Fluid Formulas
Fluid Formulas
Jens Eggers
Preliminaries
Course information
• Lecturer: Prof. Jens Eggers, Room SM2.3
• Timetable: Weeks 1-12, Tuesday 1.00 (room 1.18 Queen’s building), Thursday 1.00
(room 1.18 Queen’s bulding) and Friday 10.00 (Physics 3.21). Course made up of
32 lectures.
• Prerequisites: Mechanics 1, APDE2 and Calc2. Need ideas from vector calculus,
complex function theory, separation solutions and PDE’s.
• Homework: Questions from 10 worksheets will be set and marked during the course.
Homework sheets will be handed out each Friday, starting in the first week. Solutions
to be returned the following Friday in the box marked “Fluids 3”.
• Web: Standard unit description includes detailed course information. Lecture notes
will be posted on Blackboard, along with homework and solution sheets.
• Lectures: There is no need to take notes during the lecture, as all material relevant
for the exam will be put on the web. The main purpose of the lecture is the live
development of the material, and a chance for you to ask questions!
Recommended texts
1. A.R. Paterson, A First Course in Fluid Dynamics, Cambridge University Press.
(The recommended text to complement this course - costs ≈ £50 from Amazon;
there are 6 copies in Queen’s building Library and 3 copies in the Physics Library)
Films
There is a very good series of educational films on Fluid Mechanics available on YouTube,
produced by the National Committee for Fluid Mechanics Films in the US in the 1960’s.
Each film is also accompanied by a set of notes. I recommend them highly, and will point
out the appropriate ones throughout this course.
The following 3 sections are useful for the course. For the purposes of an examination, I
would expect you to know the definition of div, grad, curl and the Laplacian in Cartesians
and grad and the Laplacian in plane polars. Other definitions would be provided.
• The cross (or vector) product is u×v = (u2 v3 −u3 v2 )r̂+(u3 v1 −u1 v3 )ŷ+(u1 v2 −u2 v1 )ẑ
∂φ ∂φ ∂φ
• The gradient is ∇φ = , ,
∂x1 ∂x2 ∂x3
∂f1 ∂f2 ∂f3
• The divergence is ∇ · f = + +
∂x1 ∂x2 ∂x3
∂f3 ∂f2 ∂f1 ∂f3 ∂f2 ∂f1
• The curl is ∇ × f = − r̂ + − ŷ + − ẑ
∂x2 ∂x3 ∂x3 ∂x1 ∂x1 ∂x2
∂ 2φ ∂ 2φ ∂ 2φ
• The Laplacian is ∇2 φ = ∇ · ∇φ = + +
∂x21 ∂x22 ∂x23
∂f1 ∂f1 ∂f1 ∂f2 ∂f2 ∂f2 ∂f3 ∂f3 ∂f3
• (f ·∇)f = f1 ∂x1
+ f2 ∂x2
+ f3 ∂x 3
r̂+ f 1 ∂x1 + f 2 ∂x2 + f 3 ∂x3 ŷ+ f 1 ∂x1 + f 2 ∂x2 + f 3 ∂x3 ẑ
1 r̂ rθ̂ ẑ
• The curl is ∇ × f = ∂/∂r ∂/∂θ ∂/∂z .
r
fr rfθ fz
∂ 2 φ 1 ∂φ 1 ∂ 2φ ∂ 2φ
• The Laplacian is ∇2 φ = + + + 2
∂r2 r ∂r r2 ∂θ2 ∂z
fθ2
• (f · ∇)f = fr ∂f fθ ∂fr ∂fr
r̂ + fr ∂f + frθ ∂f + fz ∂f f r fθ
∂r
r
+ r ∂θ
+ f z ∂z
− r ∂r
θ
∂θ
θ
∂z
θ
+ r
θ̂ +
fr ∂f + frθ ∂f + fz ∂f
z z z
∂r ∂θ ∂z
ẑ
Formulae in spherical polar coordinates
Coordinate system is r = (r, θ, ϕ) where the relationship to Cartesians is x = r sin θ cos ϕ,
y = r sin θ sin ϕ, z = r cos θ.
The unit vectors are r̂ = r̂ sin θ cos ϕ + ŷ sin θ sin ϕ + ẑ cos θ, θ̂ = r̂ cos θ cos ϕ +
ŷ cos θ sin ϕ − ẑ sin θ and ϕ̂ = −r̂ sin ϕ + ŷ cos ϕ.
In the following, f = (fr , fθ , fϕ ) ≡ fr r̂ + fθ θ̂ + fϕ ϕ̂.
∂φ 1 ∂φ 1 ∂φ
• The gradient is ∇φ = r̂ + θ̂ + ϕ̂
∂r r ∂θ r sin θ ∂ϕ
1 ∂(r2 fr ) 1 ∂(sin θfθ ) 1 ∂fϕ
• The divergence is ∇ · f = 2
+ +
r ∂r r sin θ ∂θ r sin θ ∂ϕ
∂ 2φ
2 1 ∂ 2 ∂φ 1 ∂ ∂φ 1
• The Laplacian is ∇ φ = 2 r + 2 sin θ + 2 2
r ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂ϕ2
fθ2 +fϕ2 fϕ2 cot θ
fϕ ∂fr fϕ ∂fθ
• (f ·∇)f = fr ∂f ∂r
r
+ fθ ∂fr
r ∂θ
+ r sin θ ∂ϕ
− r
r̂+ fr ∂f∂r
θ
+ frθ ∂f
∂θ
θ
+ r sin θ ∂ϕ
+ frrfθ − r
θ̂+
fr ∂f
∂r
ϕ
+ fθ ∂fϕ
r ∂θ
+ fϕ ∂fϕ
r sin θ ∂ϕ
+ fr fϕ
r
+ fθ fϕ cot θ
r
ϕ̂
1 Introduction & Basic ideas
1.1 What is a fluid ?
In this course we will treat the laws governing the motion of liquids and gases. A liquid
or gas is characterized by the fact that there is no preferred rest state for the parts it
is composed of. If a hole is made in a water bottle, the water will flow out. If a drop
is placed on a solid surface, it will spread. By contrast, a solid retains a memory of its
original state. If one deforms a piece of metal, it will relax back to its original state once
the force is no longer applied. Collectively, we will call liquids or gases, which share this
fluid property, “fluids”.
Another important property of liquids and gases is that they are “featureless”. Viewed
from a particular point in space, all directions are equivalent. Solids, on the other hand,
often have an internal (lattice) structure. As a result, it makes a difference in which
direction they are deformed relative to their internal structure.
Fluid dynamics is an example of ‘continuum’ mechanics:
We know that this description fails if we observe matter on small enough length scales:
e.g. typical molecule size (∼ 10−9 m) or typical mean free path in a gas (∼ 10−7 m). The
miracle is that on a scale only slightly larger than that, all microscopic features can be
ignored, and we end up with a “universal” description of all things fluid. We will also
ignore all effects of incompressibility, so for the purposes of this course, liquids and gases
are the same thing. We will almost always speak of a fluid, but which can mean either a
liquid or a gas for the purposes of this course.
The motion of liquids and gases is governed by the same underlying prin-
ciples.
Figure 1 shows a jet of water from a bottle. Both the efflux of the water and the
trajectory of the resulting jet are well described by ideal fluid theory, which we will
describe.
Another spectacular success is the theory of flight. The ideal flow of air around a
wing is able to describe the lift necessary for flight, and much more. What is much more
difficult is the theory of drag. Inviscid theory suggests that there should be no energy
cost to flight at all!
Figure 3: Water waves on the surface of a lake.
The last topic we will be covering in this course is the huge area of water waves, see
Fig 3. This includes waves from the scale of millimetres up to huge tsunami waves. The
absence of any solid boundary results in very little friction, so the ideal theory works very
well.
Definition 1.3.1 (Eulerian description of the flow) . This is what the stationary
observer sees. Choose a fixed point, r to measure, for e.g. the velocity u(r, t). This pro-
vides a spatial distribution of the flow at each instant in time. This is the way continuum
equations are usually formulated, and our equation of motion will indeed be an equation
for the Eulerian field u(r, t). If the flow is steady, then u does not depend on time, t:
u = u(r).
Why do we need anything else? The reason is that to make contact with Newton’s
equations, we need to describe the flow as a moving particle would see it. This is the
Definition 1.3.2 (Lagrangian description of the flow) The observer moves with the
fluid. Choose a fluid particle (for example, we can place a small drop of ink in the fluid),
and follow it through the fluid. Measuring its velocity at a given time, t gives its ‘La-
grangian velocity’. Now we describe the whole velocity field this way, by labeling all mate-
rial points. A convenient way of doing so is to choose an initial time t0 , and to label all
fluid particle by their position r = a at that time. Then at time t > t0 , the particle is at
Of course, r(a, t) is very interesting in its own right. For example, it describes the
course of a balloon, launched at time t0 and at position a into the atmosphere. The
Lagrangian velocity is defined as
dr
v(a, t) = . (1)
dt af ixed
By definition, it is the velocity of a particle going with the flow. This is precisely how
velocity is defined in Newtonian Mechanics. As we said earlier, the Eulerian and the
Lagrangian velocity fields contain the same information; the relation between the two is:
An example: Consider logs flowing along a narrowing section of river. A fixed observer
measures the velocity by observing the velocity of logs at a given point in space. By
observing many logs at different positions, he will be able to obtain the entire Eulerian
velocity field u(r, t). Unless the flow conditions are changing, this field will be time-
independent, u(r, t) = u(r). Now imagine each log being ridden by a moving observer,
each of whom reports his velocity as time goes by. If all observers are labeled by their
position a at some reference time t0 , this will produce the Lagrangian velocity field v(a, t).
The logs travel with the fluid and will see the flow accelerating, as the river bed becomes
narrower. Thus v(a, t) is manifestly time-dependent although the Eulerian field is not!
Example 1.3.4 (Hyperbolic flow) Let us consider a very simple model flow for the
river shown above. Consider the two-dimensional, stationary Eulerian velocity field u =
(u, v, 0), defined by
u = kx, v = −ky. (3)
Figure 5: Local velocities of the flow (3). The red lines are streamlines, and denote the
bank of the river.
As can be seen in Fig. 5, the flow speeds up as the river contracts, which we can think of
as being confined by the river banks, shown as the red lines.
Example 1.4.2 (Pathlines of a steady flow) Find the pathlines for the flow field
(3).
Then the equations become
ẋ = kx, ẏ = −ky.
The initial positions a = (a1 , a2 ) at time t0 = 0, are
xy = a1 a2 , (5)
In particular, the Lagrangian velocity field is indeed time-dependent, while the Eulerian
field was steady.
Now let us consider a case where the Eulerian field is also time dependent.
(x − x0 )2
y = y0 + , (7)
2
which is a parabola, but with different coefficients, depending on the initial particle posi-
tion.
Now we come to the definition of a streamline, which describes the geometry of a flow
at a given time t0 . As a result, the pattern of streamlines will in general change in time.
dx dy dz
= = (= λ(s)ds), t = t0 . (9)
u v w
The function λ(s) does not change the shape of the streamline, but only the parameteri-
zation. To complete the picture, we also add arrows pointing in the direction of the flow
along a streamline (see Fig. 6 below).
The streamline pattern will in general be different for different times t0 , since the veloc-
ity field u(r, t0 ) is different. For a steady flow, the pattern is of course time-independent
as well. Streamlines can be visualized by taking a short-time exposure of illuminated
particles in a flow. Taking an arbitrary starting point, one always continues in the local
flow direction.
dr 1 dr
= .
dt λ(s) ds
the two definitions become identical. In other words, time is a particular parameterization
of the streamline.
0 x
Figure 6: The streamlines of the flow (3). Arrows are drawn for the case k > 0.
Since this is a steady flow, we can eliminate t from the particle path to obtain (5),
which must also be a streamline. Alternatively, we have
dx dy
=−
kx kv
according to (9). Integrating, we find ln |x| + ln |y| = C, or
|x||y| = eC ,
which is (5), the equation for a hyperbola. Any such hyperbola is a possible equation for
the river bank, which must be a streamline of the flow, see Fig. 5.
y t=0 y
t=1
0 x 0 x
Figure 7: The streamlines of the flow (10). The pattern changes in time.
We know how to measure the time derivative of a physical quantity associated with
the fluid, for example that of the temperature T (r, t), at a fixed point in space (the
∂T
Eulerian derivative); it is . This quantity will describe the change of temperature at
∂t
a fixed location, for example air temperature in Bristol. However, this quantity will not
be a measure of how a mass of air heats up or becomes colder. The reason is that air is
swept away by the prevailing flow field u(r, t). In other words, to describe the change of
temperature of a piece of air, we need to consider the rate of change of T (r, t), following
a fluid particle with trajectory r(a, t).
Definition 1.5.1 (Lagrangian time derivative)
DT
The Lagrangian derivative, denoted as , is the time derivative of a field T (r, t)
Dt
along the trajectory of a fluid particle.
Thus using the chain rule, we have
DT dT ∂T
≡ = + (ṙ · ∇)T.
Dt dt ∂t
But according to (1) and (2), ṙ = v(a, t) = u(r, t), and so
DT ∂T
= + (u · ∇)T. (11)
Dt ∂t
A particularly important example is the change in velocity (the acceleration) of a fluid
particle, which we need to apply Newton’s equations to fluid motion. Of course, the
velocity is a vector quantity, which means we have to apply (11) to each component:
Du Du1 Du2 Du3
= , , .
Dt Dt Dt Dt
The the acceleration of a fluid particle becomes
Du ∂u
= + (u · ∇)u. (12)
Dt ∂t
Note: The Lagrangian derivative (i.e. following fluid particles) is given in terms of
Eulerian (i.e. fixed point) measurements. It is vital to understand exactly how to com-
pute expressions like (u · ∇)u for a given velocity.
Example 1.5.2 (advective derivative) Consider an accelerating fluid flow, such as
the log flowing through a narrowing channel. Suppose
Then
∂u
• The flow is steady since = 0.
∂t
• The advective term is
∂ ∂
(u · ∇)u = (U + kx) + (U − ky) (U + kx, U − ky, 0).
∂x ∂y
A fluid occupies the space of which V is a subset. The fluid has velocity u(r, t) and
density ρ(r, t). Fluid can flow in and out of V , and its density can change (within V ).
ds
V S
Figure 8: The mass inside a volume V can change only because because mass flows in or
out through the surface S.
dS
Figure 9: The flux through a surface element is determined by the inner product u · n.
The flux of mass out of V (n points outward) is calculated using the following ob-
servations, see Fig. 9. In time dt the fluid on the surface element dS is transported a
distance |u|dt in the direction of u. Projected onto the direction n perpendicular to S,
this corresponds to a distance u · ndt, and in turn to a volume (u · n)dtdS. Then the
total mass flowing in or out of dS is ρ(u · n)dtdS, and the rate of mass transport is this
quantity divided by dt. This means the total flux of mass out of V is
Z
ρu · ndS.
S
Definition 1.6.1 (mass flux density) The quantity j = ρu is called the mass flux den-
sity.
Now the rate of change of mass in V must equal the rate of change of mass in/out of
V through S, which means that
Z Z
∂
ρ(r, t)dV = − j · ndS (13)
∂t V S
Since V does not change with t and using the divergence theorem for the right hand side
of (13) we get Z Z
∂ρ
dV = − ∇ · jdV
V ∂t V
or Z
∂ρ
+ ∇ · j dV = 0.
V ∂t
This is true for any fixed V , so we must have
∂ρ
+∇·j=0 (14)
∂t
at every point in the fluid. This is called the mass conservation equation or the continuity
equation.
Remark 1.6.2 (conservation laws) The structure of (14) is very general, and applies
to any conserved quantity; an equation of the form (14) is also called a conservation law.
Its basic ingredients are a conserved density (e.g. mass density or energy density), and a
flux (i.e. a mass flux or energy flux).
1.7 Incompressibility
Definition 1.7.1 (Incompressibility) A fluid is said to be incompressible if the density
Dρ
of each fluid ‘particle’ is constant, (i.e. = 0).
Dt
From (14) we have
∂ρ ∂ρ Dρ
0= + ∇ · (ρu) = + u · ∇ρ + ρ∇ · u = + ρ∇ · u
∂t ∂t Dt
So (ρ > 0) an incompressible fluid satisfies
∇ · u = 0. (15)
No fluid is completely incompressible, but even gases are often sufficiently incompress-
ible for (15) to apply to their motion. Incompressibility is a valid approximation if
Of course this leaves out everything to do with sound waves, which are due entirely to
compressible effects.
Theorem 1.8.1 (Vector potential) If ∇·u = 0, then there exists a vector field A(r, t)
s.t.
u = ∇ × A. (16)
The vector field A is called the vector potential. If u is represented like that, u is clearly
incompressible. However, A is far from unique: if the gradient of any scalar function is
added to it, the same u results. Thus to find a unique A, constraints must be applied
to it, and we are back to square one! However, the situation is different if the flow is
two-dimensional. In that case, A must point in the direction perpendicular to the plane,
and we can write A = ψ(x, y, t)ẑ, where ψ is a single scalar quantity. We will consider
the case of a Cartesian and of a polar coordinate system in the plane separately.
Cartesian coordinates : Here A = ψ(x, y, t)ẑ, so ψ is written as a function of the two
Cartesian coordinates x, y in the plane. A single scalar function ψ corresponds exactly to
the expected number of degrees of freedom of the flow: there are two components of the
velocity, and one constraint ∇ · u = 0. Then
∂ψ ∂ψ
u=∇×A= ,− ,0 ,
∂y ∂x
which determines how the velocity field is determined in terms of the stream function. In
other words, we have represented the components u, v of a two-dimensional velocity field
as
∂ψ ∂ψ
u= , v=− . (17)
∂y ∂x
By construction, this representation guarantees that ∇ · u = 0, as can be checked directly.
In addition, ψ has a very convenient and useful physical interpretation. Let (x(s), y(s)
be a streamline, parameterized by s. Then the change of ψ along the streamline is
∂ψ ∂ψ ∂x ∂ψ ∂y ∂ψ ∂ψ
= + = λ(s)u + λ(s)v = λ(s) (−vu + uv) = 0,
∂s ∂x ∂s ∂y ∂s ∂x ∂y
having used (8) in the first step, and (17) in the second. It follows that ψ(x, y, t) = const
along a streamline of the flow. This motivates the following
ψ= constant
Note: For steady flows, the streamlines do not cross each other and fluid does not
cross the streamlines.
From the first equation, ψ = a ln(x2 + y 2 )/2 + f (x), and then from the second equation
f 0 (x) = 0. Thus the streamfunction is
a
ln x2 + y 2 ,
ψ= (19)
2
only defined up to a constant, of course. Clearly, the streamlines ψ = const are circles,
as expected.
Example 1.8.4 (Vortex in polar coordinates) From the above it follows that the stream-
function (19) of Example 1.8.3 can be written in polar coordinates as
a
ln x2 + y 2 = a ln r.
ψ=
2
Example 1.8.5 (Source in polar coordinates) In the case of a source, the flow is
purely radial, see Fig. 12, so that uθ = 0 and ur is independent of θ.
The flow is incompressible, so the flux at the origin equals the flux through any closed
boundary surrounding the origin. Found, most conveniently, by measuring the flux
through a circle radius r centred at the origin. The source strength is
I Z 2π Z 2π
A
m= u · nds = u · nrdθ = rdθ = 2πA,
circle 0 0 r
where n = r̂ in polars.
This means the mass flow through a circle surrounding the origin is independent of
the radius of the circle, and equals the source of mass flux at the origin. In fact, one can
show that the mass flux through any closed surface is m, using Green’s theorem (how
exactly?). In conclusion, the two-dimensional velocity field of a source is
m
u= r̂, (21)
2πr
and the stream function is
mθ
ψ= . (22)
2π
The velocity fields and streamfunctions derived here will be summarized in Appendix D.
Note: In Cartesians,
m
ψ= arctan(y/x).
2π
Another take on the above statement is provided by the following interesting property
of the streamfunction:
Theorem 1.8.7 (Flux across a line) Let r0 and r1 be two points in the plane, and ψ
the streamfunction. Then the volume flux crossing any curve from r0 to r1 is given by
ψ(r1 ) − ψ(r0 ).
Proof: The volume flux Q across the curve C joining the points r0 and r1 is
Z r1
Q= u · n ds,
r0
where ds is the arc length along the curve. We choose n in the mathematically posi-
tive, counterclockwise direction. Since the tangent vector in the direction of the curve is
t = ( dx, dy)/ ds, this is the orthogonal unit vector n = ( dy, − dx)/ ds, pointing in the
counterclockwise direction, going from r0 to r1 . Now
Z r1 Z r1 Z r1
∂ψ ∂ψ ( dy, − dx) ∂ψ ∂ψ
Q= ,− . ds = dy + dx = dψ = ψ(r1 ) − ψ(r0 ),
r0 ∂y ∂x ds r0 ∂y ∂x r0
as claimed. In particular, if r0 and r1 lie on the same streamline, the flux across the
streamline vanishes, which is obvious. Applied to the source (21), choose r0 = (r, 0) and
r1 = (r, 2π) in polar coordinates, so that ψ(r1 ) − ψ(r0 ) = m, according to (22), which is
indeed the source strength.
2 Flow dynamics for an incompressible inviscid flow
2.1 Forces on a fluid
Fluids move in response to the forces that act on each fluid particle. In order to apply
Newton’s second law in straightforward manner, we need to define a fluid “particle” (or
fluid element) in such a way that the mass is constant. We also want it to be big enough
(encompassing a large number of molecues), such that continuum theory applies. This
can be acccomplished for example by marking each molecule in an initial lump of fluid
(for example a cube). At later times, we think of the fluid particle as the volume δV
containing the original molecules. As the molecules move about, the fluid volume will
deform, but its mass remains constant.
There are two types of forces on a fluid element:
(i) Body forces are forces on a each parcel throughout the bulk of the fluid. In a gravi-
tational potential, the force is proportional to the volume δV of the fluid element.
In the typical case of a uniform gravitational field (as near the surface of the earth)
the force is δFb = ρδV g. We will denote the force density (per unit volume) by f (r).
(ii) Surface forces are transmitted across a surface element dS of a fluid parcel, so the
force is exerted by the fluid on the exterior of δV or vice versa.
When the fluid is a rest, the surface force must be in the direction of the normal, n,
since by definition a fluid cannot sustain any shear (it would flow). When a fluid is in
motion, tangential components of the force on dS can occur. These are associated with
viscosity which describes the effect that one layer of fluid in motion has on an adjacent
layer.
Definition 2.1.1 (Inviscid fluid) A fluid is said to be inviscid (also known as an ideal
fluid), when surface forces act in the normal direction only.
For an inviscid fluid, the ‘surface stress’ (i.e. Force per unit area) is in the direction
n normal to an infinitesimal surface element dS even when the fluid is in motion. In
other words, the force exerted by the exterior fluid on any surface element dS of the fluid
element δV is dFs = −pndS, where p(r, t) is the pressure. The pressure has units of force
per unit area. It is directed inwards because fluids are usually in a state of compression.
if the fluid particle is sufficently small so ∇p can be treated as constant over its volume.
Thus δF = (f (r, t) − ∇p)δV , and since we also have δm = ρ(r, t)δV , (23) becomes
Du
ρ = −∇p + f . (24)
Dt
Using the definition of the convective derivative (12), this can be written explicitly as
∂u
ρ + u · ∇u = −∇p + f . (25)
∂t
Either (25) or (24) are known as Euler’s equation (due to Euler 1756), on which the rest
of this course will be based. Notice that in deriving (25), we have made no assumption
about the fluid being incompressible.
What at first seems strange about Euler’s equation is that the pressure p on the right
is not specified. Indeed, a casual count shows that (25) is a set of three equations for 4
unknowns (three components of u and the pressure. We need another equation! In this
course, this will be the incompressibility condition (15): ∇ · u = 0. Together with (25),
we then have 4 equations for 4 unknowns.
A good analogy for this situation comes from point mechanics: Euler’s equation (25)
plays the role of Newton’s equation ma = F. However, only the body forces f (such as
gravity) are specified explicitely, p is unknown. The reason is that p plays the role of
a constraint force, with the constraint being ∇ · u = 0. We have to find p, such that
the constraint is satisfied at all times. Just as in mechanics, solving the constraint is the
most difficult bit. While Lagrangian mechanics teaches us how to solve the problem by
choosing an appropriate coordinate system, we don’t have this luxury in fluid mechanics:
it is not possible in general to find a representation of the velocity field u with two
independent components, which satisfies the constraint identically. The only exception,
as we have seen, is the case of two-dimensional flows, for which we have the stream
function representation.
for any closed surface S inside or bounding the fluid. This is the momentum integral theorem,
of which we will make lots of use!
We now give a few simple examples of flows which are solutions of (25) in the absence of
an external force f = 0. In each case, it is crucial to also confirm that the incompressibility
condition ∇ · u = 0 is satisfied. Expressions for (u · ∇)u in various coordinate systems
are provided in chapter 0. We assume there is no external force.
Example 2.3.1 (Uniform flow) We have u = U. This must be a solution, as each
particle travels along straight lines, and thus no forces are involved.
Obviously, ∇ · u = 0, so the flow is incompressible. Turning to (25), the flow is steady,
∂u
and (u · ∇)u = 0, since u is constant. Thus ∇p = 0, and it follows that p = const,
∂t
as expected.
Example 2.3.2 (Source in three dimensions) The flow field must point radially out-
ward from the origin, and thus u = f (r)r̂. Then in spherical coordinates,
1 ∂(r2 f (r))
∇·u= .
r2 ∂r
For this to vanish, we must clearly have f (r) = A/r2 . The total flux from the origin must
equal the flux through a sphere of radius R. Since the outward normal is n = r̂, we have
Z
m= u · ndS = 4πR2 f (R) = 4πA.
SR
∂ur m2 ρ
ρ(u · ∇)u = ρur r̂ = − 2 5 r̂ = −∇p.
∂r 8π r
Thus the pressure only depends on the radial coordinate, with
∂p m2 ρ
= 2 5,
∂r 8π r
and so
m2 ρ
p = p0 − .
32π 2 r4
Thus with increasing r, the pressure increases. This is expected, as the flow slows down
in the radial direction, which means a force must be holding it back.
2.4 Hydrostatics
For a fluid at rest, u = 0 and so it follows from (24) that
∇p = f .
S n Stationary
fluid, density ρ
V
Figure 13: The force on a submerged body equals the weight of the displaced fluid.
Therefore Archimedes’ principle becomes: force on body = weight of fluid displaced is ρgV .
A body will float if its total weight is smaller than that of the fluid it displaces. According
to an anecdote, Archimedes used this principle to determine if a crown was made of pure
gold.
Figure 14: Left: the crown and a gold nugget used for reference have the same weight.
Right: The crown of lower density receives a greater upward force, so it becomes lighter
when submerged.
patm
patm
H1 p(z)
1 z p2(z) H2
Figure 15: Forces act on both sides of a lock gate; everything is per unit length (into the
page).
Example 2.4.2 (Lock gates) Consider water on both sides of a lock gate.
To the left of the gate, we can use (32), and the atmosphere surrounding the fluid
surface is at constant pressure patm . Since the pressure must be continuous across the
interface (we will derive this condition more formally in the next section), we also have
p = patm on z = H1 , just inside the fluid. It follows that
Thus the force exerted by the fluid on the gate (per unit width of the gate) is
Z H1
F1 = p1 (z)ndz, (33)
0
where n = x̂ is the vector pointing out of the fluid toward the gate. Thus the x-component
of the force from the left is
Z H1
F1 = F1 · x̂ = p1 dz = patm H1 + 21 ρgH12 ,
0
and the x-component of the force (per unit width) exerted by the fluid on the gate from
the right is Z H1
F2 = − p2 dz = −patm H1 − 12 ρgH22 ,
0
where now n = −x̂, and hence the - sign. Thus the net force in the x-direction is
1
F = F1 + F2 = ρg(H12 − H22 ),
2
which is toward the right (H1 > H2 ). In one of the problems, we will calculate the torque
on the gate.
2.5 Kinematic boundary condition
As with any PDE, it is important to understand the boundary conditions Euler’s equation
has to satisfy. The kinematic boundary condition expresses the conditions on the motion
of the fluid at the boundary. We will deal with solid boundaries first, and then generalize
to moving boundaries, for example those between two fluids or between fluid and air.
Figure 16: A solid boundary with normal pointing into the fluid.
Let us consider the flow near a solid surface, whith normal n pointing into the fluid.
We begin by assuming that the solid be at rest. At any point on the surface, it is clear
that the fluid cannot penetrate the solid. This means that the normal component of the
velocity must be zero on the surface: u(r, t) · n = 0 for any r ∈ S. Instead, the flow can
only have a component parallel to the solid.
Now let us take the slightly more general situation that the solid is moving. At any
point on the surface, we can go into a frame of reference moving with the solid, such that
the point is at rest in this coordinate system. This leads us to the requirement
Example 2.5.1 (A solid wall at rest) Let the wall occupy the plane z = 0, and the
fluid be at z > 0. Then n = ẑ, and the boundary condition is u · ẑ = 0, or uz = 0 for
z = 0.
Note that the other components of the velocity field, ux and uy , are arbitrary! The fluid
can “slip” over the solid without resistance, which is consistent with out notion of a
frictionless or ideal fluid.
Figure 17: A solid boundary with normal pointing into the fluid.
Oil
S=0
Water
Figure 18: The interface between a phase of oil and a phase of water is given by the level
curve S(r, t) = 0.
Let us apply this to the interface between two liquids, as shown in Fig. 18. Then
DS ∂S
= + u(oil) · ∇S = 0, on S = 0 from above
Dt ∂t
DS ∂S
= + u(water) · ∇S = 0, on S = 0 from below
Dt ∂t
Now ∇S is normal to the surface S = const, so n = ∇S/|∇S| is the unit normal to the
surface S = 0. It follows that
This expresses the intuitive notion that the normal component of the velocity on either
side of the interface must be equal.
Note that (36) expresses exactly the same thing as (34) for a solid boundary, except
that the velocity of the solid is now prescribed (and can only be a translation or rotation).
ur = Ṙ, (38)
which is very intuitive, and could have been written down without any calculation!
n (s) oil
Vε ε C
water
p+ = p− , (40)
which says that the pressure is continuous across an interface. This is very intuitive, and
in fact we have already used (40) in example 2.4.2. In the atmosphere, the pressure has
the constant value patm , so (40) says that the fluid pressure at a liquid-air interface must
be p = patm .
Figure 20: A jet impinging perpendicularly on a plate. The dashed line marks the control
surface S.
Now Z Z Z Z
pndS = (p − patm )ndS + patm ndS = − (p − patm )ẑdS
S S S Sw
Z Z
since patm ndS = ∇patm dV = 0 and p = patm everywhere else. Also,
S V
Z Z Z
2
ρu(u · n)dS = ρU ẑdS + ρUr2 r̂dS
S Sj Sf
since u · n = 0 on the wall and on the surface of the liquid. Thus (41) becomes
Z Z Z
2
(p − patm )ẑdS = ρU ẑdS + ρUr2 r̂dS.
Sw Sj Sf
Multiplying this by ẑ to only consider the z-component of the momentum balance one
obtains Z
(p − patm )dS = ρU 2 A,
Sw
where A is the cross section of the jet. But then the force on the wall in the ẑ-direction is
Z
Fz = − (p − patm )dS = −ρU 2 A.
Sw
assuming that the body forces possess a potential Φ. We can transform the term on the
right using the vector identity (see Appendix A)
u · ∇u = ∇( 12 u · u) − u × ω (42)
−u × ω = −∇(p/ρ + Φ + 21 u2 ). (43)
To make the left hand side disappear we take the dot-product with u on both sides to
find
u · (u × ω) = 0 = u · ∇(p/ρ + Φ + 21 u2 ).
dr
The definition of a streamline is = λ(s)u, so
ds
d dxi ∂
(p/ρ + Φ + 12 u2 ) = (p/ρ + Φ + 12 u2 ) = λ(s)u · ∇(p/ρ + Φ + 12 u2 ) = 0.
ds ds ∂xi
Consequently, the quantity
In the problems we show that the energy flux is jE = uH. Thus (44) can be seen
as a statement of energy conservation, where the Bernoulli quantity H is conserved for
a fluid element along a streamline. Note: A very nice discussion of the Bernoulli effect
and of pressure forces is presented in the film: “Pressure fields and acceleration”, found
on Youtube.
1 2 m2
u = .
2 32π 2 r4
According to (44),
m2 ρ
p=− + const,
32π 2 r4
where the constant could in principle be different for each streamline. However, the
pressure must go to the same constant far from the source, so we reach the same conclusion
as in Example 2.3.2. However, Bernoulli’s theorem is particularly powerful for more
complicated flows, for which simple solutions are not available, as the next example shows.
Example 2.7.2 (Flow out of a tank) A tank of uniform cross section A0 has a small
hole, with area Ae , and at a height h above the base, see Fig. 21. The height of the fluid
above the hole is H. What is the flow speed out of the drain ?
patm area
Ao
H Uo area
Ae
Ue
z streamlines
h
Finally we can fix a total filling height (measured from the floor) to be Hf , so that
Hf = H + h. Where do we have to place h to maximise x? To answer this question,
we maximise q
x = 2 (Hf − h)h
with respect to h. This gives h = 21 Hf , and so
xmax = Hf .
Experiment: We took data from the bottom hole, which is at h = 2.4cm. The length
of the trajectory x was measured as for the filling height Hf = H +h as given in the table:
p
Hf [cm] H [cm] 2 (Hf − h)h [cm] xexp [cm]
3.5 ± 0.1 1.1± 0.1 3.2± 0.2 2.2± 0.1
8.5 ± 0.1 6.1 ± 0.1 7.7± 0.2 6.6± 0.1
12. ± 0.1 9.6± 0.1 9.6± 0.2 8.1± 0.1
There are signifcant deviations, but they decrease for increasing Hf .
Example 2.7.3 (Flow through a slowly diverging channel) We can use Bernoulli’s
equation to calculate the pressure distribution in a pipe whose cross section is changing.
area
A2
area
A1
U1 streamline U2
Figure 24: The Venturi tube measures the pressure difference between parts of a pipe
with different cross sections.
2.8 The vorticity equation
The vorticity is the key to many phenomena in fluid mechanics. It is often concentrated
into small regions of space, for example near a point (a point vortex), a line or a sheet (a
vortex sheet). Understnding the motion of the vorticity thus often holds the key to the
entire evolution of the flow. We start with Euler’s equation having rewritten the transport
term using (42), namely
∂u
− u × ω = −∇(p/ρ + Φ + 21 u2 ).
∂t
Here we have assumed that the density ρ is constant. Taking the curl of this equation
eliminates the right hand side:
∂ω
− ∇ × (u × ω) = 0.
∂t
We can use a vector identity:
∂ω
+ (u · ∇)ω = (ω · ∇)u, (50)
∂t
or
Dω
= (ω · ∇)u (51)
Dt
in either form, (50) or (51) are called the vorticity equation. Its advantage is that we have
succeeded to eliminate the pressure from the Euler equation. However, we have payed a
heavy price: (50) contains both the velocity and the vorticity fields u, ω. Thus whenever
ω has been updated using (50), we need to reconstruct u from it.
The situation is more promising in two-dimensional flows, u = (u(x, y), v(x, y), 0) and
so, by definition ω = (0, 0, ω(x, y)) = ω(x, y)ẑ. It is then clear that the term
∂u
(ω · ∇)u = ω(x, y) =0
∂z
and so
Dω
=0 (52)
Dt
That is, vorticity is conserved as it moves with the flow.
On Sheet 1, Q. 1, we worked out how C(t) changed in time. We also showed that the
Eulerian velocity field is
u = (αy, 0, 0), (56)
which clearly is a solution of Euler’s equation.
Since s parameterizes the curve, we can calculate the circulation at any time as
Z Z
∂r
Γ= u · dl = u · ds.
C(t) 2π ∂s
Now
∂r
= a(− sin s + αt cos s, cos s, 0)
∂s
and
∂r
u(s) = = aα(sin s, 0, 0),
∂t
so that
Z Z
2 2
Γ = αa sin s(− sin s + αt cos s)ds = −αa sin2 sds = −παa2 .
2π 2π
However, we will see in a little while that problems arise with this argument when
applied near solid boundaries. The reason is that the circulation theorem applies to loops.
However, it is impossible to surround a point on a solid boundary by a loop that stays
in the flow. Thus no conclusions can be drawn on parts of the flow that originate from a
solid wall. However, even in this case Kelvin’s theorem has very important consequences.
Imagine a localised region of vorticity (shaded region) having been created by shedding
from a boundary, as seen in Fig. 26. Placing a loop Γ around it, which is convected by
the flow, both are convected together, and the vorticity will stay inside the loop. Then
Kelvin’s theorem guarantees that the total strength remaines constant. In other words,
the region of the flow which contains vorticity remains localised within a given region.
u = ∇φ. (57)
The field φ is called the velocity potential or simply the potential of u. If the velocity
field u is given, the potential is calculated from the path integral over u,
Z r
φ= u · dl,
r0
provided of course the flow is indeed potential. In particular, the result is independent of
the path taken, provided the domain is simply connected (see Section 4.2 below).
If the fluid is also incompressible then, ∇ · u = 0 and
(∇ · ∇)φ = ∇2 φ ≡ 4φ = 0. (58)
This is Laplace’s equation. When the fluid domain communicates with a (moving) solid
boundary, we impose the kinematic boundary condition (34), which can be written as
u · n = f (r), or n · ∇φ = f (r) for some given f .
The field equation (Laplace’s) and the boundary conditions are an example of a
Neumann boundary-value problem. An important feature of the equations are that they
are linear and so we can use superposition of solutions. This may seem odd, since the
Euler equation is a nonlinear equation, and does not obey the superposition principle.
However, as we will see very soon, the pressure depends on φ in a non-linear way (later),
which reflects the nonlinear character of the Euler equation.
and φ1 6= φ2 . Let ψ = φ1 − φ2
Z Z Z
2
|∇ψ| dV ≡ ∇ψ · ∇ψdV = ∇ · (ψ∇ψ)dV, since ∇2 ψ = 0
D D ZD
stagnation point
x
We begin with the two-dimensional case of a flow onto a flat plate, see Figure. We
consider the simplest case in which the flow is two-dimensional, u = (u, v, 0). There must
be a stagnation point u = 0 somewhere along the plate, which we place at the origin,
x = y = 0. Performing a Taylor expansion about the stagnation point, to lowest order
the x-component of the velocity will increase linearly as one moves away from the line of
symmetry: u = Ax. On account of incompressibility, we have
∂Ax ∂v
A= =− ,
∂x ∂y
and thus v = −Ay if one moves away from the plate against the flow. Integrating, one
finds the potential:
A
φ = (x2 − y 2 ) (59)
2
It is easy to confirm that this is a solution of Laplace’s equation, and thus satisfies the
fluid equations. More importantly, since (59) was derived from a local expansion, it is
valid locally near any two-dimensional stagnation point, regardless of the shape of the
surface, or of the exterior flow. The stream function corresponding to (59) is
ψ = Axy, (60)
as is confirmed directly from (17).
Figure 27: An axisymmetric flow around a so-called Rankine body, showing a stagnation
point near the tip of thre body to the left. The flow in the neighborhood of this point
will be described by (61).
The axisymmetric version of stagnation point flow is relevant if any axisymmetric body
is placed in an oncoming stream. The potential is in that case (exercise)
A 2
φ= (r − 2z 2 ). (61)
4
∂u
− u × ω = −∇(p/ρ + Φ + 21 u2 ).
∂t
For irrotational flows, u = ∇φ and ω = 0, and so
∂φ p 1 2
∇ + + Φ + 2 u = 0.
∂t ρ
It follows that
∂φ p
+ + Φ + 12 u2 = C(t), (62)
∂t ρ
where C(t) is an arbitrary function of time.
1. Note that (62) applies throughout the fluid, not just on a streamline.
the function C(t) can always be eliminated from the problem. The potential φ̃ leads
to the same flow field! (why?)
The Bernoulli equation (62) is best viewed as an equation for the pressure. Thus solving
a potential flow problem then becomes a two-step process:
2. Given φ, compute u = ∇φ and then (62) to compute the pressure. The pressure
depends on u in a non-linear way, a reflection of the nonlinearity of the Euler
equation.
∂φ R2 R̈ + 2RṘ2
|u|2 = Ṙ2 , and =− .
∂t R
Substituting into Bernoulli gives
which is an ODE for R(t). To integrate it, we multiply by 2R2 Ṙ, so that
d 3 2 p0 2p0 dR3
R Ṙ = 3R2 Ṙ3 + 2R̈ṘR3 = − 2R2 Ṙ = − .
dt ρ 3ρ dt
Integrating, we find
2 p0 3
R3 Ṙ2 = − R + C,
3ρ
where C = 2R03 p0 /(3ρ) is a constant, determined by the initial condition Ṙ = 0 when
R = R0 . Hence
2 p0 R03
2
Ṙ = −1 (64)
3 ρ R3
or r 1/2
2 p0 R03 − R3
dR
=− ,
dt 3ρ R3
where the negative sign must be chosen because Ṙ < 0 by assumption. Integrating up
gives
Z R(t) Z t 1/2 1/2
R3/2 dR
2 p0 2 p0
=− dt = −t .
R0 (R03 − R3 )1/2 0 3ρ 3ρ
If the bubble collapses at time t = tc , then R(tc ) = 0 implies (R = R0 u)
1 1/2
u3/2 du
Z
tc 2 p0
= (65)
0 (1 − u3 )1/2 R0 3ρ
One of the most fundamental problems of fluid mechanics is to understand the flow around
obstacles, which stand in the way of a flow. Equivalently (think of Galilean invariance)
one can think of a body moving inside a fluid which is at rest (for example an aeroplane).
Let us begin studying this problem by considering the perhaps simplest possible body, a
sphere of radius R. The sphere is placed inside a uniform stream, which we choose in the
direction of the z or symmetry axis: U = U ẑ.
Far from the body, the flow will be uniform. The sphere introduces a perturbation to
the flow which decays as r → ∞, i.e. as one moves away from the body. According to
the superposition principle (the Laplace equation is a linear equation), this perturbation
is to be added to the uniform flow:
u = U + ∇φ, (67)
where φ is the potential of the perturbation. The simplest such flow we know is a source,
with potential φ ∝ 1/r. However, such a flow cannot describe the physical situation at
hand. Far from the sphere, the total mass flow through a closed control surface should be
zero, not finite as for a source. Another way of putting it is that the flow field of a source
decays too slowly (like 1/r2 ) away from the body.
This can be helped by taking the derivative of the source solution. Taking the deriva-
tive in the µ-direction, one arrives at the potential
1 µ·r
φ=µ·∇ =− 3 , (68)
r r
which is called the dipole flow (why?). The dipole is oriented in the µ-direction and has
strength |µ|. This flow is also a solution of Laplace’s equation, since
1 1
4 µ·∇ = µ · ∇4 = 0.
r r
In summary, from (67) and (68) our ansatz for the velocity is
µ j xj µi xi
ui = U δi3 + ∂i φ = U δi3 − ∂i 3
= U δi3 − 3 + 3µj xj 5 ,
r r r
or
1 r
u = U ẑ − µ − 3µ · r . (69)
r3 r2
Now we have to satisfy the boundary condition on the surface of the sphere, which is
u · n = 0. The normal vector to the sphere is n = r/R. Now on the surface of the sphere
(r = R):
1 µ · n µ·n
u · n = U (ẑ · n) − 3 µ · n − 3 3 = U (ẑ · n) + 2 3 ,
R R R
3
UR
so choice µ = − ẑ will guarantee that this is zero for r = R. In summary, the velocity
2
field which satisfies all the boundary conditions is (z = ẑ · r):
3
U R z
u = U ẑ + ẑ − 3 n . (70)
2 r r
Now we want to calculate the pressure field, and in particular the pressure on the
surface of the sphere. This will permit us to calculate the total force on the sphere. On
the surface we have
3U z
u= ẑ − n ,
2 R
so
9U 2 z2 9U 2 z2 9U 2
2 z
1 − cos2 θ ,
u = 1 − 2 ẑ · r̂ + 2 = 1− 2 =
4 R R 4 R 4
z
using that = cos θ.
r
According to Bernoulli’s equation
u2
p+ρ = C(t),
2
everywhere in the fluid. But the pressure is p = patm far from the sphere, and the flow is
uniform: u2 = U 2 , which implies that C = patm + ρU 2 /2. As a result, the pressure on the
surface of the sphere is
ρ(U 2 − u2 ) ρU 2
9 cos2 θ − 5 ,
p = patm + = patm + (71)
2 8
Figure 29: The pressure distribution around a sphere as function of the angle.
This result has the remarkable property that the pressure distribution is symmetric
around the equator of the sphere θ = π/2. Thus while the sphere is pushed by the flow in
the downstream direction, there is an equally high overpressure in the back. As a result,
without further calculation, we can conclude that the total drag force on the sphere is
zero. By definition, the drag on a body in a uniform stream is the force in the direction
of the stream. The force perpendicular to the stream is called the lift, which is of course
zero if the flow is axisymmetric, as is the case here.
Let us check these statements by doing the calcuation explicitly. Integrating over the
surface of the sphere, the total force is
Z Z 2π Z π
2
F=− pndS = −R pn sin θdθdφ,
r=R 0 0
where n = (sin θ cos φ, sin θ sin φ, cos θ) is the normal in spherical coordinates. Performing
the φ-integral over the x and y components, we obtain
Z 2π Z 2π
cos φdφ = sin φdφ = 0,
0 0
so the first two components of F vanish. This is clear because on account of the symmetry
of the problem, there can be no component normal to the direction of the flow, which is
in the z-direction. It remains to calculate the force in the flow direction (n · ẑ = cos θ):
Z 2π Z π 2 Z π
2 2 ρU
9 cos2 θ − 5 cos θ sin θdθ =
Fz = F · z = −R p cos θ sin θdθdφ = −2πR
0 0 8 0
Z 1 1
π 2 2 2
π 2 2 9 4 5 2
− ρR U 9x − 5 xdx = − ρR U x − x = 0,
4 −1 4 4 2 −1
since the solution that decays like 1/r is a source or a sink. Now the force on the body is
calculated from integrating the pressure force over the body:
Z
F = − pndS;
S
the minus sign comes from the normal pointing outward, opposite the direction of the
∂φ
pressure force. To find p, we use the Bernoulli equation (62) with = 0, since the flow
∂t
is steady:
p + ρU 2 /2 + ρv2 /2 + ρU · v = patm + ρU 2 /2.
In other words,
Z Z Z
ρ 2
F = − patm ndS + v ndS + ρ (U · v)ndS.
S 2 S S
The first integral on the right is zero, as we have seen before. As a mathematical corollary,
we will show below that Z Z
2
v ndS = 2 v(v · n)dS. (72)
S S
But on the surface of the body the normal component of the total velocity vanishes:
n · (U + v) = 0, so the force becomes
Z Z
F = ρ [v(v · n) + (U · v)n] dS = F = −ρ [v(U · n) − (U · v)n] dS.
S S
In components, this means that
Z
Fi = −ρUj (vi nj − vj ni ) dS. (73)
S
From the symmetry of this expression it is clear immediately that the drag of the body,
which is defined as
D =F·U
is zero. However, the components of F normal to U are also zero. Let V be the volume
exterior to the body but bounded by a closed surface S∞ very far from the body (see
Figure). Then it follows from the divergence theorem that
Z Z Z
∂vi ∂vj
− dV = − (vi nj − vj ni ) dS − (vi nj − vj ni ) dS;
V ∂xj ∂xi S S∞
note that n points into the volume V . If v decays faster than 1/r2 at infinity, the integral
over S∞ goes to zero as the radius of S∞ goes to infinity. As for the integrand of the
volume integral on the left, it clearly vanishes if i = j. If i 6= j, then the integrand is (up
to a sign) ωk where k 6= i, j. But since the flow is potential, this is again zero. It follows
that the surface integral appearing on the left hand side of (73) is zero, and thus the force
is altogether zero.
The corollary
We now demonstrate (72). From the divergence theorem we conclude that
Z Z Z
2 2
v ni dS = − ∂i v dV − v2 ni dS
S V S∞
The integrand of the second integral on the right behaves like 1/r6 , so its contribution
vanishes. To deal with the volume integral, we note that according to Appendix A,
∇v2 = 2(v · ∇)v + 2v × (∇ × v), and so ∇v2 = 2(v · ∇)v for an irrotational flow. Since
the fluid is incompressible, we also have ((v · ∇)v)i = vj ∂j vi = ∂j (vi vj ). Thus
∂i v2 = 2∂j (vi vj ),
using the divergence theorem in the last step. But this proves the corollary.
This is a truly remarkable result of potential flow theory, but it also presents a series
problem, as absence of drag is clearly not in accord with observation. If our argument were
completely airtight, airline companies would surely be doing some serious overcharging!
(maybe they do anyway).
3.7 Drag
Figure 31: A wake of almost stagnant fluid forms behind an obstacle in a strong flow.
The experimental picture of a sphere in a stream reveals what the problem is: the flow
around the sphere is far from rear-aft symmetric. While the flow in front of the sphere
is nicely attached to the body and well described by (70), the back is very different. The
so-called Reynolds number of the flow is
U Rρ
Re = ≈ 1.5 × 104 (74)
η
which is a dimensionless measure of the size of the viscosity η. Since the Reynolds number
is very large, one might believe that viscous effects are small and the potential flow as-
sumption represents a faithful representation. However, this is not the case. In particular,
the flow does not remain attached to the body, and rather there are fluid particles which
originate from the surface of the sphere and the enter the liquid. This is also the place
where vorticity enters the flow. Clearly, vortices have formed in so-called wake of the flow.
Apart from the vortices, the wake appears to be a relatively stagnant region of the
flow, which is thus at constant pressure p = pcav , provided it remains closed (which it
seems to be in the experimental Figure). To estimate the value of pcav , we assume that
the pressure is continuous across the point of separation.
It is also interesting to consider the streamline that separates from the body and
bounds the cavity. Since the pressure is constant in the cavity, the pressure on this
streamline is also constant, by continuity of the pressure. Such a streamline is called
a free streamline. According to the steady Bernoulli equation (44) the velocity along
this streamline is also constant. Therefore, there is a jump of the velocity at the free
streamline to the almost vanishing velocity in the wake. (Nothing prevents the velocity to
be discontinuous if there is no viscosity). Such a surface over which the velocity changes
abruptly is called a shear layer or a vortex sheet. Now pressure in the cavity will be the
pressure on the separating streamline, as it leaves the body.
To do any calculation of the effect of the wake, one needs an estimate of the point
where the flow leaves the body, to produce the wake. A very simple criterion (which
can be justified in more detail using full viscous theory), is the concept of the “adverse
pressure gradient”. At the front of the body the pressure maximum, and then decreases
to a minimum value at the equator, where the speed is maximum; from there the pressure
decreases again. It is intuitive and can be motivated in greater detail from the viscous
flow theory, that the flow along the body is stable as long as the pressure decreases along
the path of a fluid particle. If however a fluid article has to push against an increasing
pressure (“adverse pressure gradient”), it prefers to leave the surface of the body.
Thus on the basis of the pressure distribution which we have calculated, one can give
an estimate of the place where separation is going to occur, which is at θ = π/2 in the case
of a sphere. Now we can attempt to recalculate the drag force on the sphere, assuming
separation at θs = π/2. The pressure in the cavity will be
5
pcav = patm − ρU 2 :
8
in our calculation, there is an underpressure in the cavity which pulls the sphere along.
We are now in a position to calculate the drag on the sphere, doing the front and the
back separately. The result cannot depend on patm which exerts equal forces on front and
back; therefore, we can put it to zero. Since n · ẑ = cosθ, the force from the front in the
flow direction is
1 2 2 2π π
Z Z Z
(f ront)
Fz =− p(θ) cos θdS = − ρU R sin θ cos θ(9 cos2 θ − 5)dθdφ =
f ront 8 0 π/2
Z 1
π 2 2 π
ρU R x(9x2 − 5)dx = − ρU 2 R2 ,
4 0 16
and from the back
Z 2π Z π/2 Z 1
5π 2 2 5π 2 2
Fz(back) = −R 2
sin θ cos θpcav dθ = ρU R xdx = ρU R .
0 0 4 0 8
Thus the total force or drag is
9π 2 2 9
D = Fz = ρU R = ρAU 2 ,
16 16
where A = πR2 is the projected area of the sphere.
For a general body, the drag coefficient Cd is defined by
ρ
D = U 2 ACd ,
2
and so Cd = 9/8 for our calculation of the sphere. We hasten to add that our calculation
is not particularly good, see the Figure. In the relevant range of Reynolds numbers,
CD ≈ 0.5. Our intention is to give a general idea of where the drag is coming from, and
how d’Alambert’s paradox arises.
Figure 32: The drag coefficient CD of a sphere as a function of Reynolds number.
4 Two-dimensional flows
If the flow is in the plane, and there are only two independent variables x, y, the flow prob-
lem is much simplified. Moreover, we will see that some of the physics is fundamentally
different from three dimensions. One fact we know of already is that no vorticity is created
in two dimensions, ω is simply convected with the flow. Of course, real flows are never
truly two-dimensional; however if the geometry extends very far in one direction (think
of the wing of an aeroplane), a two-dimensional description is a good approximation.
We want to find the flow around a stationary cylinder in a steady stream U = U x̂. This
is very closely analogous to the flow around a sphere; the effect of the sphere is modelled
by a (two-dimensional) dipole, which is the derivative of a source, whose velocity field is
given by (21). Thus the corresponding potential is
m
φ= ln r, (75)
2π
and a dipole in the µ-direction has potential
µ·r
φ = (µ · ∇) ln r = .
r2
The ansatz for the velocity field is thus
µ 2(µ · r)r
u = U x̂ + ∇φ = U x̂ + − .
r2 r4
Now the boundary condition for r = R is u · n = u · r/R = 0, and from the velocity field
we find (r = R):
x µ·x
u·n=U − .
R R3
Thus the boundary condition is satisfied if we choose
µ = R2 U x̂ = R2 U.
R2
u=U+ [U − 2n(U · n)]. (76)
r2
Now we plug this into Bernoulli’s equation to compute the pressure:
u2 p U 2 patm
+ = + .
2 ρ 2 ρ
On the surface
u2 r=R
= (2U − 2n(U · n))2 = 4U 2 − 4(U · n)2 = 4U 2 (1 − cos2 θ) = 4U 2 sin2 θ,
so in other words
ρU 2
p = patm + (1 − 4 sin2 θ). (77)
2
The pressure is once more symmetric about θ = π/2, i.e. about the midsection of the
cylinder. It follows that the total force on the cylinder is again zero.
3. Let D be the domain a < r < b, 0 < θ < 2π in cylindrical polars. D is not simply
connected.
i ii
Figure 34: On the left, and example of a simply connected domain: all curves in D can be
collapsed to a point. On the right, a domain which is not simply connected: some curves
can be collapsed to a point, but those going around the hole cannot.
We now place a point vortex (introduced at the beginning of chapter 3) at the centre,
which creates a flow
Γ
u= θ̂. (78)
2πr
The strength of the vortex is measured in terms of its circulation Γ. Since the flow is
irrotational, there exists a velocity potential s.t.
∂φ 1 ∂φ
ur = = 0, uθ = = Γ/(2πr).
∂r r ∂θ
The solution is
Γ
φ(r, θ) = (θ + A), (79)
2π
for any constant A.
The boundary conditions at the outer and inner walls of D are u · n = u · r̂ = ur = 0.
Clearly, (78) satisfies the boundary conditions, while the circulation Γ can assume any
value: the solution is not unique! This is related to another problem with the potential: for
example, φ(r, 0) 6= φ(r, 2π) so the potential is discontinuous along θ = 0 or, alternatively,
φ is multivalued. Evidently, whenever there is circulation
I
Γ= ∇φ · dl 6= 0,
C
a finite amount of potential is picked up as one goes around the curve. The potential of
such a flow is thus necessarily multivalued. This situation can arise only in a multiply
connected domain: if it were simply connected, the closed curve C could be reduced to a
point, and thus the integral would be zero. This can be achieved by placing a branch-cut
in D to make the domain simply connected. In simply-connected domains, φ is unique,
as seen in Fig. 35.
b
a
Needs a branch cut to
make domain simply connected
and the potential single−valued
Now assume any flow field bewteen the two cylinders, which must satisfy 4φ and
n · ∇φ = 0 on the boundaries. To any such solution, we can add the vortex (79), for any
value of the circulation Γ. The result will be another solution, which also satisfies the
boundary conditions. In other words, the flow between the cylinders can have any value
of the circulation. In general, we see that the flow in a non-simply connected domain is
determined only up to a flow with circulation Γ, where Γ can take any value.
R2
φ=U r+ cos θ (80)
r
In the problem without circulation (cf. Fig.33), the stagnation points u = (0, 0) are at
the front and the back of the cylinder, i.e. at θ = 0 and θ = π along the surface r = R. To
find the stagnation points on r = R in the general case (81), we have to look for solutions
of uθ = 0 (ur = 0 by construction):
R2
1 ∂φ Γ
0 = uθ = = −U 1 + 2 sin θ − .
r ∂θ R 2πR
Figure 36: The streamlines around a circular cylinder with circulation Γ < 4πU R, Γ =
4πU R, and Γ > 4πU R.
To calculate the force on the cylinder, we use the unsteady Bernoulli equation,
p + 12 ρu2 = C,
noting that the flow is steady (C(t) = C, ∂φ/∂t = 0). With p → patm , u2 → U 2 as
r → ∞,
p = patm + 12 ρU 2 − 21 ρu2 .
On the surface of the cylinder,
Γ
uθ = −2U sin θ − (83)
2πR
and with n = (cos θ, sin θ), the total force is
Z 2π
F=− pnR dθ.
0
Hence there is an upward lift force on the cylinder (Magnus Effect - see Annalen der Physik
164, 129 (1853)). This can explain the forces acting on spinning objects (for example
footballs or ping-pong balls). A quantitative analysis of the resulting trajectories has been
performed by Dupeux et al., New Journal of Physics 12 093004 (2010).
Figure 37: On the left, an illustration of the direction of the Magnus force. On the right,
an application to table tennis.
Note that the above calculation refers to cylinder at rest, with a flow past it. It
is equivalent to a cylinder moving to the left in a stationary medium. Thus a cylinder
moving to the right will experience a downward force and deviate in the direction of
rotation, which is fairly intuitive, as shown in the Figure. Dupeux et al. investigate the
effect more quantitatively, and calculate the resulting spiral flight path. Among others,
they show that the Magnus effect is responsible for a famous goal by the Brazilian player
Roberto Carlos, see Fig. 38.
. Proof: We have that u = ∇φ, and the stream function ψ satisfies (17). Thus
∂φ ∂ψ
u= =
∂x ∂y
(86)
∂φ ∂ψ
v= =− .
∂y ∂x
But the equations (86) are the Cauchy-Riemann equations for w = φ + iψ, and so w is
analytic. In particular, if w(z) = w(x, y) is analytic then
dw ∂φ ∂ψ ∂φ ∂ψ
= +i = −i + ,
dz ∂x ∂x ∂y ∂y
i.e. w0 (z) is independent of the direction of the derivative.
Properties of the complex potential:
(i) ∇2 φ = ∇2 ψ = 0 follow directly from (86).
∂φ ∂ψ ∂φ ∂ψ
(ii) ∇φ · ∇ψ = + = 0 meaning equipotential curves (where φ = const)
∂x ∂x ∂y ∂y
are perpendicular to streamlines, (where ψ = const).
dw
(iii) = u − iv = qe−iχ where q = (u2 + v 2 )1/2 is the speed of the flow, χ is the angle
dz
dw
the flow makes to the x-axis; is known as the complex velocity (note the minus
dz
sign)!
Then
∂ψ ∂φ
u= = 2kx =
∂y ∂x
, ⇒ φ = k(x2 − y 2 )
∂ψ ∂φ
v=− = −2ky =
∂x ∂y
Hence w(z) = φ + iψ = k(x2 − y 2 ) + i2kxy = kz 2 is the complex potential.
We have seen that the flow can be described as a superposition of a uniform stream and
a dipole. First, a uniform stream in the x-direction has the complex potential
w(z) = U z,
A line vortex intersects a plane in one point, in two dimensions one can thus describe the
vortex by a point (the position z0 = x0 + iy0 of the vortex, and its strength Γ. According
to Kelvin’s theorem, the circulation is conserved, and thus the strength of the vortex is
constant in time. The only thing we need to keep track of is the position of the vortex.
In complex notation, the velocity field generated by a vortex located at z0 is
iΓ d log(z − z0 ) iΓ 1 iΓ z − z0
u − iv = − =− =−
2π dz 2π z − z0 2π |z − z0 |2
Now each vortex moves in the velocity field generated by all the other vortices, and so
dz1
= u − iv|allothervortices
dt
In the simplest case of two vortices of strength Γ1 and Γ2 , located at z1 and z2 , respectively,
the equations of motion are
dz1 iΓ2 z1 − z2 dz2 iΓ1 z2 − z1
= , = . (89)
dt 2π |z1 − z2 |2 dt 2π |z1 − z2 |2
Decomposing into real and imaginary parts, the first equation (89) is
dx1 Γ2 y1 − y2 dy1 Γ2 x1 − x2
=− , = ,
dt 2π (x1 − x2 )2 + (y1 − y2 )2 dt 2π (x1 − x2 )2 + (y1 − y2 )2
and correspondingly for the other vortex. The generalisation for N vortices is obvious
and involves the sum over all the vortices except for the one that is being convected. How
does the motion of two vortices look like ? Let us look at the two possible cases:
Figure 40: Two vortices with the same sense of circulation rotate around a common
center, turning in opposite directions.
The vortex strengths are Γ1 and Γ2 , both being positive (the two vortices rotate coun-
terclockwise). Let the distance between the two vortices be h. The speeds of the two
vortices are then
Γ2 Γ1
U1 = , U2 = ,
2πh 2πh
respectively. From the construction it is clear that at each instant, the vortices move
in a direction perpendicular to the line connecting them. This means they have to go
around in circles, whose centre G is along the line connecting the vortices. The point G
is stationary in space. The some of the radii R1 , R2 must satisfy R1 + R2 = h. Since both
vortices make a full turn in the same time, we must have
R1 U1 Γ2
= = .
R2 U2 Γ1
Thus for example,
Γ2 h
R1 = .
Γ1 + Γ2
For two equal vortices of strength Γ1 = Γ2 = Γ, it follows that R1 = R2 = R. This
means the vortices are always at the opposing sides of a circle of radius R. The time to
complete one cycle (the period) is
2πR 8π 2 R2
T = = ,
U Γ
and so the whole configuration rotates at an angular velocity of
2π Γ
Ω= = .
T 4πR2
Figure 41: Two counterrotating vortices are created by moving a paddle in water; they
travel in the same direction.
Now consider the case that the two vortices have opposite sign. This situation is produced
easily by the stroke of a paddle, see Fig. 41. In fact since the total vorticity of the system
should remain zero according to Kelvin’s theorem, typically Γ1 = −Γ2 . After the paddle
is withdrawn, the vortices continue to move according to the other vortex’ velocity field.
Let the distance between the two vortices again be h.
Figure 42: Two vortices with the opposite sense of circulation rotate around a common
center, turning in the same direction.
The vortices now move in the same direction, and rotate about the same centre. The
radii of this circular motion obeys R1 /R2 = −Γ2 /Γ1 , noting that the circulations have
Γ2 h
opposite signs. We now have R2 − R1 = h, and thus R1 = − . Thus when the
Γ1 + Γ 2
vortices are equal and opposite (Γ = Γ1 = −Γ2 ), the radii of both circles tend to infinity.
Γ
In this case the vortices move in a straight line, and translate uniformly at speed U = .
2πh
This is what’s seen in the experimental pictures of two vortices being created behind a
paddle.
Consider the motion of a vortex located at some position z0 = a + ib in the upper half of
the complex plane. The x axis is the location of a solid wall. Thus we have to satisfy the
boundary condition of no flow through the wall: v = 0 for y = 0. Simply taking the flow
field of the vortex,
iΓ x − a − i(y − b)
u − iv = − , (90)
2π (x − a)2 + (y − b)2
does not satisfy this condition: the flow field must be modified by the presence of the wall.
The method of images is a technique that permits to find solutions to Laplace’s equa-
tion which satisfy the right boundary condition on the solid wall. Near the vortex at
z0 , the solution should look like (90). The method of solution is indicated in the figure:
imagine replacing the wall by another image vortex that is placed at an equal distance
on the other side of the wall. Its position is where an image in a mirror would appear.
And equally like a mirror image, the sense of rotation of the image vortex is reversed.
By symmetry it is clear that the v component of the image vortex has opposite sign, so
on the line of symmetry (the locus of the wall), the two contributions cancel and v = 0.
The boundary condition on the wall is satisfied and we have solved our problem! In other
words, the flow problem we want to solve (one vortex in the presence of a wall) is exactly
the same as the flow problem of two vortices (the original vortex and its image) without
the wall.
Let’s check if our reasoning was correct. It is easiest to do the calculation using
complex notation. If the vortex is at z0 , its image is at z0 . Since the image also has the
opposite sense of rotation, there is a minus sign in front of it. The total potential (vortex
+ image) becomes
iΓ iΓ
w(z) = − log(z − z0 ) + log(z − z0 ). (91)
2π 2π
We have to verify that Im{w} = ψ = const for z = x real! Now since ln(z) = ln(z)
(why?) we have
iΓ iΓ
w(z) = log(z − z0 ) − log(z − z0 ).
2π 2π
But if z is real, z = z, and thus w(x) − w(x) = 0, and so ψ = 0 along the wall.
In fact, we have shown that the problem of a vortex near a wall is mathematically
equivalent to the problem of two counterrotating vortices of equal strength, which in the
previous section we showed to move at constant speed in the direction perpendicular to
the line connecting them, in other words, parallel to the wall. Let us check this fact as
well, using complex notation. The vortex in question moves in the flow field of its image,
i.e.
iΓ x − a − i(y + b)
u − iv = ,
2π (x − a)2 + (y + b)2
evaluated at z = a + ib. Thus
iΓ −2ib Γ
u − iv = = ,
2π 4b2 4πb
Γ
and the vortex moves with a horizontal speed of , which corresponds to the earlier
4πb
result with h = 2b.
This leads us to the following result, which permits to find the flow generated by any
combination of singularities next to a wall. Let the wall be the real axis of the complex
plain, and let f (z) describe all singularities (vortices, sources, etc.) present in the flow
iΓ µ
domain y > 0. For example, f (z) = ln(z − z0 ) for a vortex and f (z) = −
2π 2π(z − z0 )
for a dipole.
Note that in the the second member of (92), we take the complex conjugate of every-
thing except of z itself.
Proof: We have to show that ψ = Im{w(z)} = 0 if z = x is real. But if this is the case
then z = z, and so
Figure 46: Two counterrotating vortices next to a wall. On the left, the configuration; on
the right the experimentally recorded trajectory of one of the vortices.
Figure 47: Via a complex mapping f (z) a corner can be mapped into the upper half plane.
Imagine we want to solve a flow problem in a given domain D. This means we are
looking for a function w(z) (the complex potential) with the following properties:
• w(z) is analytic in the domain D, except perhaps for one or more singularities at z0
(and other places).
To find such a complex potential, it is enough to find a mapping into a new domain D1
such that the flow problem can be solved in D1 . We have the following:
Figure 48: The flowlines of a source in a corner, given by the complex potential (96).
The function ζ = f (z) ≡ z 2 , z = ζ 1/2 maps the first quadrant onto the half-plane.
Namely, if z = reiθ , then ζ = r2 e2iθ = r2 eiφ . Thus the domain 0 ≤ θ < π/2 is mapped
onto 0 ≤ φ < π. The source at z0 maps to a source at ζ0 = z02 . Now the solution in the
ζ-domain is
m m
w1 (ζ) = log(ζ − ζ0 ) + log(ζ − ζ 0 )
2π 2π
as we have seen before, using the method of images. So
m
w(z) = w1 (z 2 ) = log(z 2 − z02 ) + log(z 2 − z 20 )
(96)
2π
m
= (log(z − z0 ) + log(z + z0 ) + log(z − z 0 ) + log(z + z 0 ))
2π
Thus the result can be written as a superposition of four sources of equal strength, located
at z0 , −z0 , z 0 , and −z 0 . The same result could have been obtained (albeit a bit more
laboriously), using the method of images. First, we mirror at the x-axis, then at the
y-axis. Fig. 48 is for the case z0 = 2 + i.
Figure 49: The flowlines of a uniform flow around a cylinder, which arrives under an angle
α.
Consider any flow problem, for example the flow around a cylinder, given by (88):
R2
w1 (ζ) = U ζ + .
ζ
Let ζ = ze−iα ≡ f (z), which is a rotation of the axes by α in the anticlockwise direction.
Namely, for z = reiθ , ζ = rei(θ−α) .
Then in the z-plane,
R2 iα
−iα −iα
w(z) = w1 (ze ) = U ze + e . (97)
z
b2
z=ζ+ , (98)
ζ
called the Joukowski mapping, and illustrated in Fig. 50. Consider a circle of radius R,
whose surface in the ζ-plane is described by the polar representation ζ = Reiθ . Under the
Joukowski mapping,
b2 −iθ b2 b2
iθ
z = Re + e = R + cos θ + i R − sin θ = c cos θ + id sin θ
R R R
is an ellipse with axes length 2c and 2d.
The semiaxes come out to be
b2 b2
c=R+ , d=R− . (99)
R R
Figure 51: The flowlines areound a circle are mapped to an ellipse by the Joukowski
mapping (98).
Now consider the uniform flow past an ellipse. To model an arbitrary angle α between
the direction of the flow and the semi-major axis, we consider the flow around a cylinder
that approaches the x-axis under an angle α:
R2 iα
−iα
w1 (ζ) = U ζe + e .
ζ
In principle, one can find w(z) using the inverse of the Joukowski mapping
1 √
ζ = f (z) = (z + z 2 − 4b2 ),
2
so that w(z) = w1 (f (z)).
However, the resulting expressions are often not so useful. For example, to find the
streamlines, it is much easier to find the streamlines of the w1 (ζ) in the ζ-plane, and then
to transform them using (98). This is how the Fig.51 was produced. If one wants to
calculate the velocity, one uses
dw dw1 1 U (e−iα − R2 eiα /ζ 2 )
u − iv = = = .
dz dζ dz/ dζ (1 − b2 /ζ 2 )
On the cylinder, ζ = Reiθ , so
U (e−iα − eiα e−2iθ ) 2iU sin(θ − α)
u − iv = −2iθ
= iθ . (100)
2 2
(1 − (b /R )e ) (e − (b2 /R2 )e−iθ )
This means there are stagnation points at θ = α and θ = −π + α. This point is where
a streamline leaves the surface. In other words, this streamline (plotted in red) has the
same value of the streamfunction ψ then the surface of the ellipse.
4.9 Lift
Now we want to fly! In principle, we know how to construct the flow around a wing of
arbitrary shape, we only have to find the transformation, starting from a circle. We have
seen already that the key ingredient is to have circulation around the wing. We have
calculated the lift in the case of a cylindrical cross-section, but what is it for an arbitrary
shape? To answer this question, we have the following theorem, which gives the force on
a body of arbitrary shape, using complex notation.
Figure 52: If χ is the angle made by the tangent line with the x-axis, the tangent and
normal is given by (103).
Let C be a simple, closed curve, describing a body in a stready stream U . Let the
complex potential outside of the body be w(z). Then the force F = (Fx , Fy ) on the body is
given by 2
I
iρ dw
Fx − iFy = dz. (101)
2 C dz
p = p0 − 12 ρq 2 ,
where s is the arclength along C. On the other hand, a complex line element along C is
dx dy
dz ≡ dx + idy = +i ds = t ds = eiχ ds.
ds ds
Here we have used that by definition, the tangent vector is
dr dx dy
t= = , ,
ds ds ds
and (103) in the last step.
Now we can convert everything to complex notation, where it proves convenient to
define a complex force F = Fx − iFy . Then
I I
F = Fx − iFy = − pn ds = −i pe−iχ ds
C IC I
iρ
= −i p0 (dx − idy) + q 2 e−iχ ds
C 2 C
I I 2
iρ −iχ 2 iρ dw
= (qe ) dz = dz
2 C 2 C dz
(the term proportional to p0 vanishes because the integral of a total differential over a
closed loop is zero), and we have used (102) in the last step. This proves Blasius’ theorem.
Example 4.9.2 (Cylinder with circulation) Consider the problem of Section 4.3,
R2 iΓ
w(z) = U z + U − log z,
z 2π
so
dw R2 iΓ
=U −U 2 − .
dz z 2πz
Applying (101) to the circle C gives
2
R2
I
iρ iΓ
Fx − iFy = U −U 2 − dz
2 C z 2πz
I
iρ 2 iU Γ −2 −3 −4
= U − + terms {z , z , z } dz
2 C πz
I I
dz dz
Now use Cauchy’s residue theorem = 2πi, n
= 0, n 6= 1 and
C z C z
Fx − iFy = iρU Γ,
Hence the drag (force in the direction of the stream) on an arbitrary body is zero and
the lift force is Flif t = −U ρΓ. Note that the lift force is by definition the force in the
direction normal to the flow.
Proof: Far away from the body,
iΓ
w(z) → U z − log z
2π
(origin inside C) so
dw iΓ
≈U− , z → ∞.
dz 2πz
Since there are no singularities in the flow between C and |z| = R, the integral must be
independent of any chosen closed path outside C:
I 2 I 2
dw dw
dz = lim dz.
C dz R→∞ |z|=R dz
So I 2 I 2 I
dw iΓ iΓU 1
dz = U− ,= − dz = 2U Γ,
C dz |z|=R 2πz π |z|=R z
using the same reasoning as for the cylinder. This proves the theorem.
We have developed a wonderful theory for lift. The only problem is that we don’t
know how to determine the value of Γ which determines the lift. Indeed, the wings of an
airplane do not have circular cross section: if cylinder is placed in a uniform stream (left
side of the Figure), there is no reason why the flow should turn either way, and thus there
is no circulation. The secret is to fashion the shape so as to induce lift. Namely, the tail
of wing is shaped to have a sharp corner, so as to force the flow to separate at that point.
As the simplest possible model for such a situation consider the flow around a flat
plate, which is placed in a uniform stream at an angle α. A flat plate is obtained from
an ellipse by letting d → 0. According to (99), this amounts to putting b = R in the
Joukowski transformation (98):
R2
z=ζ+ . (105)
ζ
Then the width of the plate is 2c = 4R. As is seen from the Figure, the flow separates
from the plate beyond the trailing edge, so that fluid particles very close to the plate have
to go around the sharp corner before leaving the plate.
This situation is reflected by the velocity on the surface of the plate, which is
U sin(θ − α)
u − iv = ,
sin θ
putting b = R in (100). Clearly, u diverges as as θ → 0, π, i.e. at the sharp edges of the
plate. This is a physically untenable situation; fluid particles are not likely to make such
a sharp turn without leaving the plate altogether, especially not at infinite speed! The
same conclusion can be drawn from the concept of “adverse pressure gradient”, introduced
earlier. The speed is very great at the trailing edge and decreases as one moves up the
plate. According to Bernoulli, this means that the pressure increases, in other words fluid
particles experience an adverse pressure gradient. Instead of following a path of adverse
pressure, fluid particles prefer to leave the body, producing a point of separation at the
trailing edge. This requirement is known as the Kutta condition.
Figure 54: Flow around a plate satisfying the Kutta condition (107).
As we have seen in our calculation of the flow around a cylinder with circulation, the
point of separation can be made to move by adding circulation:
R2 iα
−iα iΓ
w1 (ζ) = U ζe + e − log ζ.
ζ 2π
Then the complex velocity of the flow around a plate (b = R) is
−1
R2 iα R2
dw1 dζ −iα iΓ
u − iv = = U e − 2e − 1− 2 .
dζ dz ζ 2πζ ζ
Putting ζ = Reiθ , we find the velocity on the plate:
e−iθ
−iα+iθ iα−iθ
iΓ Γ 1
u − iv = U e −e − = U sin(θ − α) − . (106)
2πR 1 − e−2iθ 4πR sin θ
Let us focus on the trailing edge, corresponding to θ → 0. (assume for example that the
front of the plate is slightly ‘rounded’, so that separation is not as important there). Thus
we want u − iv to be finite for θ → 0, a requirement which is another incarnation of the
Kutta condition. This can be achieved by choosing Γ such that the square bracket on the
right of (106) vanishes for θ = 0. In other words,
Γ = −4πRU sin α, (107)
which is the Kutta condition for the plate. This is the value chosen for the above Figure,
where the flow is seen to leave the plate smoothly in the direction of the orientation of
the plate. Now the singularity cancels and the velocity becomes
U
u − iv = (sin(θ − α) + sin α) → U cos α, as θ → 0,
sin θ
a finite value!
Figure 55: The experimentally measured lift coefficient for a plate, compared to the
theoretical result (110).
The wonderful thing is that we have now determined the circulation uniquely, so we
can calculate the lift using Blasius’ theorem:
Flif t = −U ρΓ = 4πRρU 2 sin α, (108)
where 4R is the width of the plate. Once more, note that this is the force acting normal
to the flow, so it is at an angle α relative to the orientation of the plate. It is customary
to define a lift coefficient cL by
ρU 2
Flif t = cL Aw , (109)
2
where Aw = 4R is the area of the wing (per unit length), also called the chord. Thus we
have found that for the plate
(plate)
CL = 2π sin α ≈ 2πα (110)
In Fig.55, (110) is compared to experiment, and it works really well, as long as the angle
of attack α is small. If however α becomes too large the theory fails abruptly. The reason
is clear from the two photographs below. As long as α is small, the flow remains laminar
and attached to the wing. As α is too great, the flow separates and a completely different
type of description must be sought, see Fig. 56.
Figure 56: The flow around a wing. At small angles α, the flow separates smoothly. As
α becomes large, the flow separates from the wing and vortices are formed.
Finally, we comment on the presence of circulation around the wing, which is the
crucial ingredient needed for flying. In the first instance, there is nothing wrong with that
from the point of view of a potential flow description. However, where is this circulation
coming from when one imagines starting up the flow, with zero circulation? Since the
net circulation in a large circle around the wing must vanish initially, and the Kutta
condition requires a circulation Γ < 0 around the wing, this means that in the process
of the point of separation moving to the trailing edge, vortices of positive circulation
(rotating counterclockwise) are shed from the wing. Eventually the vortices are convected
downstream, and no longer matter for the problem.
5 Waves and free surface flows
Free surface flows are in some ways very different from what we have done before. In
all problems considered so far, the domain D in which to solve the problem is given (for
example some box or the exterior of an aeroplane wing). A free surface, on the other
hand, moves, so the domain varies in time. The key feature, however, is that the domain
does not vary according to some predetermined program. Instead, it moves in response
to the flow itself, that is in response to the flow solution (which of course itself depends
on the domain).
This leads to solutions which are quite different in character to the fixed-boundary solu-
tions we were considering so far (and generally speaking much more interesting solutions)!
Another nice feature of free surface flows is that they are easy to observe experimentally,
one just needs to track the motion of the free surface.
Density = ρ
z=−h
Figure 57: Schematics for wave motion on a fluid layer of thickness h above a solid wall.
u = ∇φ, ∇2 φ = 0,
where u = (u, 0, w) and φ = φ(x, z, t). We choose z = 0 to coincide with the undisturbed
free surface, and the bottom of the fluid is at z = −h, see Fig. 57. Let the surface of the
water in motion be given by z = ζ(x, t).
The interesting part of the problem are its boundary conditions:
(i) At the lower boundary z = −h, the vertical velocity must vanish (kineamtic b.c.):
∂φ
0 = n · ∇φ = = w.
∂z
In other words,
∂φ
=0 on z = −h. (111)
∂z
DS
(ii) On z = ζ(x, t), the kinematic boundary condition on a moving surface is = 0,
Dt
where the free surface is the zero set of S(r, t). If the surface can be represented by
a height function ζ(x, t) (i.e. we are not allowed overhangs) we can define a function
S(x, z, t) = z − ζ(x, t)
which is indeed zero at the free surface. Thus
∂ ∂ ∂ ∂ζ ∂φ ∂ζ ∂φ
0= +u +w (z − ζ(x, t)) = − − + on z = ζ,
∂t ∂x ∂z ∂t ∂x ∂x ∂z
so the kinematic boundary condition becomes
∂ζ ∂φ ∂ζ ∂φ
+ = , on z = ζ. (112)
∂t ∂x ∂x ∂z
(iii) The dynamic boundary condition at the free surface is that the pressure equals the
exterior athmospheric pressure: p = patm (const). on z = ζ. The unsteady Bernoulli
equation (62) for a constant force of gravity Φ = gz is
∂φ
p/ρ + 21 |∇φ|2 + + gz = C(t),
∂t
and so the third boundary condition becomes
∂φ
patm /ρ + 21 |∇φ|2 + + gζ = C(t), on z = ζ. (113)
∂t
This concludes our description of the fully nonlinear problem.
The extra terms are time dependent only, so they do not affect the flow velocities,
Laplace’s equation or the boundary conditions. So now (dropping primes) (iii) is
replaced with
∂φ
(iii) On z = 0, + gζ = 0.
∂t
Now we look for particular solutions to ∇2 φ = 0 with (i), (ii), (iii) motivated by observa-
tions, which describe a propagating wave.
Such a propagating wave of amplitude H is described by a solution of the form
where an arbitrary phase factor δ could have been added as well. The solution (115) is
periodic both in time (with period τ = 2π/ω) and space (with wavelength λ = 2π/k).
Most importantly, it moves to the right with speed c = ω/k, where c is called the phase
speed. To see that, let ξ = x − ct, so that
ζ = H cos(kξ)
1.5
0.5
0
-3 -2 -1 0 1 kh 2 3
u = kA cos k(x+ξ −ct) cosh k(z +η+h), w = kA sin k(x+ξ −ct) sinh k(z +η+h). (121)
Now for small ξ and η, we can put ξ = η = 0 in (121) to obtain the equations of
motion
dξ dη
= kA cos k(x − ct) cosh k(z + h), = kA sin k(x − ct) sinh k(z + h),
dt dt
which we can integrate to yield the trajectory
A A
ξ=− sin k(x − ct) cosh k(z + h), η= cos k(x − ct) sinh k(z + h). (122)
c c
Thus we find that as long as the amplitude A of the wave is small, the motion (122) of
the particles is small as well. Hence we were justified in putting ξ = η = 0 as a first
approximation. In Section 5.4 below we will pursue the expansion to the next (quadratic)
order.
Taking the square of each of the components of (122), we find
ξ2 η2 A2
+ = , (123)
cosh2 k(z + h) sinh2 k(z + h) c2
which is the equation for an ellipse with major and minor semi-axes (in the x and z
directions)
A A
a = cosh k(z + h), a = sinh k(z + h).
c c
Using the expression (117) for A and the dispersion relation (119), A/c = H/ sinh kh, and
so
cosh k(z + h) sinh k(z + h)
a=H , b=H , (124)
sinh kh sinh kh
while the ratio between the two axes is
b
= tanh k(z + h). (125)
a
This means that if the point of observation is far from the bottom (k(z + h) 1),
a ≈ b and the trajectory is a circle. This is clearly visible in Fig. 59, where particle
trajectories are visible as bright lines. The trajectories near the free surface are close to
being circles, while closer to the lower boundary (where z + h becomes small), trajectories
are flattened in the vertical direction. The size of the ellipse decreases as well.
A A kz
ξ = − ekz sin k(x − ct), η= e cos k(x − ct),
k k
this results in
dξ 2 dη 3
= Acekz cos k(x − ct) + A ce2kz , = Acekz sin k(x − ct) + O(A ).
dt dt
This is of course solved to give
A 2
ξ = − ekz sin k(x − ct) + A ce2kz t, η = Aekz cos k(x − ct). (127)
k
To leading order, the trajectories are circles, and no global transport takes place. The
only, but crucial difference is that at quadratic order, particles are pushed forward an
extra horizontal distance, so the circles do not close. On average, particles are pushed
along with the wave with a drift velocity of
2
US = A ce2kz , (128)
z=−h
x=0 x=L
Figure 60: Standing waves in a container
As a model for such situations, consider a two-dimensional rectangular box with rigid
walls at x = 0, L and a bottom on z = −h, filled with fluid to z = 0, see Fig. 60. We
use the small-amplitude theory from before, so that linearised equations are ∇2 φ = 0 for
(x, y, z) in box,
∂φ
(i) = 0 on z = −h.
∂z
∂ζ ∂φ
(ii) Kinematic: = on z = 0.
∂t ∂z
∂φ
(iii) Dynamic: = −gζ on z = 0.
∂t
Also need no-flow conditions (kinematic) on walls:
∂φ
(iv) = 0 on x = 0, L;
∂x
However now we no longer have travelling waves, but are interested in describing a
phenomenon that is periodic in time, so we write in two dimensions:
which implies
X 00 (x)Z(z) + Z 00 (z)X(x) = 0
This separates:
X 00 (x) Z 00 (z)
=− = −k 2
X(x) Z(z)
where −k 2 is the separation constant.
Solving for Z(z), with (i) gives Z(z) = A cosh k(z + h) as before. Combining (ii) and
(iii) to eliminate ζ (that’s g(ii) +∂/∂t(iii)) gives
∂ 2φ ∂φ
= −g , on z = 0
∂t2 ∂z
which implies
−ω 2 X(x)Z(0) cos ωt = −gX(x)Z 0 (0) cos ωt
and it follows that ω 2 /g = k tanh kh as before. (We expect this, as the vertical structure
of the fluid is independent of the vertical lateral walls)
ω12 L
T1 [s] h [cm] g
π tanh πh
L
4.15 0.7 0.17 0.093
2.7 1.4 0.41 0.19
2.1 2.3 0.68 0.31
The Table gives experimental data for measurements in a rectangular tank (see pic-
ture), whose length was L = 74 cm. One end of the tank was lifted, to excite the fun-
damental sloshing mode. The third and fourth columns compare experiment and theory,
as
ω12 L πh
= π tanh ,
g L
and T1 = 2π/ω1 . Remarkably, the period decreases with increasing depth h. The trend is
reproduced correctly, but there is a consistent discrepancy by a factor of two, perhaps due
to the fact that the tank is too long, so we didn’t really excite the findamental mode...
Appendix A: Vector calculus
We shall revise some vectors operations that you should have already met before this
course. We will be using suffix notation throughout this course
(drop the summation symbol on the understanding that repeated suffices imply summa-
tion.
Defn A.1.2: The Kronecker delta is defined by
1, i=j
δij =
0, i 6= j
Examples:
1. δii = 3
2. δij ui vj = uj vj ≡ u · v.
Defn A.1.3: The antisymmetric symbol ijk is defined by
• 123 = 1
• Interchanging any two suffices reverses the sign: e.g. ijk = −jik = −kji
• Above implies invariant under cyclic rotation of suffices: ijk = jki = kij
With this definition all 27 permutations are defined. There are only 6 non-zero compo-
nents,
123 = 231 = 312 = 1, 213 = 132 = 321 = −1,
Defn A.1.4: The cross product is defined by
wi = ijk uj vk
where summation over j and k occurs. [Check].
w · (u × v) = wi ijk uj vk
w · (u × v) = jki wi uj vk = u · (v × w)
= kij wi uj vk = v · (w × u)
= −jik wi uj vk = −u · (w × v)
= −ikj wi uj vk = −w · (v × u)
[Check]
w × (u × v) = (w · v)u − (w · u)v
Proof:
2. ∇ · r = ∂i xi = δii = 3
4. ∇r = rr , where r2 ≡ r · r.
√ ∂i x2j xj ∂ i xj xj δij xi
Proof: [∇r]i = ∂i xj xj = q = = = .
2 x2j r r r
3. f × (∇ × f ) = ∇( 12 f · f ) − (f · ∇)f
Proof:
ds
V S
Very important. Consider a volume V bounded by a closed surface S with outward unit
normal n. Then for a vector field f (x)
Z Z
∇ · f dV = f · ndS
V S
Corollary: Let f (r) = aφ(r) where a is an arbitrary constant vector, and φ(r) a scalar
function. Since ∇ · (aφ(r)) = a · ∇φ(r), so the divergence theorem reduces to
Z Z
a· ∇φdV = a · φndS
V S
ds
S
C
dl
and from §A.2.1(2), (∇ × aφ) = φ(∇ × a) + (∇φ × a) = (∇φ × a). Using the vector
triple product result, (∇φ × a) · n = −a · (∇φ × n) and then
I Z
a· φdl = −a · ∇φ × ndS
C S
Therefore I Z
φdl = − ∇φ × ndS
C S
Appendix B: Curvilinear coordinate systems
Many problems can be approached more simply by choosing a coordinate system that
fits a given geometry. Instead of writing the position vector r as a function of Cartesian
coordinates (x, y, z), r is now written as function of three new coordinates: r(q1 , q2 , q3 ).
The coordinate lines are swept out by varying one of the coordinates, keeping the other two
constant. We will deal only with the by far most important case of orthogonal coordinate
systems, in which the coordinate lines always intersect one another at right angles.
∂r
Evidently, is a vector which points in the direction of the i-th coordinate line,
∂qi
see the figure. If each of these vectors are normalized to unity, we obtain the local basis
system:
1 ∂r ∂r
q̂i = , hi ≡ . (129)
hi ∂qi ∂qi
The quantities hi (q1 , q2 , q3 ) are called scale factors or metric coefficients. The fact that
the coordinate system is orthogonal means that all q̂i , computed at a point (q1 , q2 , q3 ),
are orthogonal. However, the direction of q̂i of course changes as one goes along the
coordinate lines.
Remark: As the coordinates are varied by δq1 , δq2 , and δq3 , respectively, r describes a
volume whose sides are orthogonal. The length of each side is hi δqi , and thus the volume
of the cuboid is d3 x = h1 h2 h3 dq1 dq2 dq3 . Thus integration in a curvilinear coordinate
system is achieved by the formula
Z Z
3
f (r)d x = f (q1 , q2 , q3 )h1 h2 h3 dq1 dq2 dq3 . (130)
V V
In cylindrical polars, the volume element is rdrdθdz.
To do vector calculus in the curvilinear coordinate system, we have to work out what
the ∇-operator is. For any scalar function f (r) (and suspending the summation conven-
tion for the moment), we have
3
∂f X ∂xj ∂f ∂r
= = · ∇ f = hi (q̂i · ∇) f
∂qi j=1
∂q i ∂x j ∂q i
by the chain rule, and using (129). But up to the factor hi , the object on the right-
hand side is the qi component of ∇f . This means that in the local basis, ∇ has the
representation
q̂1 ∂ q̂2 ∂ q̂3 ∂
∇= + + . (131)
h1 ∂q1 h2 ∂q2 h3 ∂q3
For example in cylindrical polars, the gradient of a scalar function φ is
∂φ 1 ∂φ ∂φ
∇φ = r̂ + θ̂ + ẑ,
∂r r ∂θ ∂z
in agreement with chapter 0.
Now we are almost there. The only but crucial thing that remains to be considered
is the fact that both scale factors and unit vectors depend on the coordinates qi , so they
∂ q̂i
need to be differentiated. In particular, we want to represent in terms of the basis
∂qj
vectors. It is not difficult to write down general expressions for the derivatives of the basis
vectors if only the hi are known, see exercises. However, for a given coordinate system,
it is generally much easier to simply calculate the derivatives directly. For example, in
cylindrical polars
∂r̂ ∂ θ̂
= θ̂, = −r̂,
∂θ ∂θ
and all the other derivatives are zero.
Now it is easy to derive the other formulae in chapter 0. In cylindrical polars, the
Laplacian becomes
! !
∂ θ̂ ∂ ∂ ∂φ θ̂ ∂φ ∂φ
4φ = ∇ · ∇φ = r̂ + + ẑ r̂ + + ẑ =
∂r r ∂θ ∂z ∂r r ∂θ ∂z
!
∂ 2φ 1 1 ∂ 2φ ∂ 2φ
∂r̂ ∂φ 1 ∂ θ̂ ∂φ
(r̂ · r̂) 2 + θ̂ · + 2 θ̂ · + θ̂ · θ̂ 2 2 + (ẑ · ẑ) 2 =
∂r r ∂θ ∂r r ∂θ ∂θ r ∂θ ∂z
∂ 2 φ 1 ∂φ 1 ∂ 2φ ∂ 2φ
+ + + 2.
∂r2 r ∂r r2 ∂θ2 ∂z
u = ∇ × A.
Conversely, it is clear from this representation that the flow is incompressible. The rep-
resentation is particularly useful in two dimensions, in which case A can be written in
terms of a single scalar function, the streamfunction. By definition, the streamfunction
is constant along streamlines.
(i) Cartesians: Consider Cartesian coordinates (x, y, z), and take the flow in the
(x, y)-plane. Then A = ψ(x, y, t)ẑ, and
∂ψ ∂ψ
u= , v=− .
∂y ∂x
Now confirm that ψ is indeed constant along streamlines, which are defined by
dr
= λu,
ds
and thus dx = λuds and dy = λvds. It follows that
∂ψ ∂ψ
dψ = dx + dy = vdx − udy = 0,
∂x ∂y
if indeed ψ is varied along streamlines. Thus ψ is constant along a streamlines, as adver-
tised.
Defn: We call the function ψ(x, y, t) the streamfunction of the flow.
ψ= constant
(ii) Cylindrical polar coordinates: Now consider flow in the same two dimensional
plane z = 0, which is represented in cylindrical polars (r, θ, z). The relation to Cartesian
coordinates is x = r cos θ, y = r sin θ. Since the stream function is defined completely by
the geometry of the streamlines, it must be the same in any coordinate system. In other
words if ψc (x, y, t) is the Cartesian version, and ψp (x, y, t) the streamfunction in polar
coordinates, then ψp (r, θ, t) = ψc (r cos θ, r sin θ, t).
As before, the vector potential is A = ψ(r, θ, t)ẑ, and thus from the curl in polar
coordinates
1 ∂ψ ∂ψ
ur = , uθ = − ,
r ∂θ ∂r
where u = ur r̂ + uθ θ̂.
Now verify that ψ(r, θ) is indeed constant along streamlines. In polar coordinates,
r = rr̂, and thus
dr dr dθ dr̂ dr dθ
= r̂ + r = r̂ + r θ̂,
ds ds ds dθ ds ds
dr̂
where we have used that = θ̂. In other words, we find that dr = λur ds and rdθ =
dθ
λuθ ds. Now as before,
∂ψ ∂ψ
dψ = dr + dθ = −uθ dr + rur dθ = 0,
∂r ∂θ
if ψ is indeed varied along streamlines.
Appendix D Some simple flows and their potentials
D.1 Two dimensional flows
We will give the answer either in Cartesians or polars, depending on which is more con-
venient.
Reminder:
∂φ ∂ψ ∂φ ∂ψ
(i) in Cartesians, u = (u, v, 0) and u = = , v= =− where φ is the
∂x ∂y ∂y ∂x
velocity potential and ψ is the streamfunction.
∂φ 1 ∂ψ 1 ∂φ ∂ψ
(ii) In polars, u = (ur , uθ , 0) and ur = = , uθ = =− .
∂r r ∂θ r ∂θ ∂r
Type of flow flow field u potential φ streamfunction ψ complex pot. w
Uniform stream par- Ur̂ Ux Uy Uz
allel to x axis
U cos αr̂+
Uniform stream at U r cos(θ − α) U r sin(θ − α) U ze−iα
U sin αŷ
angle α to x axis
u = kx k 2 k 2
Stagnation point flow (x − y 2 ) kxy z
v = −ky 2 2
at origin
mr̂ m mθ m
Line source, strength log r ln z
2πr 2π 2π 2π
m, at r = 0
Γθ̂ Γθ Γ iΓ
Line vortex, circula- − log r − ln z
2πr 2π 2π 2π
tion Γ, at r = 0
µ z cos θ sin θ µ
Horizontal dipole, − ẑ − 2 r̂ −µ µ −
2πr2 r 2πr 2πr 2πz
strength µ, at r = 0
∂φ 1 ∂Ψ ∂φ 1 ∂φ
ur = =− , uz = = .
∂r r ∂z ∂z r ∂r
In spherical polar coordinates, (r, θ, ϕ), u = (ur , uθ , uϕ ); in terms of the potential φ(r, θ):
∂φ 1 ∂Ψ 1 ∂φ 1 ∂Ψ
ur = = 2 , uθ = =− .
∂r r sin θ ∂θ r ∂θ r sin θ ∂r
Type of flow flow field u potential φ coordinate system
Uniform stream U ẑ Uz cylindrical
aligned with axis
of symmetry
k
ur = r k 2
Stagnation point 2 (r − 2z 2 ) cylindrical
uz = −kz 4
flow at origin
m m
Point source, r̂ − spherical
4πr2 4πr
strength m, at
r=0
µ z cos θ
Dipole, strength µ, − ẑ − 3 r̂ −µ spherical
4πr3 r 4πr2
in ẑ-direction