given
by
dy
dh
= 0 h = h
0
=
(UH)
2/3
g
1/3
.
Note that if we set
h
y
h
0
y
0
Fig. 15 Channel ow problem: The ow depth h = h(x) and undulation height y = y(x) are
related as shown, from Bernoullis theorem. Note that y has a maximum value y
0
at height
h
0
= HF
2/3
where F = U/

gH is the Froude number.


F := U/
_
gH
then h
0
= HF
2/3
, where F is known as the Froude number. It is a dimensionless
function of the upstream conditions and represents the ratio of the oncoming uid
speed to the wave (signal) speed in uid depth H.
Note that when y = y(x) attains its maximum value at h
0
, then y = y
0
where
y
0
:= H
_
1 +
1
2
F
2

3
2
F
2/3
_
.
This puts a bound on the height of the bed undulation that is compatible with the
upstream conditions. In Fig 16 we plot the maximum permissible height y
0
the undula-
tion is allowed to attain as a function of the Froude number F. Note that two dierent
values of the Froude number F give the same maximum permissible undulation height
y
0
, one of which is slower and one of which is faster (compared with

gH).
Let us now consider and actual given undulation y = y(x). Suppose that it attains
an actual maximum value y
max
. There are three cases to consider, in turn we shall
consider y
max
< y
0
, the more interesting case, and then y
max
> y
0
. The third case
y
max
= y
0
is an exercise (see the Exercises section at the end of these notes).
In the rst case, y
max
< y
0
, as x varies from x = to x = +, the undulation
height y = y(x) varies but is such that y(x) y
max
. Refer to Fig. 15, which plots
the constraint relationship between y and h resulting from Bernoullis theorem. Since
y(x) y
max
as x varies from to +, the values of (h, y) are restricted to part
of the branches of the graph either side of the global maximum (h
0
, y
0
). In the gure
these parts of the branches are the locale of the shaded sections shown. Note that the
derivative dy/dh = 1/(dh/dy) has the same xed (and opposite) sign in each of the
branches. In the branch for which h is small, dy/dh > 0, while the branch for which
34 Simon J.A. Malham
F
y
0 / H
F=1 F>1 F<1 0
Fig. 16 Channel ow problem: Two dierent values of the Froude number F give the same
maximum permissible undulation height y
0
. Note we actually plot the normalized maximum
possible height y
0
/H on the ordinate axis.
h is larger, dy/dh < 0. Indeed note the by dierentiating the constraint condition, we
have
dy
dh
=
_
1
(UH)
2
gh
3
_
.
Using the incompressibility condition to substitute for UH we see that this is equivalent
to
dy
dh
=
_
1
u
2
gh
_
.
We can think of u/

gh as a local Froude number if we like. In any case, note that since


we are in one branch or the other, and in either case the sign of dy/dh is xed, this
means that using the expression for dy/dh we just derived, for any ow realization the
sign of 1 u
2
/gh is also xed. When x = this quantity has the value 1 U
2
/gH.
Hence the sign of 1 U
2
/gH determines the sign of 1 u
2
/gh. Hence if F < 1 then
U
2
/gH = F
2
< 1 and therefore for all x we must have u
2
/gh < 1. And we also deduce
in this case that we must be on the branch for which h is relatively large as dy/dh is
negative. The ow is said to be subcritical throughout and indeed we see that
dh
dy
=
_
dy
dh
_
1
=
_
1
u
2
gh
_
1
< 1
d
dy
(h +y) < 0.
Hence in this case, as the bed height y increases, the uid depth h decreases and vice-
versa. On the otherhand if F > 1 then U
2
/gH > 1 and thus u
2
/gh > 1. We must be
on the branch for which h is relatively small as dy/dh is positive. The ow is said to
be supercritical throughout and we have
dh
dy
=
_
1
u
2
gh
_
1
> 0
d
dy
(h +y) > 1.
Hence in this case, as the bed height y increases, the uid depth h increases and vice-
versa. Both cases, F < 1 and F > 1, are illustrated by a typical scenario in Fig. 17.
In the second case, y
max
> y
0
, the undulation height is larger than the maximum
permissibe height y
0
compatible with the upstream conditions. Under the conditions
we assumed, there is no ow realized here. In a real situation we may imagine a ow
impinging on a large barrier with height y
max
> y
0
, and the result would be some
sort of reection of the ow occurs to change the upstream conditions in an attempt
to make them compatible with the obstacle. (Our steady ow assumption obviously
breaks down here.)
Introductory uid mechanics 35
U U
F<1 F>1
Fig. 17 Channel ow problem: for the case y
max
< y
0
, when F < 1, as the bed height y
increases, the uid depth h decreases and vice-versa. Hence we see a depression in the uid
surface above a bump in the bed. On the other hand, when F > 1, as the bed height y increases,
the uid depth h increases and vice-versa. Hence we see an elevation in the uid surface above
a bump in the bed.
10 Irrotational/potential ow
Many ows have extensive regions where the vorticity is zero; some have zero vorticity
everywhere. We would call these, respectively, irrotational regions of the ow and
irrotational ows. In such regions
= u = 0.
Hence the eld u is conservative and there exists a scalar function such that
u = .
The function is known as the ow potential. In turn this implies that
K =
_
C
u dx = 0
for all simple closed curves C in the region (the reverse implication is also true).
If the uid is also incompressible, then is harmonic since u = 0 implies
= 0.
Hence for such situations, we in essense need to solve Laplaces equation = 0 subject
to certain boundary conditions. For example for an ideal ow, u n = n = /n
is given on the boundary, and this would consitute a Neumann problem for Laplaces
equation.
Example (linear two-dimensional ow) Consider the ow eld u = (kx, ky) where
k is a constant. It is irrotational. Hence there exists a ow potential =
1
2
k(x
2
y
2
).
Since u = 0 as well, we have = 0. Further, since this ow is two-dimensional,
there also exists a streamfunction = kxy.
Example (line vortex) Consider the ow eld (u
r
, u

, u
z
) = (0, k/r, 0) where k > 0
is a constant. This is the idealization of a thin vortex tube. Direct computation shows
that u = 0 everywhere except at r = 0, where u is innite. For r > 0, there
exists a ow potential = k. For any closed circuit C in this region, we have
K =
_
C
u dx = 2k N
36 Simon J.A. Malham
where N is the number of times the closed curve C winds round the origin r = 0. The
circulation K will be zero for all circuits reducible continuously to a point without
breaking the vortex.
Example (DAlemberts paradox) Consider a uniform ow into which we place an
obstacle. We would naturally expect that the obstacle represents an obstruction to the
uid ow and that the ow would exert a force on the obstacle, which if strong enough,
might dislodge it and subsequently carry it downstream. However for an ideal ow, as
we are just about to prove, this is not the case. There is no net force exerted on an
obstacle placed in the midst of a uniform ow.
We thus consider a uniform ideal ow into which is placed a sphere, radius a.
The set up is shown in Fig. 18. We assume that the ow around the sphere is steady,
incompressible and irrotational. Suppose further that the ow is axisymmetric. By
this we mean the following. Using spherical polar coordinates to represent the ow
with the south-north pole axis passing through the centre of the sphere and aligned
with the uniform ow U at innity; see Fig. 18. Then the ow is axisymmetric if it is
independent of the azimuthal angle of the spherical coordinates (r, , ). Further we
also assume no swirl so that u

= 0. Since the ow is incompressible and irrotational, it


U
U
U
U
r

Fig. 18 Consider an ideal steady, incompressible, irrotational and axisymmetric ow past a


sphere as shown. The net force exerted on the sphere (obstacle) in the ow is zero. This is
DAlemberts paradox.
is a potential ow. Hence we seek a potential function such that = 0. In spherical
polar coordinates this is equivalent to
1
r
2
_

r
_
r
2

r
_
+
1
sin

_
sin

__
= 0.
The general solution to Laplaces equation is well known, and in the case of axisym-
metry the general solution is given by
(r, ) =

n=0
_
A
n
r
n
+
B
n
r
n+1
_
P
n
(cos )
where P
n
are the Legendre polynomials; with P
1
(x) = x. The coecients A
n
and B
n
are constants, most of which, as we shall see presently, are zero. For our problem we
Introductory uid mechanics 37
have two sets of boundary data. First, that as r in any direction, the ow eld is
uniform and given by u = (0, 0, U) (expressed in Cartesian coordinates with the z-axis
aligned along the south-north pole) so that as r
Ur cos .
Second, on the sphere r = a itself we have a no normal ow condition

r
= 0.
Using the rst boundary condition for r we see that all the A
n
must be zero
except A
1
= U. Using the second boundary condition on r = a we see that all the
B
n
must be zero except for B
1
=
1
2
Ua
3
. Hence the potential for this ow around the
sphere is
= U(r +a
3
/2r
2
) cos .
In spherical polar coordinates, the velocity eld u = is given by
u = (u
r
, u

) =
_
U(1 a
3
/r
3
) cos , U(1 +a
3
/2r
3
) sin
_
.
Since the ow is ideal and steady as well, Bernoullis theorem applies and so along a
typical streamline
1
2
|u|
2
+ P/ is constant. Indeed since the conditions at innity are
uniform so that the pressure P

and velocity eld U are the same everywhere there,


this means that for any streamline and in fact everywhere for r a we have
1
2
|u|
2
+P/ =
1
2
U
2
+P

/.
Rearranging this equation and using our expression for the velocity eld above we have
P P

=
1
2
U
2
_
1 (1 a
3
/r
3
)
2
cos
2
(1 +a
3
/2r
3
)
2
sin
2

_
.
On the sphere r = a we see that
P P

=
1
2
U
2
_
1
9
4
sin
2

_
.
Note that on the sphere, the pressure is symmetric about = 0, /2, , 3/2. Hence
the uid exerts no net force on the sphere! (There is no drag or lift.) This result, in
principle, applies to any shape of obstacle in such a ow. In reality of course this cannot
be the case, the presence of viscosity remedies this paradox (and crucially generates
vorticity).
11 Dynamical similarity and Reynolds number
Our goal in this section is to demonstrate an important scaling property of the Navier
Stokes equations for a homogeneous incompressible uid without body force:
u
t
+u u = u
1

p,
u = 0.
38 Simon J.A. Malham
Note that two physical properties inherent to the uid modelled are immediately ap-
parent, the mass density , which is constant throughout the ow, and the kinematic
viscosity . Suppose we consider such a ow which is characterized by a typical length
scale L and velocity U. For example we might imagine a ow past an obstacle such a
sphere whose diameter is characterized by L and the impinging/undisturbed far-eld
ow is uniform and given by U. These two scales naturally determine a typically time
scale T = L/U. Using these scales we can introduce the dimensionless variables
x

=
x
L
, u

=
u
U
and t

=
t
T
.
Directly substituting for u = Uu

and using the chain rule to replace t by t

and x by
x

in the NavierStokes equations, we obtain:


U
T
u

+
U
2
L
u


x
u

=
U
L
2

x
u

1
L

x
p.
The incompressibility condition becomes
x
u

= 0. Using that T = L/U and dividing


through by U
2
/L we get
u

+u


x
u

=

UL

x
u

1
U
2

x
p.
If we set p

= p/U
2
and then drop the primes, we get
u
t
+u u =
1
Re
u p,
which is the representation for the NavierStokes equations in dimensionless variables.
The dimensionless number
Re :=
UL

is the Reynolds number. Its practical signicance is as follows. Suppose we want to


design a jet plane (or perhaps just a wing). It might have a characteristic scale L
1
and
typically cruise at speeds U
1
with surrounding air having viscosity
1
. Rather than
build the plane to test its airow properties it would be cheaper to build a scale model
of the aircraftwith exactly the same shape/geometry but smaller, with characteristic
scale L
2
. Then we could test the airow properties in a wind tunnel for example, by
using a driving impinging wind of characteristic velocity U
2
and air of viscosity
2
so
that
U
1
L
1

1
=
U
2
L
2

2
.
The Reynolds number in the two scenarios are the same and the dimensionless Navier
Stokes equations for the two ows identical. Hence the shape of the ows in the two
scenarios will be the same. We could also for example, replace the wind tunnel by a
water tunnel: the viscosity of air is
1
= 0.15 cm
2
/s and of water
2
= 0.0114 cm
2
/s,
i.e.
1
/
2
13. Hence for the same geometry and characteristic scale L
1
= L
2
, if we
choose U
1
= 13 U
2
, the Reynolds numbers for the two ows will be the same. Such
ows, with the same geometry and the same Reynolds number are said to be similar.
Remark 9 Some typical Reynolds are as follows: aircraft: 10
8
to 10
9
; cricket ball: 10
5
;
blue whale: 10
8
; cruise ship: 10
9
; canine artery: 10
3
; nematode: 0.6; capilliaries: 10
3
.
Introductory uid mechanics 39
12 Exercises
Exercise (streamlines: Cartesian coordinates) Sketch streamlines for the following
steady ow elds: (a) (u, v) = (U + y, 0); (b) (u, v) = (kx, ky); (c) (u, v) =
(y, x); (d) (u, v, w) = (x, y, 2z); where U, , k, and are all constants.
Exercise (streamlines: plane/cylindrical polar coordinates) Sketch streamlines for the
steady ow elds: (a) (u
r
, u

) = (0, r); (b) (u


r
, u

) = (0, k/r); (c) (u


r
, u

, u
z
) =
(t) (xy, x+y, 0)show that the streamlines are exponential spirals (Hint: convert
to polar coordinates rst). Here and k are constants, and = (t) is an arbitrary
function of t. We use (r, ) as plane polar coordinates, and (r, , z) as cylindrical polar
coordinates. Note that in these coordinates the equations for trajectories are
dr
dt
= u
r
, r
d
dt
= u

, and
dz
dt
= u
z
.
Exercise (trajectories and streamlines: two dimensions) For a given velocity ow eld
u = u(x, t) prescribed at position x and time t, the particle trajectories are given by
the solutions to the system of ordinary dierential equations
dx
dt
= u(x(t), t).
Streamlines are given by the solutions to the system of ordinary dierential equations
(where t is xed)
dx
ds
= u
_
x(s), t
_
.
(a) Explain what particle trajectories and streamlines are, and their dierence.
(b) For the two-dimensional ow in Cartesian coordinates,
(u, v) =
_
u
0
, v
0
cos(kx t)
_
,
where u
0
, v
0
, k and are constants, nd the general equation for a streamline.
Show that the streamline passing through (x, y) = (0, 0) at t = 0 is
y =
v
0
ku
0
sin(kx).
Find the equation for the path of the particle which is at (x, y) = (0, 0) at t = 0.
Comment briey on the contrast between the above streamline and particle path
in the two separate limiting cases: rst 0; second k 0.
Exercise (trajectories and streamlines: three dimensions) Find the trajectories and
streamlines when u = (xe
2tz
, ye
2tz
, 2e
2tz
). What is the track of the particle passing
through (1, 1, 0) at time t = 0?
Exercise (channel ow) Consider the two-dimensional channel ow (with U a given
constant)
u =
_
0, U
_
1
x
2
a
2
_
, 0
_
,
between the two walls x = a. Show that there is a stream function and nd it. (Hint:
a stream function exists for a velocity eld u = (u, v, w) when u = 0 and we have
40 Simon J.A. Malham
an additional symmetry. Here the additional symmetry is uniformity with respect to
z. You thus need to verify that if
u =

y
and v =

x
,
then u = 0 and then solve this system of equations to nd .)
Show that approximately 91% of the volume ux across y = y
0
for some constant
y
0
ows through the central part of the channel |x|
3
4
a.
Exercise (ow inside and around a disc) Calculate the stream function for the ow
eld
u =
_
U cos (1 a
2
/r
2
), U sin (1 +a
2
/r
2
) /2r
_
in plane polar coordinates, where U, a, are constants.
Exercise (steady oscillating channel ow) An incompressible uid is in steady two-
dimensional ow in the channel
< x < , /2 < y < /2,
with velocity
u = (1 +xsin y, cos y).
Find the equation of the streamlines and sketch them. Show that the ow has stagnation
points at (1, /2) and (1, /2).
Exercise (Flow in an innite pipe: Poiseuille ow) Consider an innite pipe with
circular cross-section of radius a, whose centre line is aligned along the z-axis. Assume
no-slip boundary conditions at r = a, for all z, i.e. on the inside surface of the cylinder.
Using cylindrical polar coordinates, look for a solution to the uid ow in the pipe
of the following form. Assume there is no radial ow, u
r
= 0, and no swirl, u

= 0.
Further assume there is a constant pressure gradient down the pipe, i.e. that p = Cz
for some constant C. Lastly, suppose that the ow down the pipe, i.e. the velocity
component u
z
, has the form u
z
= u
z
(r) (it is a function of r only).
(a) Using the NavierStokes equations, show that
C = u
z
=
_
1
r

r
_
r
u
z
r
__
.
(b) Integrating the equation above yields
u
z
=
C
4
r
2
+Alog r +B,
where A and B are constants. We naturally require that the solution be bounded.
Explain why this implies A = 0. Now use the no-slip boundary condition to deter-
mine B. Hence show that
u
z
=
C
4
(a
2
r
2
).
(c) Show that the mass-ow rate across any cross section of the pipe is given by
_
u
z
dS =
2
Ca
4
/8.
This is known as the fourth power law.
Introductory uid mechanics 41
z
a
Fig. 19 Poiseuille ow: a viscous uid ows along an innite horizontal pipe with circular
cross-section of radius a, whose centre line is aligned along the z-axis. A constant pressure
gradient is assumed, as well as axisymmetry, no radial ow and no swirl.
(From Chorin and Marsden, pp. 45-6.)
Exercise (Couette ow)Let be the region between two concentric cylinders of radii
R
1
and R
2
, where R
1
< R
2
. Suppose the velocity eld in cylindrical coordinates
u = (u
r
, u

, u
z
) of the uid ow inside , is given by
u
r
= 0, u
z
= 0, and u

=
A
r
+Br,
where
A =
R
2
1
R
2
2
(
2

1
)
R
2
2
R
2
1
and B =
R
2
1

1
R
2
2

2
R
2
2
R
2
1
.
This is known as a Couette owsee Fig. 20. Show that the:
(a) velocity eld u = (u
r
, u

, u
z
) is a stationary solution of Eulers equations of motion
for an ideal uid with density 1 (hint: you need to nd a pressure eld p that
is consistent with the velocity eld given);
(b) angular velocity of the ow (i.e. the quantity u

/r) is
1
on the cylinder r = R
1
and
2
on the cylinder r = R
2
.
(c) the vorticity eld = u = (0, 0, 2B).
two walls of the cylinders
fluid lies between the
Region of flow is :
R
1
R
2
Fig. 20 Couette ow between two concentric cylinders of radii R
1
< R
2
.
Exercise (hurricane) We devise a simple model for a hurricane.
42 Simon J.A. Malham
(a) Using the Euler equations for an ideal incompressible ow in cylindrical coordinates
(see the bath or sink drain problem in the main text) show that at position (r, , z),
for a ow which is independent of with u
r
= u
z
= 0, we have
u
2

r
=
1

0
p
r
,
0 =
1

0
p
z
+g,
where p = p(r, z) is the pressure and g is the acceleration due to gravity (assume
this to be the body force per unit mass). Verify that any such ow is indeed in-
compressible.
(b) In a simple model for a hurricane the air is taken to have uniform constant density

0
and each uid particle traverses a horizontal circle whose centre is on the xed
vertical z-axis. The (angular) speed u

at a distance r from the axis is


u

=
_
r, for 0 r a,

a
3/2
r
1/2
, for r > a,
where and a are known constants.
(i) Now consider the ow given above in the inner region 0 r a. Using the
equations in part (a) above, show that the pressure in this region is given by
p = c
0
+
1
2

2
r
2
g
0
z,
where c
0
is a constant. A free surface of the uid is one for which the pressure
is constant. Show that the shape of a free surface for 0 r a is a paraboloid
of revolution, i.e. it has the form
z = Ar
2
+B,
for some constants A and B. Specify the exact form of A and B.
(ii) Now consider the ow given above in the outer region r > a. Again using the
equations in part (a) above, and that the pressure must be continuous at r = a,
show that the pressure in this region is given by
p = c
0


0
r

2
a
3
g
0
z +
3
2

2
a
2
,
where c
0
is the same constant (reference pressure) as that in part (i) above.
Exercise (Elliptic pipe ow) Show that for a ow of a viscous uid through a pipe of
constant cross-section under pressure gradient G, the speed u satises

2
u
x
2
+

2
u
y
2
=
G

,
where x and y are coordinates in a plane of cross-section. State the boundary conditions
for this elliptic partial dierential problem.
Show for an elliptic pipe of semi-axes a, b, that u = A + Bx
2
+ Cy
2
for suitable
choices of A, B and C. Verify that the ux of uid through the pipe is a
3
b
3
G/4(a
2
+
b
2
). Deduce that for a given elliptical cross-sectional area, the optimal choice of a and
b to maximize the discharge rate is a = b.
Introductory uid mechanics 43
Exercise (Wind blowing across a lake) Wind blowing across the surface of a lake of
uniform depth d exerts a constant and uniform tangential stress S. The water is initially
at rest. Find the water velocity at the surface as a function of time for t d
2
. (Hint:
solve for the vorticity using the vorticity equation for a very deep lake.)
Suppose now that the wind has been blowing for a suciently long time to es-
tablish a steady state. Assuming that the water velocity can be taken to be almost
uni-directional and that the horizontal dimensions of the lake are large compared with
d, show that the water velocity at the surface is Sd/4. (Hint: A pressure gradient
would be needed to ensure no net ux (why?) in the steady state, and this pressure
gradient leads to a small rise in the surface elevation of the lake in the direction of the
wind.)
Exercise (Venturi tube) Consider the Venturi tube shown in Fig. 21. Assume that the
ideal uid ow through the construction is homogeneous, incompressible and steady.
The ow in the wider section of cross-sectional area A
1
, has velocity u
1
and pressure
p
1
, while that in the narrower section of cross-sectional area A
2
, has velocity u
2
and
pressure p
2
. Separately within the uniform wide and narrow sections, we assume the
velocity and pressure are uniform themselves.
(a) Why does the relation A
1
u
1
= A
2
u
2
hold? Why is the ow faster in the narrower
region of the tube compared to the wider region of the tube?
(b) Use Bernoullis theorem to show that
1
2
u
2
1
+
p
1

0
=
1
2
u
2
2
+
p
2

0
,
where
0
is the constant uniform density of the uid.
(c) Using the results in parts (a) and (b), compare the pressure in the narrow and wide
regions of the tube.
(d) Give a practical application where the principles of the Venturi tube is used or
might be useful.
streamline
A , u , p
1 1 1
A , u , p
1 1 1
A , u , p
2 2 2
Fig. 21 Venturi tube: the ow in the wider section of cross-sectional area A
1
has velocity
u
1
and pressure p
1
, while that in the narrower section of cross-sectional area A
2
has velocity
u
2
and pressure p
2
. Separately within the uniform wide and narrow sections, we assume the
velocity and pressure are uniform themselves.
Exercise (Clepsydra or water clock) A clepsydra has the form of a surface of revolution
containing water and the level of the free surface of the water falls at a constant rate,
as the water ows out through a small hole in the base. The basic setup is shown in
Fig. 22.
44 Simon J.A. Malham
(a) Apply Bernoullis theorem to one of the typical streamlines shown in Fig. 22 to
show that
1
2
_
dz
dt
_
2
=
1
2
U
2
gz
where z is the height of the free surface above the small hole in the base, U is the
velocity of the water coming out of the small hole and g is the acceleration due to
gravity.
(b) Assuming that the constant rate at which the level surface is falling is very slow,
explain why we can deduce that
U
_
2gz.
(c) If S is the cross-sectional area of the hole in the bottom, and A is the cross-sectional
area of the free surface, explain why we must have
A
dz
dt
= S U.
(d) Combine the results from (b) and (c) above, to show that the shape of the container
that guarantees that the free surface of the water drops at a constant rate must
have the form z = C r
4
in cylindrical polars, where C is a constant.
r
z
P
0
= air pressure
U
typical streamline
r
z
free surface
P
0
Fig. 22 Clepsydra (water clock).
Exercise (coee in a mug) A coee mug in the form of a right circular cylinder (diam-
eter 2a, height h), closed at one end, is initially lled to a depth d >
1
2
h with static
inviscid coee. Suppose the coee is then made to rotate inside the mugsee Fig. 23.
(a) Using the Euler equations for an ideal incompressible homogeneous ow in cylin-
drical coordinates show that at position (r, , z), for a ow which is independent of
with u
r
= u
z
= 0, the Euler equations reduce to
u
2

r
=
1

p
r
,
0 =
1

p
z
+g,
where p = p(r, z) is the pressure, is the constant uniform uid density and g is
the acceleration due to gravity (assume this to be the body force per unit mass).
Verify that any such ow is indeed incompressible.
Introductory uid mechanics 45
(b) Assume that the coee in the mug is rotating as a solid body with constant angular
velocity so that the velocity component u

at a distance r from the axis of


symmetry for 0 r a is
u

= r.
Use the equations in part (a), to show that the pressure in this region is given by
p =
1
2

2
r
2
gz +C,
where C is an arbitrary constant. At the free surface between the coee and air,
the pressure is constant and equal to the atmospheric pressure P
0
. Use this to show
that the shape of the free surface has the form
z =

2
2g
r
2
+
C P
0
g
.
(c) Note the we are free to choose C = P
0
in the equation of the free surface so that
it is described by z =
2
r
2
/2g. This is equivalent to choosing the origin of our
cylindrical coordinates to be the centre of the dip in the free surface. Suppose that
this origin is a distance z
0
from the bottom of the mug.
(i) Explain why the total volume of coee is a
2
d. Then by using incompressibility,
explain why the following constraint must be satised:
a
2
z
0
+
_
a
0

2
r
2
2g
2r dr = a
2
d.
(ii) By computing the integral in the constraint in part (i), show that some coee
will be spilled out of the mug if
2
> 4g(h d)/a
2
. Explain briey why this
formula does not apply when the mug is initially less than half full.
O
d
2a
h
P
0
free surface
coffee
Fig. 23 Coee mug: we consider a mug of coee of diameter 2a and height h, which is initially
lled with coee to a depth d. The coee is then made to rotate about the axis of symmetry
of the mug. The free surface between the coee and the air takes up the characteristic shape
shown, dipping down towards the middle (axis of symmetry). The goal is to specify the shape
of the free surface.
Exercise (Channel ow: Froude number) Recall the scenario of the steady channel
ow over a gently undulating bed given in Section 9. Consider the case when the
maximum permissible height y
0
compatible with the upstream conditions, and the
46 Simon J.A. Malham
actual maximum height y
max
of the undulation are exactly equal, i.e. y
max
= y
0
. Show
that the ow becomes locally critical immediately above y
max
and, by a local expansion
about that position, show that there are subcritical and supercritical ows downstream
consistent with the continuity and Bernoulli equations (friction in a real ow leads to
the latter being preferred).
Exercise (Bernoullis Theorem for irrotational unsteady ow) Consider Eulers equa-
tions of motion for an ideal homogeneous incompressible uid, with u = u(x, t) de-
noting the uid velocity at position x and time t, the uniform constant density,
p = p(x, t) the pressure, and f denoting the body force per unit mass. Suppose that
the ow is unsteady, but irrotational, i.e. we know that
u = 0
throughout the ow. This means that we know there exists a scalar potential function
= (x, t) such that u = . Also suppose that the body force is conservative so
that f = for some potential function = (x, t).
(a) Using the identity
u u =
1
2
(|u|
2
) u (u),
show from Eulers equations of motion that the Bernoulli quantity
H :=

t
+
1
2
|u|
2
+
p

+
satises H = 0 throughout the ow.
(b) From part (a) above we can deduce that H can only be a function of t throughout
the ow, say H = f(t) for some function f. By setting
V :=
_
t
0
f() d
throughout the ow show that the Bernoulli quantity
H :=
V
t
+
1
2
|u|
2
+
p

+
is constant throughout the ow.
Exercise (rigid body rotation) An ideal uid of constant uniform density
0
is contained
within a xed right-circular cylinder (with symmetry axis the z-axis). The uid moves
under the inuence of a body force eld f = (x + y, x + y, 0) per unit mass,
where , , and are independent of the space coordinates. Use Eulers equations of
motion to show that a rigid body rotation of the uid about the z-axis, with angular
velocity (t) given by =
1
2
( ) is a possible solution of the equation and boundary
conditions. Show that the pressure is given by
p = p
0
+
1
2

0
_
(
2
+)x
2
+ ( +)xy + (
2
+)y
2
_
,
where p
0
is the pressure at the origin.
Introductory uid mechanics 47
Exercise (vorticity and streamlines) An inviscid incompressible uid of uniform density
is in steady two-dimensional horizontal motion. Show that the Euler equations are
equivalent to
H
x
= v and
H
y
= u,
where H = p/ +
1
2
(u
2
+v
2
), where p is the dynamical pressure, (u, v) is the velocity
eld and is the vorticity. Deduce that is constant along streamlines and that this
is in accord with Kelvins theorem.
Exercise (vorticity, streamlines with gravity) An incompressible inviscid uid, under the
inuence of gravity, has the velocity eld u = (2y, x, 0) with the z-axis vertically
upwards; and is constant. Also the density is constant. Verify that u satises
the governing equations and nd the pressure p. Show that the Bernoulli function
H = p/ +
1
2
|u|
2
+ is constant on streamlines and vortex lines, where is the
gravitational potential.
13 Notes
13.1 Streaklines
A streakline is the locus of all the uid elements which at some time have past through
a particular point, say (x
0
, y
0
, z
0
). We can obtain the equation for a streakline through
(x
0
, y
0
, z
0
) by solving the equations
d
dt
x(t) = u(x(t), t),
assuming that at t = t
0
we have
_
x(t
0
), y(t
0
), z(t
0
)
_
= (x
0
, y
0
, z
0
). Eliminating t
0
be-
tween the equations generates the streakline corresponding to (x
0
, y
0
, z
0
). For example,
ink dye injected at the point (x
0
, y
0
, z
0
) in the ow will trace out a streakline.
13.2 Ideal uid ow and conservation of energy
We show that an ideal uid that conserves energy is necessarily incompressible. We
have derived two conservation laws thusfar, rst, conservation of mass,

t
+ (u) = 0,
and second, balance of momentum,

Du
Dt
= p + f.
If we are in three dimensional space so d = 3, we have four equations, but ve
unknownsnamely u, p and . We cannot specify the uid motion completely without
specifying one more condition.
Denition 6 (Kinetic energy) The kinetic energy of the uid in the region D is
E :=
1
2
_

|u|
2
dV.
48 Simon J.A. Malham
The rate of change of the kinetic energy, using the transport theorem, is given by
dE
dt
=
d
dt
_
1
2
_

t
|u|
2
dV
_
=
1
2
_

D|u|
2
Dt
dV
=
1
2
_

D
Dt
(u u) dV
=
_

t
u
Du
Dt
dV
=
_

t
u
_

Du
Dt
_
dV.
Here we assume that all the energy is kinetic. The principal of conservation of energy
states (from Chorin and Marsden, page 13):
the rate of change of kinetic energy in a portion of uid equals the rate at which
the pressure and body forces do work.
In other words we have
dE
dt
=
_

t
p u ndS +
_

t
u f dV.
We compare this with our expression above for the rate of change of the kinetic energy.
Equating the two expressions, using Eulers equation of motion, and noticing that the
body force term immediately cancels, we get
_

t
p u ndS =
_

t
u p dV

t
(up) dV =
_

t
u p dV

t
u p + ( u) p dV =
_

t
u p dV

t
( u) p dV = 0.
Since and therefore
t
is arbitrary we see that the assumption that all the energy
is kinetic implies
u = 0.
Hence our third conservation law, conservation of energy has lead to the equation of
state, u = 0, i.e. that an ideal uid is incompressible.
Hence the Euler equations for a homogeneous incompressible ow in D are
u
t
+u u =
1

p +f,
u = 0,
together with the boundary condition on D which is u n = 0. Note, as we did for
the NavierStokes equations, since is constant, it is convenient to re-label p/ to be
p, thus removing from the equations above completely.
Introductory uid mechanics 49
13.3 Isentropic ows
A compressible ow is isentropic if there is a function , called the enthalpy, such that
=
1

p.
The Euler equations for an isentropic ow are thus
u
t
+u u = +f

t
+ (u) = 0,
in D, and on D, u n = 0 (or matching normal velocities if the boundary is moving).
For compressible ideal gas ow, the pressure is often proportional to

, for some
constant 1, i.e.
p = C

,
for some constant C. This is a special case of an isentropic ow, and is an example of
an equation of state. In fact we can actually compute
=
_

p

(z)
z
dz =
C

1
,
and the internal energy (see Chorin and Marsden, pages 14 and 15)
= (p/) =
C

1
.
A Multivariable calculus identities
We provide here some useful multivariable calculus identities. Here and are generic scalars,
and u and v are generic vectors.
1. u = det

i j k
/x /y /z
u v w

w
y

v
z
u
z

w
x
v
x

u
y

.
2. () =
2
=

2

x
2
+

2

y
2
+

2

z
2
.
3. () 0.
4. (u) 0.
5. (u) = ( u)
2
u.
6. () = +.
7. (u v) = (u )v + (v )u +u (v) +v (u).
8. (u) = ( u) +u .
9. (u v) = v (u) u (v).
10. (u) = u + u.
11. (u v) = u( v) v ( u) + (v )u (u )v.
50 Simon J.A. Malham
B NavierStokes equations in cylindrical polar coordinates
The incompressible NavierStokes equations in cylindrical polar coordinates (r, , z) with the
velocity eld u = (u
r
, u

, u
z
) are
u
r
t
+ (u )u
r

u
2

r
=
1

p
r
+

u
r

u
r
r
2

2
r
2
u

+f
r
,
u

t
+ (u )u

+
u
r
u

r
=
1
r
p

+
2
r
2
u
r

r
2

+f

,
u
z
t
+ (u )u
z
=
1

p
z
+u
z
+f
z
,
where p = p(r, , z, t) is the pressure, is the mass density and f = (f
r
, f

, f
z
) is the body
force per unit mass. Here we also have
u = u
r

r
+
u

+u
z

z
and
=
1
r

r

r

+
1
r
2

2
+

2
z
2
Further the gradient operator and the divergence of a vector eld u are given in cylindrical
coordinates, respectively, by
=


r
,
1
r

,

z

and
u =
1
r

r
(ru
r
) +
1
r
u

+
u
z
z
.
Lastly in cylindrical coordinates u is given by
u =

1
r
u
z

z
u
r
z

u
z
r
1
r

r
(ru

)
1
r
u
r

.
C NavierStokes equations in spherical polar coordinates
The incompressible NavierStokes equations in spherical polar coordinates (r, , ) with the
velocity eld u = (u
r
, u

, u

) are (note is the angle to the south-north pole axis and is


the azimuthal angle)
u
r
t
+ (u )u
r

u
2

u
2

r
=
1

p
r
+

u
r
2
u
r
r
2

2
r
2
sin

(u

sin )
2
r
2
sin
u

+f
r
,
u

t
+ (u )u

+
u
r
u

u
2

cos
r sin
=
1
r
p

+
2
r
2
u
r

r
2
sin
2

2
cos
r
2
sin
2

+f

,
u

t
+ (u )u

+
u
r
u

r
+
u

cos
r sin
=
1
r sin
p

+
2
r
2
sin
u
r

+
2 cos
r
2
sin
2

r
2
sin
2

+f
z
,
Introductory uid mechanics 51
where p = p(r, , , t) is the pressure, is the mass density and f = (f
r
, f

, f

) is the body
force per unit mass. Here we also have
u = u
r

r
+
u

+
u

r sin

and
=
1
r
2

r
2

r

+
1
r
2
sin

sin

+
1
r
2
sin
2

2
.
Further the gradient operator and the divergence of a vector eld u are given in spherical
coordinates, respectively, by
=


r
,
1
r

,
1
r sin

and
u =
1
r
2

r
(r
2
u
r
) +
1
r sin

(sin u

) +
1
r sin
u

.
Lastly in spherical coordinates u is given by
u =

1
r sin

(sin u

)
u

1
r sin
u
r


1
r

r
(ru

)
1
r

r
(ru

)
1
r
u
r

.
Acknowledgements These lecture notes have to a large extent grown out of a merging of,
lectures on Ideal Fluid Mechanics given by Dr. Frank Berkshire [2] in the Spring of 1989,
lectures on Viscous Fluid Mechanics given by Prof. Trevor Stuart [17] in the Autumn of 1989
(both at Imperial College) and the style and content of the excellent text by Chorin and
Marsden [3]. They have also benetted from lecture notes by Prof. Frank Leppington [11] on
Electromagnetism. SJAM would also like to thank Prof. Andrew Lacey for his suggestions and
input. Lastly, thanks to all the students who pointed out typos to me along the way!
References
1. Batchelor, G.K. 1967 An introduction to uid mechanics, CUP.
2. Berkshire, F. 1989 Lecture notes on ideal uid dynamics, Imperial College Mathematics
Department.
3. Chorin, A.J. and Marsden, J.E. 1990 A mathematical introducton to uid mechanics,
Third edition, SpringerVerlag, New York.
4. Doering, C.R. and Gibbon, J.D. 1995 Applied analysis of the NavierStokes equations,
Cambridge Texts in Applied Mathematics, Cambridge University Press.
5. Evans, L.C. 1998 Partial dierential equations, Graduate Studies in Mathematics, Volume
19, American Mathematical Society.
6. Fulton, W. and Harris, J. 2004 Representation theory: A rst course, Graduate Texts in
Mathematics 129, Springer.
7. Gurtin, M.E. 1981 An introduction to continuum mechanics, Mathematics in Science and
Engineering, Volume 158, Academic Press.
8. Keener, J.P. 2000 Principles of applied mathematics: transformation and approximation,
Perseus Books.
9. Krantz, S.G. 1999 How to teach mathematics, Second Edition, American Mathematical
Society.
10. Lamb, H. 1932 Hydrodynamics, 6th Edition, CUP.
11. Leppington, F. 1989 Electromagnetism, Imperial College Mathematics Department.
12. McCallum, W.G. et. al. 1997 Multivariable calculus, Wiley.
13. Majda, A.J. and Bertozzi, A.L. 2002 Vorticity and incompressible ow, Cambridge Texts
in Applied Mathematics, Cambridge University Press.
14. Marsden, J.E. and Ratiu, T.S. 1999 Introduction to mechanics and symmetry, Second
edition, Springer.
52 Simon J.A. Malham
15. Meyer, C.D. 2000 Matrix analysis and applied linear algebra, SIAM.
16. Saman, P.G. 1992 Vortex dynamics, Cambridge Monographs in Mechanics and Applied
Mathematics, Cambridge University Press.
17. Stuart, J.T. 1989 Lecture notes on Viscous Fluid Mechanics, Imperial College Mathematics
Department.

You might also like