Fluid Dynamics
Fluid Dynamics
1
Contents
1 Fluid Mechanics 12
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.1.2 Continuum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.2 Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.2.1 Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.6.3 Simplifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.7.1 Pathlines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.7.2 Streamlines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2
1.8.1 Streamfunction in 2D . . . . . . . . . . . . . . . . . . . . . . . . . . 26
1.9.2 Circulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.4 Doublets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.5.2 Source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3
3.5.3 Doublet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.2.1 Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.2.2 Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.4 Alternative Form of the Navier-Stokes Equation and the Vorticity Equation 58
5 Streamfunction in 2-Dimensions 61
5.1 Cartesians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4
6.3.1 Uniform Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.4 Singularities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
6.4.2 Vortex . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
7 Boundary Conditions 91
5
9.2.1 2-D Poiseuille Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
6
11.3.4 Stress tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
11.3.6 Calculating the force on the sphere by doing an integral at infinity . 140
A Glossary 142
7
C.4 The Curl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
8
List of Figures
1.1 Rotation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.5 Rotation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.7 A pthline. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.8 Streamlines. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.10 Stress. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
1.11 Say two points on a curve have tangent vector dx~ex + dy~ey then dy~ex −dx~ey
points ‘outward’ - this can be seen here following from the fact that dx and
dy are positive. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.2 Stress. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
9
6.4 Flow at a stagnation point. . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
6.6 Extended solution from flow at a corner with walls at θ = 0 and θ = 2π/n. 74
6.12 Doublet. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6.17 Flow around a rotating cylinder results in an upward force on the cylinder. 89
B.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
B.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
B.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
B.5 When adding together the circulation of each loop the only remaing con-
tribution to the line integral comes from the outside edge. . . . . . . . . . 155
B.6 We have some surface bounded by the loop Γ. The surface is divided into
many small areas, each approximately a square. . . . . . . . . . . . . . . . 156
10
C.1 The gradient of ~v . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
11
Chapter 1
Fluid Mechanics
1.1 Introduction
The defining property of a fluidis that it cannot withstand, shearing forces, however small,
without sustained motion. Since both gases and liquids have this property, both are fluids
and subject to a unified treatment as far as their macroscopic motion is concerned.
The rate at which the fluid deforms continuously depends not only on the magnitude of
the applied force but also on a property of the fluid called its viscosity - its resistance to
deformation and flow.
Solids also deform when sheared, but a position of equilibrium is soon reached in which
elestic forces induced by the deformation of the solid exactly counterbalance the applied
shear force, and further deformation ceases.
1.1.2 Continuum
There are a large number of atoms or molecules in any given volume. For example the
number of gas molecules in (1µcm)3 is 2.687 × 107 . Therefore gases and liquids can be
considered as continuum, for example water appears continuous, wind appears continuous
(even though we cant see it).
What are the physical properties of a fluid? Density, viscocity and temperature and
pressure, velocity and acceration.
12
Density ̺ = ̺(~x, t)
Voscosity µ = µ(~x, t)
Temperature T = T (~x, t)
Voscosity p = p(~x, t)
Temperature ~u = ~u(~x, t). (1.1)
1.2 Kinematics
1.2.1 Velocity
d~x
~u = . (1.2)
dt
Lagrangian
(a) In the Lagrangian description we follow a particle which starts at position ~x = ~x0 at
time t = t0 and we follow it and observe its velocity ~q(~x0 ; t, t0 ). In this case ~x and t are
dependent variables.
Newton’s laws and conservation of mass and energy, apply directly to each particle. How-
ever, fluid flow is a continuum phenomenon and it is not possibly to to track each “par-
ticle”. Let’s consider a second way of describing fluid motion.
Eulerian
(b) In the Eulerian description we define the velocity of the fluid particle which is situated
at ~x at time t to be ~u(~x, t).
13
with ~x and t as independent variables.
Here one is not concerned about the location or velocity of any particular particle, but
rather about the velocity of whatever particle happens to be at a particular location of
interest at a particular time.
∂~q
~a = (~x0 ; t, t0 ) = . (1.4)
∂t
(b) Eulerian.
The velocity field ~u gives the velocity of the fluid at someplace in the fluid. The accelera-
tion is not simply given by ∂~u/∂t as this represents the rate at which the velocity ~u(x, y, z)
changes at a fixed point in space. What we need is the rate at which the velocity changes
for a particular piece of fluid. After time dt the piece of fluid will have moved to a new
position. The particle at position (~x, t) moves to (~x + d~x, t + dt) where d~x = ~u(~x, t)dt
∂~u ∂~u
~u(~x + d~x, t + dt) = ~u(~x, t) + dt + dx (1.6)
∂t ∂xi i
so that
∂~u
~a(~x, t) = + ~|u ·{z
∇~u} . (1.8)
∂t
|{z} advective
local change
14
D ∂
= + ~u · ∇. (1.9)
Dt ∂t
Consider an infintesimal volume element of the fluid at time t0 . After an infintesimal time
later δt this volume element will have suffered a deformation that can be considered the
combination of a translation, linear expansion, rotation and shear (see fig 1.1).
= + + +
A translation is when the position of a fluid element is changed without changing its size
or shape, see fig 2. After a time δt the fluid element will be shifted in the x−direction an
amount ux δt and shifted in the y−direction an amount uy δt - see fig 1.2.
δy
O′
uy ux uy δt
O δx
ux δt
Expansion
Let us consider the case where we have linear expansion in the x−direction. At time t0
the x−component of the velocity at position (x + δx, y) where δx is small is given by the
15
first order Taylor expansion
∂u
u(x + δx, y, z, t0 ) = u(x, y, z, t0) + δx.
∂x
δy δy
∂ux
ux ux + ∂y
δy
O δx A O δx A A′
∂ux
δl = ∂x
dx δt
1 dδV ∂u
= x.
δV dt ∂x
If velocity gradients ∂uy /∂y and ∂uz /∂z are also present, it can easily be shown that
16
Rotation and shear
We now consider the effects of rotation and shear. For simplicity we consider motion in
the x − y plane, however these considerations are easily extended to the more general
case. The situation is the one shown in fig 1.4. We have that the change in angle of the
line OA is
∂uy
δα = δx δt
∂x
∂ux
δβ = δy δt.
∂y
∂ux
∂y
δy δt
∂ux
ux + ∂y
δy
B B B′
uy
δβ
δy δy
δα
uy ux ux ∂uy A′ ∂uy
uy + ∂x
δx ∂x
δx δt
A
O δx A O δx
Note that if ∂uy /∂x is positive, ωOA will be counterclockwise. If ∂ux /∂y is positive then
ωOB will be clockwise.
Rotation
We consider pure rotations first where after time δt the element has rotated by a small
angle θ in which case
∂uy
δxδt ∂uy
ϑ = δα = ∂x = δt
δx ∂x
∂ux
δxδt
∂y ∂u
= −δβ = − = − x δt. (1.11)
δx ∂y
17
Rotations are obviously volume preserving.
Shear
Shear type strains are where the initial fluid element is changed into a parallelogram.
∂uy
ϑ = δα = δt
∂x
∂u
= δβ = x δt. (1.12)
∂y
The area of the parallelogram is |OA||OB| sin(π − 2δα) or |OA||OB| cos(2δα). Taylor
expanding we get
1 2
|OA||OB| 1 − (2δα) + . . .
2!
When we have general ‘angular’ deformations we extract the part that corresponds to
rotation by taking the average of the angular velocities with counterclockwise rotation
considered positive,
1 δα δβ
ω = −
2 δt δt
1 ∂u y ∂ux
= − (1.13)
2 ∂x ∂y
1 δα δβ
σ = +
2 δt δt
1 ∂uy ∂ux
= + (1.14)
2 ∂x ∂y
For pure rotation (1.14) is zero, but for pure shear we have δα = δβ and (1.13) is zero.
18
1.4 Rate of Strain Tensor
∂ui
ui (~x + ~r) = ui (~x) + rj + O(r 2 ) (1.15)
∂xj
!
1 ∂ui ∂uj
eij = + (1.16)
2 ∂xj ∂xi
ǫxx ǫxy ǫxz
ǫyx ǫyy ǫyz (1.17)
ǫzx ǫzy ǫzz
Define
!
1 ∂ui ∂uj
ωij = − = −ωji . (1.18)
2 ∂xj ∂xi
This has three independent components. Behave like a vector. Only happens in three
dimensions.
∂uk
ωi = ǫijk (1.19)
∂xj
We can now see clearly the meaning of the term ωij rj in the Taylor series expansion for
the displacements
1
ωij rj = ǫijk ωj rk (1.21)
2
19
1
~ω × d~r. (1.22)
2
1 1
~ × d~r = ωr~j.
ω
2 2
z ~ω
1
ui (~x + ~r) = ui (~x) + eij rj + ǫijk ωj rk + O(r 2 ) (1.23)
2
! !
∂ui 1 ∂ui ∂uj 1 ∂ui ∂uj
= + + −
∂xj 2 ∂xj ∂xi 2 ∂xj ∂xi
= eij + ωij
1 1 1
= ekk δij + eij − ekk δij + ǫijk ωk
3 3 2
1 1
= θδij + γij + ǫijk ωk (1.24)
3 2
X ∂u
i
=θ
i
∂xi
20
P
where we have used i γii = 0. The term
1
θδ
3 ij
in (1.24) represents the isotropic expansion part, that is it is the volume change ‘averaged’
in each direct equally. Also from (1.24) we have
!
∂ui ∂uj
− = ǫijk ωk
∂xj ∂xi
If i 6= j then
∂ui ∂uj
+ .
∂xj ∂xi
1
γii = eii − θ (no summation implied)
3
as eii (no summation implied) is the volume change in the ith−direction and is minus the
ith part of isotropic expansion.
Z
̺dV
V
and conservation of mass means the rate of change of mass equals the mass flux across
the boundary
Z Z
∂
̺dV = − ̺~u · ~ndS (1.25)
∂t V S
21
The minus sign is there as mass flux leaing the volume would reduce the mass. Using the
Divergence Theorem
Z
∂̺
+ ∇ · (̺~u) dV = 0 (1.26)
V ∂t
∂̺
+ ∇ · (̺~u) = 0 (1.27)
∂t
This is the differential form of the continuity equation.
∂(̺ux )
̺ux dydz [̺ux + ∂x
dx]dydz
∂
xf lux = ̺ux + (̺ux )dx dydz − ̺ux dydz
∂x
∂
= (̺ux )dxdydz.
∂x
∂
yf lux = (̺uy )dxdydz
∂y
∂
zf lux = (̺uz )dxdydz (1.28)
∂z
22
The total net mass outflux must balance the rate of decrease of mass within the cubical
element which is
∂̺
− dxdydz
∂t
Combining the above equations, the balance of mass outflux with mass decrease is then
described by the equation
∂̺ ∂ ∂ ∂
+ (̺ux ) + (̺uy ) + (̺uz ) dxdydz = 0.
∂t ∂x ∂y ∂z
Or
∂̺ ∂ ∂ ∂
+ (̺ux ) + (̺uy ) + (̺uz ) = 0.
∂t ∂x ∂y ∂z
1.6.3 Simplifications
∂̺
+ ̺∇ · ~u + ~u · ∇̺ = 0 (1.29)
∂t
D̺
+ ̺∇ · ~u = 0
Dt
A simplification would be steady flow in which case the continuity equation becomes
∇ · (̺~u) = 0.
If the density is constant ̺ = ̺0 , that is, the fluid is incompressible the continuity equation
becomes
∇ · ~u = 0 (1.30)
23
~ such that
So ~u divergence free and hence solenoidal, that is, there is a vector A
~
~u = ∇ × A.
We exclusively deal with incompressible flows in these notes. Note that this equation can
be expressed in terms of the rate of strain tensor as ekk = 0.
1.7.1 Pathlines
~
The locus of an actual particle X(t) is called a pathline. Pathlines are associated with
the Lagrangian description. The corresponding acceleration of the particle is then
dX~
= ~q = ~u(~x, t) . (1.31)
dt |{z} | {z }
Lagrangian Eulerian
X(t)
1.7.2 Streamlines
We can draw lines which are always tangent to the fluid velocity at an instant in time,
these are called streamlines. With steady flow (this is when the velocity of fluid doesn’t
change at any point with time) the streamlines dont change with time and in this case the
particle’s path coincides with the streamline. If we have unsteady flow the streamlines
vary from instant to instant and in general the particle’s path does not coindide with a
streamline as the particle can end up on another streamline. That is with time unsteady
flow pathlines and streamlines are not necessarily coincident.
24
Figure 1.8: Streamlines.
d~x
= λ~u(~x, t) (1.32)
ds
or
dx dy dz
~x + ~y + ~z = λ(ux~x + uy ~y + uz ~z) (1.33)
ds ds ds
where s is a parameter along the streamline. The parameter s should not be confused
with time, the above equations are integrated while keeping time fixed. The resulting
curves give us the streamlines at an instant in time.
The streamline passing through a closed curve which does not lie on a surface generated
by streamlines form a tubular surface. The fluid contained in such surface is called a
stream tube.
∇ · ~u = 0 (1.34)
25
Here we breifly introduce the method using the so-called streamfucntion. We will go into
this in more detail in future chapters. The streamfunction method is introduced for 2D
motion. It was extended to axisymmetric flow in thre dimensions - then called Stoke’s
streamfucntion.
1.8.1 Streamfunction in 2D
Imcompressibility condition
∇ · ~u = 0. (1.35)
∂u ∂v
+ = 0. (1.36)
∂x ∂y
If we choose
∂ψ ∂ψ
u= , v=− (1.37)
∂y ∂x
∂ ∂ψ ∂ ∂ψ
− = 0.
∂y ∂y ∂x ∂y
We now prove that ψ = Const. is a streamline in steady flow. First note tat in general for
two infinitesimally close points a = (x, y) and b = (x + dx, y + dy) we have from calculus
that
∂ψ ∂ψ
ψ(x + dx, y + dy) − ψ(x, y) = dx + dy
∂x ∂y
= ∇ψ · d~r. (1.38)
26
where d~r is the vector frompoint a to point b. If a and b are points on a curve defined by
ψ = C where C is a constant, then d~r is tangent to the curve ψ = C at a and
Implying that the vector ∇ψ is normal to the curve ψ = C. If we can show that everywhere
~u · ∇ψ = 0, using the formula for ~u in terms of ψ given by (1.37), then we will have proved
the result. This easily follows
∂ψ ∂ψ ∂ψ ∂ψ
~u · ∇ψ = + − = 0. (1.39)
∂y ∂x ∂x ∂y
We have already defined the flux through a surface by summing over ~u · ~ndS where dS is
the area of an infintesimal area element with unit normal vector ~n. The two dimensional
analogue of the volume flux as the sum of terms
~u · ~ndl
along a given curve and where dl is the length of an infintesimal line element.
We first consider the simple case of the volume flux Φ through the curve x = Const. in
the x−direction from say a to b,
Z b Z b
∂ψ
Φ= udy = dy = ψ|b − ψ|a . (1.40)
a a ∂y
x
a
Now consider any curve Γ from point P to point Q. The volume flux through a curve is
the integral of the dot product of the flow velocity vector (u, v) and the normal to the
27
curve. If (dx, dy) is the infinitesimal displacement from one point of the curve to another,
the unit normal ~n to the curve there multiplied by dl is (dy, −dx), i.e.
and
The vector (1.41) points ‘outward’ as can be understood from considering fig (1.8).
y
dx
dy
~ndl
x
Figure 1.11: Say two points on a curve have tangent vector dx~ex + dy~ey then dy~ex − dx~ey
points ‘outward’ - this can be seen here following from the fact that dx and dy are positive.
The volume flux through the curve Γ with end points P and Q is
Z Q
Φ = ~u · ~ndl
P
Z Q
= (u~ex + v~ey ) · (dy~ex − dx~ey )
P
Z Q
= (udy − vdx)
P
Z Q
∂ψ ∂ψ
= dx + dy
P ∂x ∂y
Z Q
= ∇ψ · d~r (1.42)
P
28
where we have used (1.37). By the theorem proved in section B.4 we then have
Note that the flux only depends on the values ψ take at the end points, and so the flux
is independent of which curve we choose between the end points.
Since it is axisymmetric it can be written in terms of two coordinates. There are two
different choices of coordinates systems. This will be much expanded on in future chapters.
~u = ∇φ (1.44)
∇2 φ = 0 (1.45)
A line to which vorticity vectors are tangent at all its points is called a vortex line. The
differential equations for a vortex line are,
d~x
= λ~ω . (1.46)
ds
29
Voretx lines have a direction of ~ω and have a density in any region proportional to the
magnitude of ~ω . As ~ω = ∇ × ~u, we have that ∇ · ~ω = 0. So vortex lines are like lines of a
magnetic force - they will form closed loops.
1.9.2 Circulation
Z B Z B
Γ= ~u · d~x = uidxi . (1.47)
A A
There is a simple relationship between the circulation around a closed curve and the
vorticity of the fluid over any surface bounded by that curve.
Z Z
~u · d~x = ∇ × ~u · dS~
C ZS Z
= ~ω · ~ndS = ωn dS (1.48)
S S
30
Chapter 2
If Fi denotes the component of body forces per unit mass acting in the xi −direction, the
i − th component of the resulting body force acting on V is
Z
̺Fi dV. (2.1)
V
Let Fi denote body forces per unit mass in the i − th direction acting upon a surface
whose normal is in the j − th direction and denoted by σij
An ideal (or perfect) fluid is one which can exert no shearing stress across any surface
31
−p 0 0
σ = 0 −p 0 (2.4)
0 0 −p
Z Z
σij dSj = − pdSi (2.5)
S S
this is done ignoring viscous effects, that is µ = 0 and the fluid is inviscid.
We can go through the same process for momentum as with did for mass. The total
momentum in the volume V is:
Z
Πi = ̺ui dV (2.6)
V
where i runs over the three components of the momentum. We allow the proof to include
compressible fluids however will only look at incompressible solution. It’s rate of change
is just
Z
∂Πi ∂(̺ui )
= dV (2.7)
∂t V ∂t
In the absense of any forces a momentum change within the volume V can occur by
momentum flowing across the boundary,
Z
∂Πi
=− (̺ui)~u · ~ndS (2.8)
∂t S
Using the Divergence theorem the surface integral can be written as a volume integral
Z
∂(̺ui )
+ ∇ · (̺ui~u) dV = 0 (2.9)
V ∂t
∂(̺ui )
+ ∇ · (̺ui~u) = 0. (2.10)
∂t
32
Note the analogy with the mass conservation equation (1.27).
d~p d
F~ = = (m~u).
dt dt
It must be beared in mind that the we write d/dt rather than ∂/∂t, and that the laws of
motion apply to particles, not to the control volume. We have “surface forces on S” plus
“body forces” in V but assume viscous forces are negligible so surface forces are purely
due to pressure. Then we can write for an ideal fluid
Z Z Z Z
∂
̺ui dV + ui ̺uk dSk = − pdSi + ̺Fi dV (2.11)
∂t V S S V
where the second term is the outlet momentum flux and the RHS is the sum of surface
and body forces. Converting again the flow contributions to a volume integral and then
deducing local equations one obtains
∂(̺ui )
+ ∇j (̺ui uj ) = −∇i p + ̺Fi . (2.12)
∂t
∂(̺ui ) ∂u ∂̺
+ ∇ · (̺ui~u) = ̺ i + ui + [ui ∇ · (̺~u) + ̺(~u · ∇)ui]
∂t ∂t ∂t
∂ui ∂̺
= ̺ + (~u · ∇)ui + ui + ∇ · (̺~u) .
∂t ∂t
∂ui Dui
̺ + (~u · ∇)ui ≡ ̺ . (2.13)
∂t Dt
Dui ∂p
̺ + − ̺Fi = 0 (2.14)
Dt ∂xi
33
These are known as the Euler equations of motion.
D~u
̺ = −∇p + ̺F~ (2.15)
Dt
∇ · ~u = 0. (2.16)
p′ = p + ̺Ω
D~u
̺ = −∇p′ . (2.18)
Dt
~u = ∇φ (2.19)
∇2 φ = 0. (2.20)
For potential flow the velocity field can thus be found by solving Laplace equation for
the velocity potential. Since the Laplace equation is linear we can superimpose solutions.
In particular we can take simple flows and superimpose them to form more complicated
flows of interest such as the flow around solid objects. In chapter 3 we find simple exact
34
solutions of the Laplace equation in three dimensions, simple flows, and construct more
complex flows. We do the same for plane flows in chapter 6.
In this chapter we have not included viscosity. An important aspect of viscous fluids is
the non-slip condition at the surface of a solid object. This states that the velocity of the
fluid at a solid boundary is zero relative to the boundary. As we shall see potential flows
often dont satisfy this non-slip condition - however all liquids have a non-zero viscosity.
Can potential flows be of physical interest?
As it turns out for small viscocity (more accurately large Reynolds number, see chapter 8)
the viscous effects are concentrated in a thin boundary layer around the surface. Outside
the bounary layer the fluid acts as an invisid fluid and so potential flow theory provides
both the velocity at the outer edge of the boundary layer and the pressure there (we
explain how to find the pressure in the next section). The pressure is not significantly
effected by the thin boundary layer, so the pressure calculated from the invisid flow gives
the pressure on the surace, and then for example this can be integrated over the body’s
surface to calculate a lift.
Having constructed potential flows, the Euler equation can then be used to find the
pressure. Substitute ~u = ∇φ into the Euler equation (2.18),
D~u ∂~u
̺ =̺ + ~u · ∇~u = −∇p′ . (2.21)
Dt ∂t
!
∂ui ∂u ∂p′
̺ + uj i =− (2.22)
∂t ∂xj ∂xi
∂φ
but ui = ∂xi
!
∂ ∂φ ∂φ ∂ 2 φ ∂p′
̺ + =− (2.23)
∂t ∂xi ∂xj ∂xi ∂xj ∂xi
or
!2
∂ ∂φ 1 ∂φ ∂p′
̺ + =− (2.24)
∂xi ∂t 2 ∂xj ∂xi
35
or
!2
∂ ∂φ 1 ∂φ p′
+ + =0 (2.25)
∂xi ∂t 2 ∂xj ̺
!2
∂φ 1 ∂φ p′
+ + = Const. = p∞ (2.26)
∂t 2 ∂xj ̺
Therefore
∂φ 1
p′ = p0 − ̺ − ̺|~u|2 . (2.27)
|{z} |2 {z }
∂t
(1) (2)
This is the Bernoulli pressure, where (1) is the rate of change of potential with time and
(2) is the dependence quadratically on velocity.
We will prove later that potential flow is not influenced by viscous forces and vorticity.
36
Chapter 3
As mentioned in the previous chapter, for potential flow the velocity field can be found
by solving Laplace equation for the velocity potential. Since the Laplace equation is
linear we can superimpose solutions. We can take simple flows and superimpose them
to form more complicated flows of interest such as the flow around solid objects. The
momentum conservation equation, that is the Euler’s equations of motion, can then be
used to calculate the pressure.
Exact solutions are easily found and described in spherical or cylindrical coordinates,
not just in this chapter but also chapters to come. In the next section we introduce the
mathematics for spherical and cylindrical coordinates with more details and derivations
given in the appendices D and E.
x = r sin θ cos ϕ
y = r sin θ sin ϕ
z = r cos θ (3.1)
37
~er
θ ~eϕ
r ~eθ
ϕ
Figure 3.1:
When we work with vectors in spherical polar coordinates, we use a new local vector
basis different from ~i, ~j and ~k of Cartesian coordinates. Instead we specify vectors as
components in the basis (~er , ~eθ , ~eϕ ) of unit vectors. These basis vectors may be visulalised
as follows. To see the direction of ~er , keep θ and ϕ and increase r. To see the direction
of ~eθ , keep r and ϕ fixed and increase θ. To see the direction of ~eϕ , keep r and θ fixed
and increase ϕ. If ~r corresponds to the position vector then mathematically these unit
vectors are written
∂~r
hi =
∂qi
∂~r
= sin θ cos ϕ~ex + sin θ sin ϕ~ey + cos θ~ez
∂r
∂~r
= r cos θ cos ϕ~ex + r cos θ sin ϕ~ey − r sin θ~ez
∂θ
∂~r
= −r sin θ sin ϕ~ex + r sin θ cos ϕ~ey (3.4)
∂ϕ
38
From these we can check that the basis vectors are indeed orthogonal to each other. We
calculate
to obtain
hr = 1, hθ = r, hϕ = r sin θ (3.6)
We have
1 ∂~r
~ex · ~ei = ~ex ·
hi ∂qi
1 ∂
= ~ex · x~ex + y~ey + z~ez
hi ∂qi
1 ∂x
= .
hi ∂qi
Substutuiting this and analogous results for ~ey and ~ez into (3.7) gives
1 ∂φ ∂x 1 ∂φ ∂y 1 ∂φ ∂z
(∇φ)i = + +
hi ∂x ∂qi hi ∂y ∂qi hi ∂z ∂qi
1 ∂φ
=
hi ∂qi
Therefore we have
39
3
X 1 ∂φ
∇φ = ~ei (~r) . (3.8)
i=1
hi ∂qi
∂φ 1 ∂φ 1 ∂φ
∇φ = ~er + ~eθ + ~eϕ . (3.9)
∂r r ∂θ r sin θ ∂ϕ
∂φ 1 ∂φ 1 ∂φ
ur = , uθ = , uϕ =
∂r r ∂θ r sin θ ∂ϕ
x = ρ cos ϕ
y = ρ sin ϕ
z = z (3.10)
These basis vectors may be visulalised as follows. To see the direction of ~eρ , keep ϕ and z
and increase ρ. To see the direction of ~eϕ , keep ρ and z fixed and increase ϕ. To see the
direction of ~ez , keep ρ and ϕ fixed and increase z. If ~r corresponds to the position vector
then mathematically these unit vectors are written
40
~ez
ρ ~eϕ
z
~eρ
∂~r
= cos ϕ~ex + sin ϕ~ey
∂ρ
∂~r
= −ρ sin ϕ~ex + ρ cos ϕ~ey
∂ϕ
∂~r
= ~ez (3.13)
∂z
From these we can check that the basis vectors are indeed orthogonal to each other. We
calculate
to obtain
hρ = 1, hϕ = ρ, hz = 1 (3.15)
3
X 1 ∂φ
∇φ = ~ei (~r) . (3.16)
i=1
hi ∂qi
41
∂φ 1 ∂φ ∂φ
∇φ = ~eρ + ~eϕ + ~ez . (3.17)
∂ρ ρ ∂ϕ ∂z
∂φ 1 ∂φ ∂φ
uρ = , uϕ = , uz =
∂ρ ρ ∂ϕ ∂z
φ = Ux (3.18)
ψ = Uy. (3.19)
Let us consider Laplace’s equation in spherical polar coordinates (r, θ, ϕ), that is
∇2 φ = 0 (3.20)
or
1 ∂ 2 ∂φ 1 ∂ ∂φ 1 ∂2φ
r + 2 sin θ + 2 2 =0 (3.21)
r 2 ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂ϕ2
Consider the problem with radial symmetry, that is φ is a function of r only, that is
φ = φ(r). Then Laplace’s equation reduces to
1 ∂ 2 ∂φ
r =0 (3.22)
r 2 ∂r ∂r
42
so that
A
φ=− +B r 6= 0 (3.23)
r
m
φ=− (3.24)
4πr
∂φ m
ur = = (3.25)
∂r 4πr 2
and this is outflow. We see that the minus sign in (3.24) corresponds to outward flow
from the origin, i.e. a source of material. So (3.24) is the velocity potential due to a
simple source of strength m at r = 0. If we wanted a sink would make the replacement
m → −m and we would have inflow.
The volume flux Φ through a spherical surface at r = a is radial velocity times the area
of the sphere,
m
Φ= × 4πa2 = m.
4πa2
In cylindrical coordinates
m
φ=− p
4π x2 + ρ2
∂φ mx 1 ∂ψ
ux = = 2 2 3/2
=
∂x 4π(x + ρ ) ρ ∂ρ
∂φ mρ 1 ∂ψ
uρ = = 2 2 3/2
=− (3.26)
∂̺ 4π(x + ρ ) ̺ ∂x
43
so that ψ(x, ρ) is the streamfunction, that is
mx
ψ(x, ρ) = − p . (3.27)
4π x2 + ρ2
m
φ=− p (3.28)
4π (x − x0 )2 + (y − y0 )2 + (z − z0 )2
3.4 Doublets
If a source of strength m is situated at (d, 0, 0) and a sink of the same strength at (−d, 0, 0),
the potential is a superposition of the two, that is
!
m 1 1
φ= p −p (3.29)
4π (x + d)2 + y 2 + z 2 (x − d)2 + y 2 + z 2
1 1 ∂ 1
p = +d .
(x + d)2 + y 2 + z 2 r ∂x r
m ∂ 1 2md x
φ= 2d =−
4π ∂x r 4π r 3
limd→0 2md = µ
−µx
φ= (3.30)
4πr 3
This satisfies Laplaces’s equation and is called a doublet. The axis of a doublet is the
lines drawn from the sink to the source, in the case the x axis. Likewise with y and z
directions
µ2 y
φ = −
4πr 3
µ z
φ = − 33 (3.31)
4πr
44
In general
~µ · ~r
φ=− (3.32)
4πr 3
∂φ 1 ∂ψ
ur = = 2
∂r r sin θ ∂θ
1 ∂φ 1 ∂ψ
uθ = =−
r ∂θ r sin θ ∂r
The formula for the velocities in terms of the streamfunctions will be given in chapter 10.
∂φ 1 ∂ψ
uρ = =−
∂ρ ρ ∂z
∂φ 1 ∂ψ
uz = =
∂z ρ ∂ρ
Again the formula for the velocities in terms of the streamfunctions will be given in chapter
10.
45
3.5.1 Uniform Flow
φ = Ur cos θ
1
ψ = Ur 2 sin2 θ
2
as
∂φ 1 ∂φ
ur = = U cos θ and uθ = = −U sin θ
∂r r ∂θ
and
∂φ 1 ∂ψ 1 ∂ψ
ur = = 2 = U cos θ and uθ = − = −U sin θ.
∂r r sin θ ∂θ r sin θ ∂r
φ = Ux
The streamfunction is
1
ψ = Uρ2
2
as
∂φ ∂φ
uρ = =0 and ux = =U
∂ρ ∂x
and
1 ∂ψ 1 ∂ψ
uρ = − =0 and ux = =U
ρ ∂x ρ ∂ρ
46
3.5.2 Source
m
φ=−
4πr
The streamfunction is
m
ψ=− cos θ
4π
m
φ=− p
4π z 2 + ρ2
which is the same function as given above but expressed in cylindrical polar coordinates.
1 ∂ψ m 1 ∂ψ
ur = = and uθ = − = 0.
r 2 sin θ ∂θ 4πr 2 r sin θ ∂r
∂φ ∂ m mρ
uρ = =− p = p
∂ρ ∂ρ 4π z + ρ
2 2 4π z 2 + ρ2
∂φ mz
uz = = p
∂z 4π z 2 + ρ2
3.5.3 Doublet
µ cos θ
φ=−
4πr 2
47
µ sin2 θ
ψ=
4πr
The velocity potential in spherical polar coordinates via the vector potential is
µz
φ=−
4πρ3/2
∂φ ∂ µz 3 µz
uρ = =− 3/2
=
∂ρ ∂ρ 4πρ 2 4πρ5/2
∂φ µ
uz = =−
∂z 4πρ3/2
We can use singular solutions to “build up” the fow around different shape bodies. First
example is the so-called “half-body”.
1 m
ψ(r, θ) = Ur 2 sin2 θ − cos θ (3.33)
|2 {z } | 4π{z }
unif orm source
48
m 1 m
= Ur 2 sin2 θ − cos θ
4π 2 4π
that is
m 1
(1 + cos θ) = Ur 2 sin2 θ
4π 2
2
m 2 θ 1 2 θ θ
2 cos = Ur 2 sin cos
4π 2 2 2 2
which gives
m 1/2 θ
r= cosec
4π 2
A closed body is generated by superimposing a uniform stream plus sources and sinks
such that the outflow due to the sources is exactly balanced by the inflow due to the sinks,
with the line joining the sources and sinks along the stream direction.
+ −
−d d
Figure 3.3: .
m m
φ = Ur cos θ − + (3.34)
4πr1 4πr2
49
and streamfunction
1 m cos θ1 m cos θ2
ψ = Ur 2 sin2 θ − + (3.35)
2 4π 4π
Important problem - many particles are spherical or nearly spherical. Ball games involve
spherical objects. Spherical inhomogeneties are used to change flow characteristics for
example to “trip” boundary layers or improve mixing in jet engines.
Consider a frame of reference such that the sphere is at rest and the fluid flow at ∞ is U
in the x−direction.
a
U
Figure 3.4: .
Consider the Rankine body - push the soure and sink together, so that we have a dipole
(source-doublet). The velocity potential φ and streamfunction are
µ cos θ
φ = Ur cos θ −
r2
1 2 2 µ sin2 θ
ψ = Ur sin θ + (3.36)
2 r
The first terms correspond to the uniform stream and the second to the dipole.
∂φ 2µ
vr = = U cos θ + 3 cos θ. (3.37)
∂r r
We require that this velocity is zero on the sphere (a physical consideration) so that the
1
boundary condition is ur = 0 on r = a so that µ = − Ua3 . Then
2
50
a3
φ = U r + 2 cos θ
2r
1 2 a3
ψ = U r − sin2 θ (3.38)
2 r
a3
ur = U 1 − 3 cos θ
r
1 ∂φ a3
uθ = = −U 1 + 3 sin θ (3.39)
r ∂θ 2r
Notice that
1
p = p0 − ̺|~u|2
2
1 1
p∞ = p0 − ̺U 2 implying p0 = p∞ + ̺U 2
2 2
therefore
1 1
p = p∞ + ̺U 2 − ̺|~u|2 .
2 2
3
On r = a, ur = 0 and uθ − U sin θ therfore
2
51
p − p∞ 1 2 9 2
= U 1 − sin θ (3.40)
̺ 2 4
If the minimum pressure, pmin , is less than the saturated vapour pressure pc then liquid
π
will caviate. The minimum pressure for a sphere occurs at θ =
2
5
pmin = p∞ − ̺U 2 (3.41)
8
We have pmin ≤ pc if
5
p∞ − ̺U 2 ≤ pc
8
or
r 1/2
8 p∞ − pc
U≥ . (3.42)
5 ̺
5 p − pc
≥ ∞
1
4 2
̺U 2
p∞ − pc
κ= 1
2
̺U 2
The next step is to consider an infinite set of singularities each infinitesimally close to its
neighbour over a finite line segment.
52
For example uniform distribution of sources
m(x − s) m(x − s)
ψ=− =− p (3.43)
4πr 4π (x − s)2 + y 2 + z 2
Let the total strength of the sources be m, take the strength of the source in element δs
as mδs/a, with m constant for a uniform distribution. For a continuous distribution from
s = 0 to s = a if we let δs → 0
Z a
1 m(s)(x − s) m
ψ=− p ds = [R − R] (3.44)
4πa 0
2 2
4π (x − s) + y + z 2 4πa 2
where R2 is the distance from the point (a, 0, 0) and R the distance to the origin.
1 m m
ψ = Ur 2 sin2 θ + (R − R2 ) − cos θ
2 4πa 4π
53
Chapter 4
Or
Z Z Z
D~u
̺ dV = σ · ~ndS + ̺f~dV (4.2)
V Dt S V
by taking the derivative inside the integral, so that it becomes a substantial dreivative.
On applying the Divergence theorem,
Z Z
~
∇ · AdV = ~ · ~ndS,
A
V S
Z
D~u ~
̺ − ̺f − ∇ · σ dV = 0 (4.3)
V Dt
D~u
̺ = ̺f~ + ∇ · σ (4.4)
Dt
54
or in subscript notation
Dui ∂σij
̺ = ̺fi + (4.5)
Dt ∂xj
4.2.1 Pressure
For a real fluid (that is viscous, heat conducting, ...) in motion, we shall define the
(mechanical) pressure by
1 1
p = − T r(σij ) = − σii .
3 3
In Newtonian fluid there is only a linear dependence on the rate of strain tensor in which
case,
1
However by definition p = − σii , so wemust have the following relation between λ and µ
3
2
λ = − µ.
3
Therefore the constitutive relation for a Newtonianfluid isgiven by
1
σij = −pδij + 2µ eij − ekk δij
3
This relation also tells us that the principal axis for σij and eij coincide.
4.2.2 Viscosity
55
4.2.3 Incompressible Newtonian Fluid
ekk = ∇ · ~u = 0 (4.7)
so
This is the constitutive relation on which is nearly all these nots are based.
Dui ∂σij
̺ = ̺fi + (4.9)
Dt ∂xj
which together with the incompressibility condition for the continuity equation
∂uj
=0 (4.10)
∂xj
!
∂ui ∂uj
σij = −pδij + µ +
∂xj ∂xi
" !#
Dui ∂p ∂ ui 2
∂ ∂uj
̺ = ̺fi − +µ 2
+ (4.11)
Dt ∂xi ∂xj ∂xi ∂xj
The last term on the RHS of this equation is identically zero because of the continuity
equation. The Navier-Stokes equations emerges
56
Dui ∂p ∂2u
̺ = ̺fi − + µ 2i (4.12)
Dt ∂xi ∂xj
or in vector notation
D~u
̺ = −∇p + µ∇2~u + ̺f~ (4.13)
Dt
f~ = −∇Ω (4.14)
where Ω is the potential energy per unit mass. For example gravity
f~ = −g~k (4.15)
Ω = gz. (4.16)
p′ = p + ̺Ω (4.17)
D~u
̺ = −∇p′ + µ∇2~u (4.18)
Dt
It is convenient to write the equations as
D~u 1
= − ∇p′ + ν∇2 ~u (4.19)
Dt ̺
∂~u 1
+ ~u · ∇~u = − ∇p′ + ν∇2~u (4.20)
∂t ̺
57
4.4 Alternative Form of the Navier-Stokes Equation
and the Vorticity Equation
We first derive a couple of identities using ǫijk ǫki′ j ′ = (δii′ δjj ′ − δij ′ δji′ ):
[~u × ~ω ]i = ǫijk uj ωk
= ǫijk uj (ǫki′ j ′ ∂i′ uj ′ )
= ǫijk ǫki′ j ′ uj ∂i′ uj ′
= (δii′ δjj ′ − δij ′ δji′ )uj ∂i′ uj ′
= uj ∂i uj − uj ∂j ui
1 2
= ∂ u − ~u · ∇ui
2 i
1
= [ ∇~u2 − ~u · ∇~u]i
2
and
[∇ × ~ω ]i = ǫijk ∂j ωk
= ǫijk ∂j (ǫki′ j ′ ∂i′ uj ′ )
= ǫijk ǫki′ j ′ ∂j ∂i′ uj ′
= (δii′ δjj ′ − δij ′ δji′ )∂j ∂i′ uj ′
= ∂j ∂i uj − ∂j ∂j ui
= [−∇2 ~u + ∇(∇ · ~u)]i .
1
~u · ∇~u = ∇|~u|2 − ~u × ~ω (4.21)
2
and
∇ × ~ω = −∇2~u. (4.23)
58
Substitution of (4.21) and (4.23) into (4.20)
∂~u 1 2 p
− ~u × ω
~ = −∇ |~u| + − ν∇ × ~ω (4.24)
∂t 2 ̺
”This is the ‘best’ form of the equation to use when considering coordinate systems other
than Cartesian coordinates”.
∂~ω
+ ~u · ∇~ω = ~ω · ∇~u + ν∇2 ~ω (4.25)
∂t
~ω · ∇~u = 0. (4.26)
D~ω
= ν∇2 ~ω . (4.27)
Dt
It is possible to show that without viscous forces there cannot be a vorticity fieled in the
flow which varies with coordinate directions. This is expressed in the following relationship
∂τij ∂ωk
= µ∇2 ui = −µǫijk
∂xj ∂xj
(recall σij = −pδij + τij where τij = 2µeij ). The equality of the last expression with the
central expression can be derived in the following manner
59
∂ωk ∂ ∂
−µǫijk = −µǫijk (ǫkmn u )
∂xj ∂xj ∂xm n
∂ ∂
= −µǫijk ǫkmn u
∂xj ∂xm n
∂ 2 un
= −µ(δim δjn − δin δjm )
∂xj ∂xm
∂ ∂uj ∂ 2 ui
= −µ( − )
∂xi ∂xj ∂xj ∂xj
= µ∇2 ui . (4.28)
If invisid flow starts with no vorticity then no vorticity will be produced. To understand
this intuitively we note that of the three types of force that can act on a cubic fluid
element, the pressure, body forces, and viscous forces, only the viscous shear forces are
able to give rotary motion. Hence if the viscous effects are nonexistent, vorticity cannot
be introduced.
60
Chapter 5
Streamfunction in 2-Dimensions
If
∇ · ~u = 0 (5.1)
~ x, t) such that
then it follows that there exists a vector field A(~
~
~u = ∇ × A. (5.2)
5.1 Cartesians
∂u ∂v
+ =0
∂x ∂y
is solved by choosing
∂ψ ∂ψ
u= and v=− .
∂y ∂x
~ for A
The continuity equation is solved if we write ~u = ∇ × A ~ = ψ(x, y, t)~z, then
61
~ex ~ey ~ez
~ = ∂ ∂ ∂
~u = ∇ × A
∂x ∂y ∂z
0 0 ψ(x, y, t)
∂ψ ∂ψ
= ~ex − ~e (5.3)
∂y ∂x y
thus
∂ψ ∂ψ
u= , v=− (5.4)
∂y ∂x
x = r cos θ, y = r sin θ.
From section 3.1.2 we see that the formula for the velocity components ur and uθ in terms
of the velocity potential
∂φ 1 ∂φ
ur = and uθ = . (5.5)
∂r r ∂θ
1 ∂ 1 ∂uθ
(rur ) + =0 (5.6)
r ∂r r ∂θ
1 ∂ψ ∂ψ
ur = and uθ = − (5.7)
r ∂θ ∂r
62
We obtain these also from the curl in cylindrical polar coordinates
We demonstrate that ψ is indeed constant along the streamlines using the formula (5.7)
that we have obtained relating the velocity components to this streamfunction. The
differential equation for streamlines are
d~x
= λ~u.
ds
and thus dr = λur ds and rdθ = λuθ ds. Using this find
∂ψ ∂ψ
dψ = dr + dθ
∂r ∂θ
= −uθ dr + rur dθ
= −uθ λur ds + ur λuθ ds = 0. (5.9)
Since the streamfunction is defined completely by the geometry of the streamlines, it must
be the same in any coordinates system. So if ψC (x, y, t) is the streamfunction in Cartesian
coordinates, the stream function in polar coordinates will be
63
∂~r
= cos θ~i + sin θ~j.
∂r
∂~r
= −r sin θ~i + r cos θ~j
∂θ
~u = ur~er + uθ~eθ
= ur [cos θ~i + sin θ~j] + uθ [− sin θ~i + cos θ~j]
= [u cos θ − u sin θ]~i + [u sin θ + u cos θ]~j
r θ r θ (5.11)
∂ψ
u = = ur cos θ − uθ sin θ
∂y
∂ψ
v = − = ur sin θ + uθ cos θ (5.12)
∂x
∂ψC ∂ψ
ur = cos θ − sin θ C
∂y ∂x
∂ψ ∂ψ
uθ = − sin θ C − cos θ C (5.13)
∂y ∂x
64
Now using the chain rule we can write formula for the derivatives
∂ ∂x ∂ ∂y ∂
= +
∂r ∂r ∂x ∂r ∂y
∂ ∂
= cos θ + sin θ (5.14)
∂x ∂y
and
∂ ∂x ∂ ∂y ∂
= +
∂θ ∂θ ∂x ∂θ ∂y
∂ ∂
= −r sin θ + r cos θ . (5.15)
∂x ∂y
∂ ∂ 1 ∂
= cos θ − sin θ
∂x ∂r r ∂θ
∂ ∂ 1 ∂
= sin θ + cos θ (5.16)
∂y ∂r r ∂θ
∂ 1 ∂ ∂ 1 ∂
ur = cos θ sin θ + cos θ ψ − sin θ cos θ − sin θ ψ
∂r r ∂θ P ∂r r ∂θ P
1 ∂ψ
= (5.17)
r ∂θ
∂ 1 ∂ ∂ 1 ∂
uθ = − sin θ sin θ + cos θ ψ − cos θ cos θ − sin θ ψ
∂r r ∂θ P ∂r r ∂θ P
∂ψ
= − (5.18)
∂r
In polar coordinates the formula for the velocity components in terms of the stream
function are
1 ∂ψ ∂ψ
ur = , uθ = − (5.19)
r ∂θ ∂r
where ~u = ur~er + uθ~eθ .
65
Chapter 6
∂uk ∂2φ
ωi = ǫijk = =0
∂xj ∂xj ∂xk
b) Potential flow is not influenced by viscous forces. Recall the stress tensor can be
written: σij = −pδij + τij where τij is the viscous stress tensor - see (4.9). The viscous
term in the Navier-Stokes equation is
∂τij
= µ∇2 ui .
∂xj
From ∇ × ~ω = ∇ × (∇ × ~u) = ∇(∇ · ~u) − ∇2~u = −∇2 ~u (where we have used ∇ · ~u = 0),
∂ωk
µ∇2 ui = −µǫijk = 0.
∂xj
0 = ∇ · ~u = ∇ · (∇φ) = ∇2 φ.
66
6.2 Complex Representation
We will show in the following that any analytic function represents a two dimensional
flow.
~ = ψ~k such that
Incompressible - ∇ · ~u = 0 implies that there exists a vector field A
~u = ∇ × ψ~k or
∂ψ ∂ψ
u= , v=− (6.1)
∂y ∂x
∂φ ∂φ
u= , v= . (6.2)
∂x ∂y
∂u ∂v
+ = ∇2 φ = 0.
∂x ∂y
∂v ∂v
−ω = − + = ∇2 ψ = 0.
∂x ∂y
Thus both the velocity potetial φ and the streamfrunction ψ the Laplace equation.
These imply
∂ψ ∂φ ∂ψ ∂φ
= , − = (6.3)
∂y ∂x ∂x ∂y
which are the Cauchy-Riemann equations. These are necessary condition for the complex
function φ(x, y) + iψ(x, y) to be analytic, as we shall now show. First we must note that
requiring that a derivative of a complex function should be the same independent of the
direction we take the limit in the complex plane. We have
67
1
f ′ (z) = lim [f (z + ∆z) − f (z)]
∆z→0 ∆z
1
= lim [φ(x + ∆x, y) − φ(x, y) + iψ(x + ∆x, y) − iψ(x, y)]
∆x→0 ∆x
1
= lim [φ(x, y + ∆y) − φ(x, y) + iψ(x, y + ∆y) − iψ(x, y)] (6.4)
i∆y→0 ∆x
Equating the real parts of thes last two give lines give
∂φ ∂ψ
=
∂x ∂y
∂ψ ∂φ
=− .
∂x ∂y
Thus we have that the Cauchy-Riemann equations as a necessary requirement for a com-
plex function to be analytic.
dw ∂φ ∂ψ
= +i
dz ∂x ∂x
dw
= u − iv (6.5)
dz
Note that
dw √
= u2 + v 2
dz
68
(c) Flow around a corner
ux = u = U, v=0
which implies that φ = Ux+k, ψ = Uy +k ′ where k and k ′ are arbitrary amd do not effect
the velocity or Cauchy-Riemann equations as these all involve derivatives, and hence can
be dropped without loss of generality. The complex potential is w(z) = Uz and
dw
= U = U + i0
dz
We now calculate the velocity field from the complex potential w = Ue−iα z.
dw
= Ue−iα = U(cos α − i sin α)
dz
u = U cos α
v = U sin α. (6.6)
69
y
√
Obviously u2 + v 2 = U.
It is easy to show that multiplying any complex potential w(z) by e−iα rotates the velocity
vector anti-clockwise by an angle of α at each point in space: as
dw(z)
= u(x, y) − iv(x, y)
dz
then
d(w(z)e−iα )
= [u(x, y) − iv(x, y)][cos α − i sin α]
dz
= u(x, y) cos α − v(x, y) sin α − i[u(x, y) sin α + v(x, y) cos α]
cos α − sin α u(x, y)
= . (6.7)
sin α cos α v(x, y)
Note that we are rotating the velocity vector at each point in space rather than rotating
about the origin.
dw
= 0. (6.8)
dz z=zs
For some potential w(z) (analytic about zs ) consider a Taylor expansion of w(z) about zs
to invetigate flow near a stagnation point.
70
1
w(z) = w(zs ) + (z − zs )w ′ (zs ) + (z − zs )2 w ′′ (zs ) + O(|z − zs |3 )
2
so that
1
w(z) = w(zs ) + (z − zs )2 w ′′ (zs ) + O(|z − zs |3 ). (6.9)
2
Assuming that w ′′ (zs ) 6= 0 it may be written in polar coordinate form w ′′ (zs ) = Aeiα
A > 0 and α real. The term w(zs ) is a constant and so does not effect the resulting
velocity field and so can be neglected. The flow resulting from (6.9) is the same as the
flow resulting from (1/2)z 2 Aeiα + O(|z|3 ) but shifted by an amount xs in the x−direction
and an amount ys in the y−direction. We are considering the flow near the stagnation
point and so the terms represented by O(|z − zs |3 ) are to be neglected. We know that a
factor eiα rotates the velocity vector by an angle of α clock-wise at each point in space.
In the case of monomials in z − zs , we will find that we can obtain the flow by performing
a rotation about the point zs . First put z − zs = (x − xs ) + i(y − ys ). Consider the (x′ y ′ )
axis defined by x′ = x − xs and y ′ = y − ys . Then consider this axis rotated through −α/2
and becoming the X − Y axis so that
′
x cos(α/2) sin(α/2) X
= . (6.10)
y ′
− sin(α/2) cos(α/2) Y
(see fig 6.3 and note for example that X = a and Y = 0 corresponds to x′ = a cos(α/2)
and y ′ = −a sin(α/2)) this is effected via
iα
(z − zs ) = (X + iY )e− 2
= [X cos(α/2) + Y sin(α/2)] + i[−X sin(α/2) + Y cos(α/2)] (6.11)
1
w = C + (X + iY )2
2
1
= Re(C) + (X 2 − Y 2 ) + i[Im(C) + AXY ]. (6.12)
2
Using
71
y′ Y
y
α/2 x′
zs
X
w = φ̃(X, Y ) + iψ̃(X, Y )
and
∂ ψ̃(X, Y ) ∂ ψ̃(X, Y )
uX = and uY = − . (6.14)
∂Y ∂X
So
So the flow in the neighbourhood of zs in the X −Y frame is the same as the flow generated
1
by the complex potential z 2 .
2
w(z) = Az n , (6.16)
72
Y
φ = Ar n cos (nθ)
ψ = Ar n sin (nθ) (6.18)
dw
= Anz n−1 . (6.19)
dz
1 ∂ψ
ur =
r ∂θ
∂ψ
uθ = −
∂r
Consider the streamlines Ψ = 0. This obviously implies that uθ = 0, which could concur
with there being a wall through which there is no flow. In this case 0 ≤ θ ≤ π/n.
73
At θ = 0 and θ = π/n (n 6= 1) we have uθ = 0, while at θ = π/2n ur = 0 and uθ has
its maximum value. Flow is parallel to the walls at the walls - the velocity component
perpendicular to the wall is zero as it should be. At the point r = 0 we have ur = uθ = 0
so the velocity of the fluid is zero there and we have a stagnation point.
Ψ=0
π
n
Ψ=0
These solutions corresponing to integer value n can be extended by having the wall at
2π/n as shown in fig (5.4) while retaining the same velocity formula (6.20) but now with
the variable θ extended to the range 0 ≤ θ ≤ 2π/n. Where there was a wall there is
now a stagnation point streamline. The streamfunction is zero on the walls and on the
stagnation point streamline.
Ψ=0
Ψ=0
π
n
Ψ=0
Figure 6.6: Extended solution from flow at a corner with walls at θ = 0 and θ = 2π/n.
74
dw(z)
= 2Az
dz
= 2A(x + iy).
Implying
u = 2Ax
v = −2Ay
Ψ=0
Or if we allow
In chapter 7 we will reconsider this flow but for non-zero viscocity where we demand that
u = v = 0 at the wall but that the flow is still described by ψ = Axy as y → ∞.
2π n−1
ϑ = 2π − = 2π
n n
75
Figure 6.9: Stagnation point flow when n = 2.
Uniform flow over and below a plate for n = 1 with walls at θ = 0, θ = π and
θ = 2π
6.4 Singularities
(b) Vortex
76
6.4.1 Source (and Sink)
w(z) = m ln z (6.21)
dw m
= .
dz z
Recall that in polar coordinates m ln z = m(ln r + iθ), so that the streamfunction is given
by ψ = mθ
So the streamlines (ψ is a constant) are straight lines radiating from the source
dw m dw m
= e−iθ , =
dz r dz r
dw
From u − iv = dz
we have
m m
u= cos θ, v= sin θ (6.22)
r r
6.4.2 Vortex
w(z) = Γi ln z, (6.23)
with derivative
dw iΓ
=
dz z
As
77
1 ∂ψ
ur = =0
r ∂θ
and
∂ψ Γ
uθ = − =−
∂r r
I
~u · d~r
C
I Z 2π
Γ
~u · d~r = uθ adθ = − a2π = −Γ2π (6.24)
C 0 a
Place a source and sink on the x−axis a distance d from the origin. The complex potential
is
!
d
1+ z
w(z) = m ln d
.
1− z
m −m
x
−d d
78
We have the Taylor expansions
x2 x3
ln(1 + x) = x − + − ...
2 3
x2 x3
ln(1 − x) = −x − − − ...
2 3
d d
w(z) = m ln 1 + − m ln 1 −
z z
2 3
d 1d 1d d 1 d2 1 d3
= m − + − ... − m − − − + ...
z 2 z2 3 z3 z 2 z2 3 z3
2md 1 d2
= 1+ + ... (6.25)
z 3 z2
Now we consider the dual limit of d → 0 whilst m → ∞ and call the limit of 2md µ so
we have
µ
w(z) = (6.26)
z
µ µ µ(x − iy)
w(z) = = = 2
z x + iy x + y2
µy
ψ=−
x2 + y2
µy
x2 + y 2 = −
ψ0
or
2
2 µ µ2
x + y+ =
2ψ0 4ψ02
79
µ µ
which are circles with centre at (0, ) and radius .
2ψ0 2|ψ0 |
More generally the complex potential for a dipole at z0 = x0 + iy0 which makes an angle
α with the x−axis is given by
µeiα
w(z ∗ ) = . (6.27)
z ∗ − z0∗
µ
z∗
µ
.
z
Then substituting
x 7→ x′ = x − x0 y 7→ y ′ = y − y0
shifts the origin to (x0 , y0 ) and this would be given by the complex potential
µ
.
z ∗ − z0∗
80
Then multiplying by eiα gives
′
X cos α − sin α x
= . (6.29)
Y sin α cos α y′
µ
w(z) = Uz + (6.30)
z
dw µ
(z) = U − 2 (6.31)
dz z
a2 a2
w(z) = U z + = U x + iy + (6.32)
z x + iy
a2 y x2 + y 2 − a2
ψ=U y− = Uy (6.33)
x2 + y 2 x2 + y 2
The streamlines are given by ψ being a constant and a dividing streamline is when ψ = 0
for x2 + y 2 = a2 which is the equation of a circle of radius a, if we consider this dividing
81
y
ψ=0
Figure 6.14: Superposition of uniform flow and a dipole results in ψ = 0 as a solid object,
a cylinder.
streamline to be solid, we then have a cylinder. In terms of polar coordinates the cylinder
is given by z = aeiθ so that
dw a2
= U 1− 2
dz z
−2iθ
= U(1 − e )
= Ue−iθ (eiθ − e−iθ )
= 2iU sin θe−iθ (6.34)
therefore
2
dw
= 4U 2 sin2 θ
dz
By Bernoulli’s equation,
p 1 2
+ |~u| = k,
̺ 2
at infinity
p∞ 1 2
+ U =k
̺ 2
and at cylinder
82
p 1 2 2
+ 4U sin θ = k,
̺ 2
p p 1
= ∞ + U 2 (1 − 4 sin2 θ). (6.35)
ρ ρ 2
ψ=0 y
ψ=0 a x ψ=0
w(z) = Uz
Figure 6.15: Flow around a cylinder. We have a dividing streamline for ψ = 0 as a circle
of radius a about the origin. Another solution for ψ = 0 is the line y = 0. The points
denoted by black circles denote stagnation points, namely where dw/dz = 0. The x’s
denote points of minimum pressure. Cavitation occurs here first.
Notice that minimum pressure occurs when θ = ±π/2 that is when sin2 θ = 1. Therefore
from (6.35) we have
3
pmin = p∞ − ̺U 2 .
2
3
p∞ ≥ ̺U 2
2
When p∞ is equal to 32 ̺U 2 cavitation will occur and the equations will no longer apply.
Note that pmax occurs at θ = 0 and at π i.e. at the stagnation points, this is always the
case for steady flow.
83
6.6 Flow Around a Rotating Cylinder
We can obtain flow around a rotating cylinder if we add a vortex to the previous solution.
a2 z
w(z) = U z+ + iΓ ln (6.36)
z a
which is equivalent to
a2
w(z) = U z + + iΓ ln z − iΓ ln a} .
z | {z
constant
The derivative is
dw a2 iΓ
=U 1− 2 + . (6.37)
dz z z
Now
a2 y
ψ = Im(w) = U y− 2 + Γ ln r − Γ ln a (6.38)
x + y2
ψ(x, y = 0) = Γ ln x − Γ ln a 6= const.
The radial and angular velocity components can be obtained from the formula
1 ∂ψ ∂ψ
ur = , uθ = − .
r ∂θ ∂r
84
The streamfunction (6.38) written in terms of r and θ is
a2
ψ(r, θ) = Ur 1 − 2 sin θ + Γ ln r − Γ ln a
r
a2
ur = U 1 − 2 cos θ
r
a2 Γ
uθ = −U 1 − 2 sin θ − (6.39)
r r
dw
=0
dz
iΓ
z 2 − a2 + z=0
U
s
iΓ Γ2
zs = − ± − + a2 . (6.40)
2U (2U)2
2 Γ2 2 Γ2
|zs | = − 2
+a + 2
= a2
(2U) (2U)
and hence the stagnation points lie on the cylinder’s surface. As 0 < sin β < 1 for
0 < β < π/2, there is a β in this range such that Γ = 2aU sin β. We must solve
85
Γ
a sin β =
2U
resulting in
q
zs = −ia sin β ± a2 − a2 sin2 β
= −ia sin β ± a cos β. (6.41)
s
iΓ Γ2
zs = − +i − a2 .
2U (2U)2
s
iΓ Γ2
zs = − −i − a2
2U (2U)2
dw iΓ
= U 1 − e−2iθ + e−iθ
dz z=aeiθ a
dw −iθ Γ
= ie 2U sin θ +
dz a
86
Γ=0 0 < Γ < 2aU
Figure 6.16: Flow around a clockwise rotating cylinder for different values of Γ.
and so
2 2
dw 2 Γ
= ~u = 2U sin θ +
dz a
This splits as
4ΓU Γ2
~u2 = 4U 2 sin2 θ + sin θ + 2
a a
The first parts are the previous velocity and the other is the addad circulation, from
Bernoulli’s equation ̺p + 21 U 2 = k
p 1 2
By Bernoulli’s equation, + |~u| = k first at infinity
̺ 2
p∞ 1 2
+ U =k
̺ 2
p 1 2 2 4ΓU Γ2
+ 4U sin θ + sin θ + 2 = k,
̺ 2 a a
87
p p 1 1 4ΓU Γ2
= ∞ + U2 − 2
4U sin θ + 2
sin θ + 2 (6.42)
̺ ̺ 2 2 a a
Z
F~ = −p~ndS
S
for a cylinder
Z 2π
F~ = − p(θ)~er adθ (6.43)
0
1 4ΓU
− ̺ sinθ~er
2 a
So that
Z 2π
F~ = 2̺ΓU sin θ(cos θ~ex + sin θ~ey )dθ (6.44)
0
R 2π R 2π
The integral 0
sin θ cos θdθ = 0 by symmetry. We are left with the integral 0
sin2 θdθ,
Z 2π Z 2π Z 2π
2 1 1
sin θdθ = (1 − cos 2θ) = dθ = π.
0 2 0 2 0
F~ = 2̺ΓUπ~ey . (6.45)
We are familiar with the lift generated by a curving ball (backspin and top spin). The
flow associated with a rotating cylinder two dimensional and is more easy to understand.
88
Figure 6.17: Flow around a rotating cylinder results in an upward force on the cylinder.
µeiα µei(π−α)
w = − −
z−a z+a
iα
µe µe−iα
= − + (6.47)
z−a z+a
µeiα µe−iα
w= +
a − iy a + iy
which is obviously real (being the sum of a complex number and its complex conjugate)
and hence corresponds to ψ = 0.
89
6.7.2 Continuous Distributions
Z a
m(s)
w(z) = ln(z − s)ds (6.48)
0 a
and
Z a
dw m(s) 1
= u − iv = ds (6.49)
dz 0 a z−s
we can consider the case we have already seen, which is m is a constant. The intragl is
easy to perform
Z a
1
ds = [ln(z − s)]a0 = ln(z − a) − ln z,
0 z−s
so we end up with
dw m z−a
= ln (6.50)
dz a z
dw m r1 eiθ1 m r m
= ln = ln( 1 ) + i (θ1 − θ)
dz a reiθ a r a
Therefore u = m
a
ln( rr1 ) and v = m
a
(θ1 − θ).
Consider the angle preserving transformation ξ = ξ(z). We can use this technique to find
the complex potential for a large variety of geometries.
90
Chapter 7
Boundary Conditions
D~u 1
= − ∇p + ν∇2~u
Dt ̺
is a second order differential equation in space. This implies that each of the 3 velocity
components must have two boundary conditions for the problem to be well-posed.
There are two sets of boundary conditions here, one on hte body and the other at ∞. At
∞ the flowtends to the free stream of constant velocity U∞ int the x−direction, that is
~ → (U , 0, 0)
U ∞
(~u · ~n)S = 0.
2. Non-Slip condition
In this case the velocities in tangental and binormal direction (the binormal is the cross
product of the tangential and normal vector and so is orthogonal to the normal and given
tangential vector. It also tangential to the surface but orthognal to the given tangential
vector) are zero (that is the fluid ‘sticks’ to the surface - hence the non-slip condition).
91
Thus the velocity on the surface can be set identically equal to zero because the velocity
in all three orthogonal directions (i.e. normal tangential and binormal) are zero, that is
~u = 0 on S.
~u → at ∞
~u = U~ on S. (7.1)
3. Free-surface conditions.
(a)
Let F (~x, t) = 0 be the equation of any surface moving with the fluid (for example ocean
surface), then a prticle of fluid remains on the surface. This tells us that the total
(substantial) derivative of F must be indentically zero, that is
DF ∂F
= + ~u · ∇F = 0.
Dt ∂t
For example with two-dimensional water waves we might define the free surface height by
y = ζ(x, t)
FIG HERE
F = ζ(x, t) − y = 0
becomes
∂ζ ∂ζ
+u − v = 0.
∂t ∂x
(b)
92
Chapter 8
Let us now scale the Navier-Stokes equation with respect to the following characteristic
scales of the flow field
U = Velocity scale
L = Characteristic length
ω = Characteristic frequency. (8.1)
∂~u 1 2
+ ~u · ∇~u = − ∇p + ν∇
| {z ~u} (8.2)
∂t
| {z } ̺
| {z } viscous
Inertial pressure
νU
3. ν∇2 ~u = O L2
.
1.
93
ωU ωL2
O =O
νU/L2 ν
2.
U 2 /L UL
O =O
νU/L2 ν
2
The quantity ωLν is often referred to as the oscillatory Reynolds number (which is related
to the Strouhal number) while,
UL UL
R= =̺ (8.3)
ν µ
is the famous Reynolds number which represents the ratio of inertial forces to viscous
forces. If R >> 1 then inertial forces dominate the flow field. Conversely, if R << 1 then
viscous forces dominate the flow field.
We can make the Navier-Stokes equation dimensionless if the units of time, distance and
velocity are scaled in accordance with
L ′
t = t
U
~x = L~x′
~u = U~u′ (8.4)
That is, distance is measured in multiples of L, time in multiples of U/T and velocity in
multiples of U. We also have
F U2 1 1
p= = (̺L3 ) F ′ 2 ′ = ̺U 2 p′ .
A L L A
With these new variables, the derivatives get changed from ∂/∂x to (1/L)∂/∂x′ , and so
on. We have for the terms of the Navier-Stokes equation
94
∂~u ̺U 2 ∂~u
̺ →
∂t L ∂t
̺U 2 ′
̺~u · ∇~u → ~u · ∇′~u′
L
̺U 2 ′ ′
∇p → ∇p
L
µU ′ 2 ′
µ∇2~u = ∇ ~u (8.5)
L2
∂~u′ ′ ′ ′ 1 ′ ′ 1 ′2 ′
+ ~
u · ∇ ~
u = − ∇ p + ∇ ~u . (8.6)
∂t′ ̺ R
Whatever the scale, flows with the same R “look” the same - in terms of appropriate
scaled x′ , y ′ , z ′ and t′ .
The basis of tests on scaled down models in a wind or water tunnel based on the concept
of dynamical similarity that is experiments should be conducted at the same Reynolds
number.
We are able to predict the values quantities to be expected on a prototype from measure-
ments on a model.
For example: for a ship the Reynolds number R is the same in a water channel as an
ocean. As ν is the same because the fluid is the same, then if L decreases, U must increase
to keep R constant.
Usefull in experiments to use the same liquid for example but to change the Reynolds
number by varying other characteristic quantities.
These equations
∇p = µ∇2~u (8.7)
are called the Stokes flow equations where viscous effects totally dominate over inertial
forces. The flow is characterised by either extreme viscosities or microscopic length scales
and velodcities.
95
Chapter 9
As we shall see here the nonlinear term (~u · ∇)~u of the Navier-Stoke’s equations vanishes.
This has the significant advantage that the equation is linear allowing the use of powerfull
mathematical techniques for linear analysis.
(b) Circular streamlines. Example of where the nonlinear term decouples from the equa-
tion which determines the velocity vector field.
∂u ∂v
+ = 0.
∂x ∂y
D
∇2 Ψ = ν∇4 Ψ (9.1)
Dt
expanding the substantial derivative
D ∂ ∂ ∂
∇2 Ψ = ∇2 Ψ + u ∇2 Ψ + v ∇2 Ψ
Dt ∂t ∂x ∂y
∂ ∂Ψ ∂ ∂Ψ ∂
= ∇2 Ψ + ∇2 Ψ − ∇2 Ψ
∂t ∂y ∂x ∂x ∂y
2
∂(∇ Ψ, Ψ)
=
∂(x, y)
= ν∇4 Ψ (9.2)
96
where the Jacobian is defined by
∂(∇2 Ψ, Ψ) ∂Ψ 2 ∂Ψ ∂Ψ 2 ∂Ψ
= ∇ − ∇ . (9.3)
∂(x, y) ∂y ∂x ∂x ∂y
The RHS of this is nonlinear and again is a majour source of difficulty when trying to
solve the Navier-Stokes equations.
Unidirectional (or rectilinear) flow is when the velocity vector field is given by ~u = (u, 0, 0),
that is
~u = u(~x, t)
∂u
=0 (9.4)
∂x
u = u(y, z, t).
∂u
~u · ∇~u = u =0
∂x
because (9.4).
As there are no y and z components of velocity there can be no forces on the fluid in the
the y or z direction, implying
∂p ∂p
= 0, = 0. (9.5)
∂y ∂z
97
p = p(x, t).
∂u 1 ∂p
=− + ν∇2 u (9.6)
∂y ̺ ∂x | {z }
|{z} | {z } (3)
(1) (2)
for unidirectional flows. It has the significant advantage that the equation is linear al-
lowing the use of powerfull mathematical techniques for linear analysis. If we look more
closely at (9.6), we see that parts (1) and (3) are functions of y, z and t whereas (2) is
a function of x and t. As x is an independent variable ∂p/∂x cannot be a function of x
because neither of the other two terms are. It can be at most a function of t. Therefore
the pressure term can be at most be a function of t, that is
1 ∂p
− = G(t). (9.7)
̺ ∂x
∂u
= G(t) + ν∇2 u. (9.8)
∂t
In practice the pressure gradient is usually prescribed and the problem is to determine u,
subject to specified boundary conditions.
(i) 2-D Poiseuille flow - the flow between two flat plates a distance b apart.
98
9.2.1 2-D Poiseuille Flow
The no-slip boundary condition at the surface of the cylinder r = a provides us with one
boundary condition
u = 0 on r = a.
A “smoothness” condition at r = 0 of
∂u
= 0 on r = 0.
∂r
This says that there is no “stress jump” along the axis of symmetry.
∂u G1 A
=− r+
∂r ν2 r
G 2
u(r) = − r +B (9.10)
4ν
G 2
u(r) = (a − r 2 ) (9.11)
4ν
Same as with (9.9) but now the non-slip conditions aplly to both cylinders, that is
u = 0 on r = a, b.
99
G 2
u(r) = − r + A ln r + B (9.12)
4ν
Using the non-slip boundary conditions (where a is the radius of the outer cylinder and
b is the radius of the inner cylinder)
G 2
0 = − a + A ln a + B (9.13)
4ν
G
0 = − b2 + A ln b + B (9.14)
4ν
Subtracting we get
G (a2 − b2 )
A=
4ν ln a/b
G 2 G (a2 − b2 )
B= a − ln a
4ν 4ν ln a/b
G 2 2 (a2 − b2 ) ln r/a
u(r) = a −r + . (9.15)
4ν ln a/b
In the limit b → 0, ln a/b → −∞ and we recover the parabolic profile given in the previous
section.
∂2u
=0
∂y 2
100
u = 0 on y = 0
u = U on y = b
The solution is
Uy
u= (9.16)
b
~i ~j ~k
U
~ω = ∇ × ~u = ∂
∂x
∂
∂y
∂
∂z = − ~k. (9.17)
Uy b
b
0 0
This simple flow is exploited in the design of the Couette viscometer which is used to
measure the viscocity of a liquid.
∂u ∂2u
=ν 2
∂t ∂y
101
u(∞, t) = 0
while the no slip boundary condition for flow at the plate for t > 0 implies
The equation is the diffusion equation which is known to have a similarity solution.
It is easy
∂u ′ ∂η ′ 1
= U0 f (η) = U0 f (η) − η
∂t ∂t 2t
and
∂2u 1
2
= U0 f ′′ (η)
∂y 4νt
f ′′ + 2ηf ′ = 0 (9.20)
or
2
(e−η f )′ = 0
so
2
f ′ = Ae−η
102
Z η
2 2
erf (η) = √ e−ξ dξ. (9.22)
π 0
We now find C and D from the boundary conditions, recalling that u(y, t) = U0 f (η) and
1/2
η = y/(2(νt)
R ∞ −ξ2 . From
√ no flow at infinity u(∞, t) = 0 we get C = −D (where we have
used 0 e dξ = π/2). From u(0, t > 0) = U0 we get D = U0 . Finally we have
The solution can be generalised to a general velocity U(t), t > 0, of the infinite plate. We
begin with a couple of trivial cases that will instructive in devloping the general solution.
First consier the case where we have the impulsively started plate that moves at speed
(U1 − U0 ) after t = t1 , then write
then
∂u ′ ∂η ′ 1
= (U1 − U0 )f (η) = (U1 − U0 )f (η) − η
∂t ∂t 2(t − t1 )
and
∂2u 1
2
= (U1 − U0 )f ′′ (η)
∂y 4ν(t − t1 )
We obtain f (η) = Cerf (η) + D where η = y/(2(t − t1 )1/2 ). No flow at infinity implies
again that C = −D. And u(0, t > t1 ) = U1 − U0 implies D = U1 − U0
103
Twice impulsively motioned plate
Second consider the case where we have the impulsively started plate that moves at speed
U0 after t = 0 and then later at speed U1 for t > t1 then the we have for the plate
Let t > t1
X U(τ + dτ ) − U(τ )
u(t) = U(0)Θ(t) + dτ ( )Θ(t − τ − dτ )
dτ
Zτ =0∞
dU
= U(0)Θ(t) + dτ Θ(t − τ ) (9.26)
0 dτ
Z ∞
0 dU(τ )
u(y, t) = U(0)(1 − erf (η ))Θ(t) + Θ(t − τ )(1 − erf (η τ ))dτ
0 dτ
= U(0)(1 − erf (η 0 ))Θ(t) + [U(τ )Θ(t − τ )(1 − erf (η τ ))]∞
0
Z ∞
d
− U(τ ) [Θ(t − τ )(1 − erf (η τ ))]dτ
0 dτ
Z ∞
∂
= U(τ ) [Θ(t − τ )(1 − erf (η τ ))]dτ (9.27)
0 ∂t
where
p
η τ = y/(2 ν(t − τ )).
104
The final solution for an infinite plate.
Z ∞
∂
u(y, t) = U(τ ) [Θ(t − τ )(1 − erf (η τ ))]dτ. (9.28)
0 ∂t
∂u ∂2u
=ν 2
∂t ∂y
we obtain
d2 U1 d2 U2
−U1 (y)σ sin σt + U2 (y)σ cos σt = ν cos σt 2 + sin σt 2 . (9.31)
dy dy
Implying
d2 U1 σ d2 U2 σ
2
= U2 (y), 2
= − U1 (y) (9.32)
dy ν dy ν
Or in matrix form
d2 U1 (y) 0 σ/ν U1 (y)
= . (9.33)
dy 2 U2 (y) −σ/ν 0 U2 (y)
a
eαy (9.34)
b
105
in (9.33)
a 2 bν/σ
α = (9.35)
b −aν/σ
a 1 2 a 1
= (for α = +iν/σ), = (for α2 = −iν/σ)
b i b −i
and
1+i 1−i
α=± √ , α=± √ .
2 2
r r
U1 (y) 1 σ 1 σ
= A exp (1 + i) y +B exp −(1 + i) y
U2 (y) i 2ν i 2ν
r r
1 σ 1 σ
+ C exp (1 − i) y +D exp −(1 − i) y
−i 2ν −i 2ν
(9.36)
At this point we can use the boundary condition at infinity which is that there is no flow
ther to get A = C = 0, so that
r r
U1 (y) 1 σ 1 σ
=B exp −(1 + i) y +D exp −(1 − i) y (9.37)
U2 (y) i 2ν −i 2ν
We get
106
√ √ √
U1 (y) 1 σ
y 1 σ σ
= B e−i 2ν
+D ei 2ν y e− 2ν y
U2 (y) i −i
r r
1 σ σ
= [B (cos y − i sin y )
i 2ν 2ν
1
r
σ
r
σ
√σ
+D (cos y + i sin y )]e− 2ν y
−i 2ν 2ν
r r √
B+D σ 0 σ σ
= [ cos y + sin y ]e− 2ν y
0 2ν B+D 2ν
r r √
0 σ −B + D σ σ
+ i[ cos y + sin y ]e− 2ν y .
B−D 2ν 0 2ν
(9.38)
r r r
U1 (y) 2B σ 0 σ σ
=[ cos y + sin y ] exp − y (9.39)
U2 (y) 0 2ν 2B 2ν 2ν
r r r
σ σ σ
u(y, t) = [2B cos y cos σt + 2B sin y sin σt] exp − y (9.40)
2ν 2ν 2ν
r r r
σ σ σ
u(y, t) = V (cos y sin σt + sin y sin σt) exp − y (9.41)
2ν 2ν 2ν
Using cos(α + β) = cos α cos β − sin α sin β, the solution can be written in the final form
r r
σ σ
u(y, t) = V cos σt − y exp − y . (9.42)
2ν 2ν
107
G(t) = G cos σt (9.43)
while requiring the non-slip condition to be satisfied on the surface of the cylinder r = a.
∂u ∂ 2 u 1 ∂u
= G cos σt + ν + (9.44)
∂t ∂r 2 r ∂r
which leads to
d2 U1 (r) 1 dU1 (r)
−U1 (r)σ sin σt + U2 (r)σ cos σt = G cos σt + ν + cos σt
dr 2 r dr
2
d U2 (r) 1 dU2 (r)
+ν + sin σt (9.46)
dr 2 r dr
d2 U1 (r) dU (r) σ 2 G
r2 2
+r 1 − r U2 (r) = −r 2
dr dr ν ν
2
d U2 (r) dU (r) σ 2
r2 2
+r 2 + r U1 (r) = 0 (9.47)
dr dr ν
which implies
108
d2 U1 (r) dU1 (r) σ 2
r2 + r − r Ũ2 (r) = 0
dr 2 dr ν
2
d Ũ2 (r) dŨ (r) σ 2
r2 2
+r 2 + r U1 (r) = 0 (9.50)
dr dr ν
If we write
d2 z(r) dz(r) σ
r2 2
+r − i r 2 z(r) = 0 (9.52)
dr dr ν
r
1/2 σ
z(r) = CI0 (i r) (9.53)
ν
where I0 is the modified Bessel function of the first kind which satisfies the differential
equation
d2 y(x) dy(x)
x2 2
+ x − x2 y(x) = 0.
dx dr
From I0 (x) = J0 (ix) we have I0 (i1/2 x) = J0 (i3/2 x) (where J0 is the Bessel function of the
first kind). The real and imaginary parts of J0 (i3/2 x) are the so-called Kelvin functions,
r r
σ σ
z(r) = C[ber0 ( r) + ibei0 ( r)]
ν ν
or
r r
σ σ
U1 (r) − iŨ2 (r) = (CR + iCI )[ber0 ( r) + ibei0 ( r)].
ν ν
109
Written out
G
U1 (r) − i(U2 (r) − )
rσ r
σ σ
= (CR + iCI )(ber0 ( r) + i bei0 ( r))
ν ν
r r r r
σ σ σ σ
= CR ber0 ( r) − CI bei0 ( r) + i[CI ber0 ( r) + CR bei0 ( r)]
ν ν ν ν
(9.55)
r r
σ σ
U1 (a) = CR ber0 ( a) − CI bei0 ( a) = 0
ν ν
r r
G σ σ G
U2 (a) − = −CI ber0 ( a) − CR bei0 ( a) = − (9.56)
σ ν ν σ
r r
σ σ
CR = A bei0 ( a), CI = A ber0 ( a)
ν ν
r r
2 σ 2 σ G
A ber0 ( a) + bei0 ( a) = .
ν ν σ
Therefore we have:
pσ pσ p p
G bei0 ( ν
a)ber0 ( r) − ber0 ( σν a)bei0 ( σν r)
U1 (r) = pν p (9.57)
σ ber02 ( σν a) + bei20 ( σν a)
and
pσ p p p
G [ber0 ( ν
r)ber0 ( σν a) + bei0 ( σν r)bei0 ( σν a)] G
U2 (r) = − p p + (9.58)
σ ber02 ( σν a) + bei20 ( σν a) σ
Let us introduce
110
r
σ
α= a (9.59)
ν
G
u(r, t) = [(Bi ber0 (αr/a) − Br bei0 (αr/a)] cos σt +
σ
G
+ [1 − Br ber0 (αr/a) + Bi bei0 (αr/a)] sin σt. (9.60)
σ
" √ ! #
G I0 ( iαr/a) iσt
u(r, t) = Re 1− √ e (9.61)
iσ I0 ( iα)
√ !
G1 I ( iαr/a)
1− 0 √
σ i I0 ( iα)
G1 ber0 (αr/a) + ibei0 (αr/a)
= 1−
σ i ber0 (α) + ibei0 (α)
G1 [ber0 (αr/a) + ibei0 (αr/a)][ber0 (α) − ibei0 (α)]
= 1−
σ i ber02 (α) + bei20 (α)
G1
= (1 − [ber0 (αr/a) + ibei0 (αr/a)][Br − iBi ])
σ i
G1
= (1 − Br ber0 (αr/a) − Bi bei0 (αr/a) + i [−Br bei0 (αr/a) + Bi ber0 (αr/a)])
σ i
G G
= [Bi ber0 (αr/a) − Br bei0 (αr/a)] − i [1 − Br ber0 (αr/a) + Bi bei0 (αr/a)] .
σ σ
(9.62)
111
9.4 Circular Streamlines
Let us consider steady flow between two rotating cylinders. The inner cylinder has radius
r1 and angular velocity Ω1 whereas the outer cylinder has radius r2 and angular velocity
Ω2 .
v = Ω1 r1 on r = r1
v = Ω2 r2 on r = r2 . (9.63)
∂ ∂
~e = −~er , ~e = ~eφ (9.65)
∂φ φ ∂φ r
2 1 ∂ ∂F 1 ∂2F
∇F = r + 2 2.
r ∂r ∂r r ∂φ
∂ 1 ∂
~u · ∇~u = (uφ~eφ ) · (~r + ~eφ )(u ~e )
∂r r ∂φ φ φ
v(r) ∂
= (v(r)~eφ )
r ∂φ
v2
= − ~er (9.66)
r
We inspect the effect of the Laplacian on the vector v~eφ employing (9.65) and using
v = v(r),
1 ∂ ∂ 1 ∂2 ~ = ~e 1 ∂ ∂ 1 ∂
r + 2 2 (v φ) φ r v+ (−v~er )
r ∂r ∂r r ∂φ r ∂r ∂r r 2 ∂φ
1 ∂ ∂ v
= ~eφ r v− ~e
r ∂r ∂r r2 φ
2 1
= ~eφ ∇ − 2 v (9.67)
r
112
(this illustrates the fact that in non-Cartesian coordinates the vector Laplacian resolved
in these non-Cartesian coordinates does not coincide with the Laplacian of a scalar in the
non-Cartesian coordinates).
1 dp
~u · ∇~u = − + ν∇2~u
̺ dr
becomes
v2 1 dp 2 1
− ~er = − ~e + ν~eφ ∇ − 2 v
r ̺ dr r r
1 dp v2
= (9.68)
̺ dr r
and
2 1
∇ − 2 v(r). (9.69)
r
Rewriting (9.69)
2
2 d d
r + r − 1 v(r) = 0
dr 2 dr
(λ(λ − 1) + λ − 1)r λ = 0
B
v(r) = Ar + (9.70)
r
B B
Ω1 r1 = Ar1 + and Ω2 r2 = Ar2 +
r1 r2
113
implying Ω2 r22 − Ω1 r12 = A(r22 − r12 ) and Ω1 r1 (r1 r22 ) − Ω2 r2 (r2 r12 ) = B(r22 − r12 ), so that
Ω2 r22 − Ω1 r12
A =
r22 − r12
(Ω − Ω )r 2 r 2
B = − 2 2 1 2 1 2. (9.71)
r2 − r1
Special cases:
Z 2π
Ω1 r12
Γ= (rdθ) = 2πΩ1 r12 (9.72)
0 r
Consider the situation when r1 → 0 while Ω1 r12 = Const. and invoke Stokes theorem
I Z
~u · d~x = ~ω · ~ndS,
C S
ψ = Axy
114
∂ψ
u= = Ax
∂y
∂ψ
v=− = −Ay
∂x
This will not be a solution for viscous flow because u 6= 0 on the boundary y = 0.
We require the far-field to look like the above solution whereas on the plane y = 0,
u = v = 0. In summary
u = v = 0 on y = 0
ψ = Axy as y → ∞. (9.74)
∂(∇2 ψ, ψ)
= ν∇4 ψ (9.75)
∂(x, y)
ψ = xF (y) (9.76)
∂ψ
u = = xF ′ (y)
∂y
∂ψ
v = − = F (y). (9.77)
∂x
∇2 ψ = xF ′′ ∇4 ψ = xF ′′′′
∂(∇2 ψ, ψ)
= x(F ′′ F ′ − F ′′′ F ). (9.78)
∂(x, y)
115
F ′′ F ′ − F ′′′ F = νF ′′′′
((F ′ )2 − F F ′′ )′ = νF ′′′′
(F ′ )2 − F F ′′ = νF ′′′ + c
(F ′ )2 − F F ′′ = νF ′′′ + A2
This equation has two parameters ν and A. If we apply the following scaling
r
A √
η= y, F (y) = AνG(η)
ν
G(0) = G′ (0) = 0
r √
dF A d AνG)
(∞) = A or (∞) = A
dy ν dη
or
G′ (∞) = 1.
This may be solved numerically. The solution of the equation is now of the form
116
r !
√ A
ψ= AνxG y
ν
117
Chapter 10
Mathematical Theorems
The theorems derived here pertain to the approximations to the Navier-Stokes equation
given by
∇p = µ∇2~u
∇ · ~u = 0.
These are the Stokes flow equations for slow viscous flow of an incompressible fluid.
!
1 ∂ui ∂uj
eij = +
2 ∂xj ∂xi
and that the incompressibilty condition ∇ · ~u = 0 is equivalent to ekk = 0. Also that the
constituative relation then for incompressible flow is
∂σij
=0 (10.1)
∂xj
118
using the continuity equation in component form
∂uj
= 0.
∂xj
∇p = µ∇2~u ∇ · ~u = 0 (10.2)
~ (x) on S
~u = U
(x ∈ S, S may consist of interior as well as exterior boundaries). Let (~u′ , p′ ) also satisfy
(10.2); then
~u ≡ ~u′ in V. (10.3)
P roof
Let
~u˜ = ~u − ~u′
ẽij = eij − e′ij
σ̃ij = σij − σij′ (10.4)
Z Z !
1 ∂ ũi ∂ ũj
ẽij σ̃ij dV = + σ̃ij dV
V V 2 ∂xj ∂xi
Z
∂ ũi
= σ̃ dV (10.5)
V ∂xj ij
119
The last line on the RHS is obtained because of the symmetry of σ̃ij namely σ̃ij = σ̃ji .
Using the alternative form of Stokes flow equation, (10.1), we can write (10.5) as
Z
∂
(ũ σ̃ )dV
V ∂xj i ij
which becomes
Z
ũiσ̃ij nj dS = 0
S
~u˜ = ~u − ~u′ = 0
on S, because both the velocity fields satisfy the boundary conditions. We have thus
shown that (10.5) vanishes.
Now because
and
Z
2µ ẽij ẽij dV
V
which we have shown is zero. Clearly the integrand must be zero, that is
ẽij = 0
~u˜ := ~u − ~u′
120
induces at most a uniform translation or rotation. However since ~u = ~u′ on S it implies
that
~u = ~u′ = 0 in V.
Z
∂Ekin 1 ∂(̺~u · ~u)
= (10.6)
∂t V 2 ∂t
In the absense of any forces an kinetic Energy change within the volume V can occur by
Energy flowing across the boundary,
Z
∂Ekin 1
=− ( ̺~u · ~u)~u · ~ndS (10.7)
∂t S 2
In the pressence of viscous and body forces the rate of change of energy is
Z Z Z
∂Ekin 1
=− ( ̺~u · ~u)~u · ~ndS = ui σij nj dS + ̺ui fi dV (10.8)
∂t S 2 S V
121
Z Z " #
∂E ∂
+ ∇ · (E ~u) dV = (u σ ) + ̺ui fi dV
V ∂t V ∂xj i ij
Z " #
∂ui ∂σij
= σ + ui + ̺ui fi dV
V ∂xj ij ∂xj
Z " #
∂σij
= eij σij + ui + ̺ui fi dV
V ∂xj
Z " #
∂
= eij (−pδij + 2µeij ) + ui (−pδij + 2µeij ) + ̺ui fi dV
V ∂xj
Z h i
= 2µeij eij − ~u · ∇p + µui ∇2 ui + ̺ui fi dV.
V
(10.9)
1∂ 1 ∂ 1 1 ∂̺ 1 1
(̺u2 ) + ∇ · (̺u2~u) = ̺ ( u2 ) + u2 + ̺~u · ∇( u2 ) + u2 ∇ · (̺~u)
2 ∂t 2 ∂t
2 2 ∂t 2 2
∂ 1 2 1 2 1 2 ∂̺
= ̺ ( u ) + ~u · ∇( u ) + u + ∇ · (̺~u)
∂t 2 2 2 ∂t
∂ 1 2 1 2
= ̺ ( u ) + ~u · ∇( u )
∂t 2 2
D 1 2
= ̺ ( u ). (10.10)
Dt 2
As we are assuming an incompressible fluid we can bring the density ̺ inside the sub-
stantial derivative. We obtain an Energy balance equation, Energy is not conserved and
there is a viscous dissipation of Energy on the RHS
D 1 2
( ̺u ) + ~u · ∇p − ̺~u · f~ − µ~u · ∇2~u = 2µeij eij . (10.11)
Dt 2
Z
2µ eij eij dV. (10.12)
V
122
T heorem Let (~u, p) be the unique flow satisfying
∇p = µ∇2~u, ∇ · ~u = 0
~
~u = U on S (10.13)
∇ · ~u′ = 0 and ~
~u′ = U on S (10.14)
then the rate of dissipation of energy is least in the flow satisfying (10.13).
P roof
Let
~u˜ = ~u − ~u′
ẽij = eij − e′ij (10.15)
where the ~u′ is the one defined in this theorem and e′ij corresponds to this ~u′ . We do a
caluculation similar to one done in the previous theorem,
Z Z
∂ ũi
ẽij σij dV = σij dV
V V ∂xj
Z
∂
= (ũi σij )dV
V ∂xj
Z
= ũi σij nj dS = 0. (10.16)
S
As ~u − ~u′ = 0 on S. Substituiting σij = −pδij + 2µeij into the first integral implies
Z
2µ ẽij eij dV = 0. (10.17)
Z Z
2µ e′ij e′ij dV = 2µ (eij − ẽij )2 dV
V
ZV Z Z
= 2µ eij eij dV + 2µ ẽij ẽij dV − 4µ eij ẽij dV (10.18)
V V V
123
Since the last term is zero by (10.17) and ẽij ẽij ≥ 0 we must have
Z Z
2µ e′ij e′ij dV ≥ 2µ eij eij dV (10.19)
V V
which proves the theorem. That is the Stokes flow is the one of minimum dissipation for
given boundary conditions.
Assume ~u is an irrotaional velocity field field described by a velocity potential, φ, and ~u′
is an arbitrary solenoidal rotational velocity field. The velocity field corresponding to an
irrotational flow has a least amount of kinetic energy. We impose the boundary conitions
on these velocity fields
Assuming, the same, constant density ̺ the continuity equation for both these velocity
fields are ∇ · ~u = ∇ · ~u′ = 0. The difference in kinetic energies of the two flows, ∆Ekin =
Ekin (~u′) − Ekin (~u), is
Z
1
∆Ekin = ̺ (~u′ · ~u′ − ~u · ~u)dV
2 V
Z Z
1
= ̺ (~u − ~u) · (~u − ~u) + ̺ (~u′ − ~u) · ~udV.
′ ′
(10.21)
2 V V
Using ~u = ∇φ and the divergence theorem on the last term on the RHS gives
Z Z
′
(~u − ~u) · ∇φdV = ∇ · [(~u′ − ~u)φ]dV
V ZV
= φ(~u′ − ~u) · ~ndS, (10.22)
S
which is zero by the boundary condition (10.20). This implies that the RHS of (10.21)
is positive and therefore the kinetic energy of rotational flow with velocity ~u′ is greater
than the kinetic energy of the corresponding irrotational flow with velocity ~u.
124
Chapter 11
Stoke’s Streamfunction
∇ · ~u = 0
implies
~
~u = ∇ × A (11.1)
~u = uρ~eρ + uz~ez
~ = Ψ(ρ, z) ~e
A (11.2)
ϕ
ρ
125
~u = ∇ × (Aρ~eρ + Aϕ~eϕ + Az ~ez )
~eρ ρ~eϕ ~ez
1 ∂ ∂ ∂
= ∂r ∂ϕ ∂z
ρ
Aρ ρAϕ Az
~eρ ρ~eϕ ~ez
1 ∂ ∂ ∂
= ∂ρ ∂ϕ ∂z
ρ
0 Ψ(ρ, z) 0
1 ∂Ψ 1 ∂Ψ
= − ~eρ + ~e (11.3)
ρ ∂z ρ ∂ρ z
1 ∂Ψ 1 ∂Ψ
ur = − , uz = (11.4)
ρ ∂z ρ ∂ρ
d~x
= λ~u
ds
∂Ψ ∂Ψ
dΨ = dρ + dz
∂r ∂z
= ρuz dρ − ρuρ dz
= ρuz λuρ ds − ρuρ λuz ds = 0. (11.5)
By using the formula for the curl in cylindrical polar coordinates (an example of an
orthogonal curvilinear coordinate system)
~ = ∇(∇ · A)
∇ × ~u = ∇ × (∇ × A) ~ − ∇2 A.
~
~
∇ × ~u = −∇2 A
126
∇ × ~u = ∇ × (uρ~eρ + uθ~eθ + uϕ~eϕ )
~eρ ρ~eϕ ~ez
1 ∂ ∂ ∂
= ∂ρ ∂ϕ ∂z
ρ
uρ ρuϕ uz
~eρ ρ~eϕ ~ez
1 ∂ ∂ ∂
= ∂ρ ∂ϕ ∂z
ρ
− ρ1 ∂Ψ
∂z
0 ρ1 ∂Ψ ∂ρ
∂ 1 ∂Ψ 1 ∂2Ψ
= − − ~eϕ
∂ρ ρ ∂ρ ρ ∂z 2
1 ∂ 2 Ψ 1 ∂Ψ ∂ 2 Ψ
= − − + ~eϕ (11.6)
ρ ∂ρ2 ρ ∂ρ ∂z 2
1
∇ × ~u = − D 2 Ψ~eϕ . (11.7)
ρ
hence
∂2 1 ∂ ∂2
D2 = − + (11.8)
∂ρ2 ρ ∂ρ ∂z 2
That ∇p = −µ∇2~u and from the identity ∇×(∇p) = 0, we have that ∇×(∇×(∇×~u)) = 0
but
where we have used the identity ∇ · (∇ × ~u) = 0. We have already deomstrated that
~ = ∇ × (∇ × Ψ (−D 2 Ψ)
∇ × (∇ × A) ~eϕ ) = ~eϕ .
ρ ρ
(−D 2 Ψ) (−D 4 Ψ)
∇ × (∇ × (∇ × ~u)) = ∇ × (∇ × ~eϕ ) = ~eϕ .
ρ ρ
127
The streamfucntion equation is then
D 4 Ψ = 0. (11.10)
Pressure
From ∇ × (∇ × ~u) = ∇(∇ · ~u) − ∇2~u and using ∇ · ~u = 0 we can rewrite ∇p = −µ∇2~u as
The usual formula for the grad operator in cylindrical polar coordinate basis (here no φ
component)
∂p ∂p
∇p = ~eρ + ~ez (11.12)
∂ρ ∂z
and
1
∇ × (∇ × ~u) = ∇ × (− D 2 Ψ~eϕ )
r
~eρ ρ~eϕ ~ez
1 ∂ ∂ ∂
=
ρ ∂ρ ∂ϕ ∂z
0 −D 2 Ψ 0
1 ∂ 2 ∂ 2
= − (−D Ψ)~eρ + (−D Ψ)~ez (11.13)
ρ ∂z ∂ρ
Using (11.11), (11.12) and (11.13) we obtain equations for the pressure in terms of the
stream function,
∂p ∂ −D 2 Ψ
=µ ( ) (11.14)
∂ρ ∂z ρ
and
∂p 1 ∂ −D 2 Ψ
= −µ ρ( ). (11.15)
∂z ρ ∂ρ ρ
128
11.1.2 Streamfunction in Spherical Polar Coordinates
~u = ur~er + uθ~eθ
~ = Ψ(r, θ) ~e
A (11.16)
r sin θ ϕ
~ we find
Computing ∇ × A
1 ∂Ψ 1 ∂Ψ
ur = , uθ = − (11.18)
r2 sin θ ∂θ r sin θ ∂r
d~x
= λ~u
ds
∂Ψ ∂Ψ
dΨ = dr + dθ
∂r ∂θ
= −r sin θuθ dr + r 2 sin θur dθ
= −r sin θuθ λur ds + r sin θur λuθ ds = 0. (11.19)
129
Therefore Ψ = Const. on streamlines (again stream surfaces as it is 3D flow).
1
∇ × ~u = − D 2 Ψ~eϕ . (11.21)
r sin θ
hence
∂2
2 sin θ ∂ 1 ∂
D = 2+ 2 (11.22)
∂r r ∂θ sin θ ∂θ
D 4 Ψ = 0. (11.23)
Pressure
Again from ∇×(∇×~u) = ∇(∇·~u)−∇2~u and using ∇·~u = 0 we can rewrite ∇p = −µ∇2~u
as
The usual formula for the grad operator in spherical polar coordinate basis (here no ϕ
component)
130
∂p 1 ∂p
∇p = ~er + ~e (11.25)
∂r r ∂θ θ
and
1
∇ × (∇ × ~u) = ∇ × (− D 2 Ψ~eϕ )
r sin2 θ
~er r~eθ r sin θ~eϕ
1 ∂ ∂ ∂
= ∂r ∂θ ∂ϕ
r 2 sin θ
0 0 − sin1 θ D 2 Ψ
1 ∂ 1 2 ∂ 1 2
= 2 − D Ψ~er + r D Ψ~eθ (11.26)
r sin θ ∂θ sin θ ∂r sin θ
Using (11.24), (11.25) and (11.26) we obtain equations for the pressure in terms of the
stream function,
∂p 1 ∂ −D 2 Ψ
= −µ ( ) sin2 θ (11.27)
∂r r sin θ ∂θ r sin2 θ
and
1 ∂p 1 ∂ −D 2 Ψ
=µ r( ) sin θ. (11.28)
r ∂θ r ∂r r sin2 θ
Suppose the rigid boundary OA is scraped along the plane OB at a constant angle α with
velocity V Relative to O the flow is steady.
2
4 1 ∂ ∂ 1 ∂2
∇ ψ= r + 2 2 ψ=0
r ∂r ∂r r ∂θ
1 ∂ψ
ur = = −V on θ = 0 (11.29)
r ∂θ
and
131
1 ∂ψ
ur = = 0 on θ = α. (11.30)
r ∂θ
From the first of the ur boundary conditions it is clear that the solution is of the form
ψ = rV f (θ) (11.31)
where
Now we have
2 ∂
1 ∂ 1 ∂2
∇ ψ = r + 2 2 rV f (θ)
r ∂r
∂r r ∂θ
V V F (θ)
= (f + f ′′ ) =: (11.34)
r r
and
4 1 ∂ ∂ 1 ∂2 V F (θ)
∇ψ = r + 2 2
r ∂r ∂r r ∂θ r
V
= 3
(F + F ′′ ) = 0. (11.35)
r
F = Ã cos θ + B̃ sin θ
f + f ′′ = Ã cos θ + B̃ sin θ
132
The remainder of the solution is found from considering f + f ′′ = 0, this as we already
know is of the form A cos θ + B sin θ. This leads to the general solution
Multiplying the first by α sin α and the second by α cos α + sin α gives
(−α2 sin2 α)C + (α cos α + sin α)(α sin α)D = α cos α sin α
(α2 cos2 α − sin2 α)C + (α sin α)(α cos α + sin α)D = (α cos α sin α + sin2 α)
− sin2 α
C= .
sin2 α − α2
Substituting this into the second equation of (11.37) and dividing both sides by − sin α
gives
From which we easily get the expression for D. We easily find the expression for B by
substituting the expression for C into B = −C − 1. Altogather we have,
133
α2 − sin2 α cos α sin α − α
B= C= D=
sin2 α − α2 sin2 α − α2 sin2 α − α2
Simplifying we obtain
Here we present the classic solution of the Stokes equations representing uniform motion
of a sphere of radius a in an infinitge expanse of fluid.
Z
Fi = σij nj dS (11.39)
S r=a
134
The net force FD on a sphere is in the z−direction and given by F~ · ~z . Using have
~er · ~ez = cos θ and ~eθ · ~ez = − sin θ we have
Z 2π Z π
2
FD = a [σrr cos θ − σrθ sin θ] sin θdθ (11.40)
0 0
At infinity
ur = 0, uθ = 0 (11.42)
Using separation of variables and from the behaviour at infinity we try the solution of the
form Ψ(r, θ) = f (r) sin2 θ. Calculating D 2 Ψ first,
2 2 ∂2 sin θ ∂ 1 ∂
D [f (r) sin θ] = + 2 f (r) sin2 θ
∂r 2 r ∂θ sin θ ∂θ
2 ∂2 f (r) ∂ 1 ∂
= sin θ 2 f (r) + 2 sin θ sin2 θ
∂r r ∂θ sin θ ∂θ
∂2 f (r)
= sin2 θ 2 f (r) − sin2 θ 2
2 ∂r r
d 2
= 2
− 2 f (r) sin2 θ. (11.44)
dr r
135
2
4 2 2 d2 2
D [f (r) sin θ] = sin θ − f (r) (11.45)
dr 2 r 2
2
d2 2
2
− 2 f (r) = 0. (11.46)
dr r
A
f (r) = + Br + Cr 2 + Dr 4 . (11.47)
r
We can express the velocity components in terms of Ψ via (11.4). Using the above
boundary conditions on the velocity components allows us to determine the constants
A, B, C and D.
1 ∂Ψ
ur =
r2
sin θ∂θ
2 cos θ A 2 4
= + Br + Cr + Dr
r2 r
2A 2B 2
= cos θ + + 2C + 2Dr (11.48)
r3 r
1 ∂Ψ
uθ = −
r sin θ∂r
A B 2
= − sin θ − 3 + + 2C + 4Dr (11.49)
r r
2A 2B
ur |r=a = cos θ + +U =0
a3 a
A B
uθ |r=a = − sin θ − 3 + + U = 0. (11.50)
a a
136
Solving for A and B gives
1 3
A = Ua3 , B = − Ua.
4 4
1 1 a3 3
Ψ = U( r 2 + − ar) sin2 θ (11.51)
2 4 r 4
Velocities
1 a 3 3 a
ur = U cos θ 1 + − (11.52)
2 r 2 r
and
1 a 3 3 a
uθ = −U sin θ 1 − − . (11.53)
4 r 4 r
Expressions for the rate of strain tensor in spherical polar coordinates are given in the
appendix. We will need the components ǫrr and ǫrθ .
∂ur
ǫrr = (11.54)
∂r
∂ 1 a 3 3 a
ǫrr = U cos θ 1+ −
∂r 2 r 2 r
3
U −3a a
= cos θ 4
+3 2 (11.55)
2 r r
r ∂ uθ 1 ∂ur
ǫrθ = + (11.56)
2 ∂r r 2r ∂θ
137
r ∂ 1 1 a3 3 a 1
ǫrθ = −U sin θ − 4
− 2 − U sin θ ur
2 ∂r r 4 r 4r 2r
3
3
U sin θ 1 a 3a 1 1a 3a
= − − + 4+ 2 + + −
2 r r 2r r 2 r4 2 r2
3U sin θ a3
= − (11.57)
4 r4
11.3.3 Pressure
Differential equations for the pressure p come from (11.27) and (11.28). We need to
2
calculate rDsin[Ψ]
2 θ from (11.51). Using (11.44) we can write
D 2 [Ψ] 1 d2 2
2 = − f (r)
r sin θ r dr 2 r 2
2
U d 2 1 2 1 a3 3
= − ( r + − ar)
r dr 2 r 2 2 4 r 4
3 3
U 1a 1a 3a
= (1 + ) − (1 + − )
r 2 r3 2 r3 2 r
3a
= U 2 (11.58)
2r
∂p 1 ∂ 3a
= −µ (−U 2 ) sin2 θ
∂r r sin θ ∂θ 2r
1
= 3µUa 3 cos θ. (11.59)
r
This implies
3 cos θ
p = − µUa 2 + g(θ) (11.60)
2 r
1 ∂p 1 ∂ 3a
= µ r(−U 2 ) sin θ.
r ∂θ r ∂r 2r
3 1
= µUa 3 sin θ. (11.61)
2 r
138
This implies
3 cos θ
p = − µUa 2 + h(r) (11.62)
2 r
where h(r) is an arbitray function of r. Comparing (11.60) and (11.62) noting that we
must have that g(θ) = h(r) = Const. = p∞ we obtain the final result
3 cos θ
p = − µUa 2 + p∞ (11.63)
2 r
−3a3 a
σrr = −p + µU cos θ 4
+3 2 (11.67)
r r
3 a3
σrθ = −µU sin θ (11.68)
2 r4
Z 2π Z π
2
FD = a [σrr cos θ − σrθ sin θ] sin θdθ (11.69)
0 0
Z π
2 3 cos θ −3a3 a
FD = 2πa µUa 2 − p∞ + µU cos θ +3 2 cos θ
0 2 r r4 r r=a
3 a3
+ µU sin θ 4 sin θ sin θdθ (11.70)
2 r r=a
139
The p∞ term vanishes because of the integral
Z π Z −1
−1
u2
cos θ sin θdθ = − udu = = 0.
0 1 2 1
Z π Z π
2
FD = 3πµaU cos θ sin θdθ + 3πµaU sin3 θdθ (11.71)
0 0
| {z } | {z }
pressure viscous
Z π Z −1
2
−1
2 u3 2
cos θ sin θdθ = − u du = =
0 1 3 1 3
Z π Z π
3 2 4
sin θdθ = sin θ(1 − cos2 θ)dθ = [− cos θ]π0 − = .
0 0 3 3
That is, one third of the drag is due to pressure forces, two-thirds are due to viscous
forces.
!
∂ui ∂uj
σij = −pδij + µ +
∂xj ∂xi
140
!
∂σij ∂ ∂ 2 ui ∂ ∂uj
=− p+µ +
∂xj ∂xi ∂x2j ∂xi ∂xj
∂uj
From the continuity equation ∂xj
= 0, so
∂σij ∂ ∂2u
=− p + µ 2i
∂xj ∂xi ∂xj
∂σij
= 0.
∂xj
By Gauss theorem
Z Z Z
∂σij
σij nj dS + σij nj dS = dV = 0.
S r=a S r=∞ V ∂xj
Implying that
Z
Fi = − σij nj dS. (11.74)
S r=∞
141
Appendix A
Glossary
• Body force: Force acting on an infintesimal volume of fluid per unit mass.
• Boundary layer flow: There will be a layer close to the solid wall where the viscous
terms are important even for very high Reynolds number flow. The non-slip condition
holds at the surface while a height above the fluid motion is uneffected. This layer will
be very thin and the flow in that region is called the boundary layer flow.
∂ui ∂σij
• Cauchy’s equation: ̺ + (~u · ∇)ui = ̺fi + where σij is the stress tensor.
∂t ∂xj
p∞ − pv
σ= 1
2
̺V 2
is less than the critical cavitation number σcrit , which depends on the geometry and the
Reynolds number.
∂̺
• Continuity equation: The continuity equation is + ∇ · (̺~u) = 0 where ̺ is the density
∂t
and ~u is the fluid velocity. The continuity equation expressing the conservation of mass.
For constant density, or incompressible flow the continuity equation reduces to ∇ · ~u = 0.
• Density: Denoted ̺ here, is the mass of fluid per unit volume. It can be a function fo
space and time, but oftent the assumption that it is constant is made.
• Euler equations of motion: For flows with high Reynolds number R, where the viscous
effects are small, most of the flow can be considered inviscid and a simpler set of equations
142
can be solved, which corresponds to the R → ∞ limit in the Navier-Stokes equations.
∂
They read ρ ~u + (~u · ∇)~u = −∇p + ρf~.
∂t
R H
• Guass Divergence Theorem: V ∇ · AdV ~ = SA ~ · ~ndS where V is a volume, S is the
surface bounding the volume V , and ~n is an outward pointing vector normal to the surface
S.
• Ideal (or perfect) fluid: This is a fluid which can exert no shearing stress across any
surface σij = −pδij that is the stress tensor for a perfect fluid is diagonal. This is done
by ignoring viscous effects.
• Incompressible fluid is a fluid which always has constant density. The continuity equa-
tion is then ∇ · ~u = 0.
• Kinematic viscocity: The kinematic viscocity is the ratio of viscocity and the density:
ν = µ/ρ.
• Lagrangian description:
∂~u
• Navier-Stokes equation: The Navier-Stokes equation is ρ +(~u·∇)~u = −∇p+µ∇2~u+ρf~
∂t
together with the continuity equation
∇ ·~u = 0. Another form of the Navier-Stokes
∂~u 1 2 p
equation is − (~u × ω) = −∇ |~u| + ~ + f~ where ~ω is the vorticity.
− ν∇ × ω
∂t 2 ρ
• Non-slip condition: Dictates that the speed of the fluid at a solid boundary is zero
relative to the boundary.
• Newtonian fluid:
• Slender body theory: Methodology that can be used to take advantage of the slenderness
of a body to obtain an approximation to the fluid field surrounding it and/or the net effect
of the field on the body.
• Stagnation point: A point in the flow field where the local velocity of the fluid is zero.
143
• Stokes equation: For Stoke’s flow the Navier-Stokes equation reduces to Stoke’s equation
∇p = µ∇2~u.
• Stokes flow: A type of flow where advective inertial forces are small compared with
viscous forces. The Reynolds number is low.
•R Stokes Theorem:
H Stayes that for any surface S encloed by the the closed curve C
∇×A ~= A~ · d~l holds.
S C
D ∂
• Substantial derivative: = + (~u · ∇)
Dt ∂t
• Surface forces:
• Vorticity: ~ω = ∇ × ~u
144
Appendix B
which means
a1 a2 a3
~a · (~a × ~b) = a1 a2 a3 = 0.
b1 b2 b3
~a × ~b = −~b × ~a.
145
We consider the length of a cross product. We can simplify things by writing ~a = |~a|~ex
and ~b = |~b| cos θ~ex + |~b| sin θ~ey , then
Thus ~a × ~b has length |~a||~b| sin θ and a direction ~ez perpedicular to the plane containing
~a and ~b.
∂φ ∂φ ∂φ
dφ = dx + dy + dz
∂x ∂y ∂z
As the the LHS is a scalr and dx, dy and dz are components of a vector the three numbers
∂φ ∂φ ∂φ
, ,
∂x ∂y ∂z
∂φ ∂φ ∂φ
∇φ = ~ex + ~ey + ~ez
∂x ∂y ∂z
We call ∇φ the gradient of a scalar. Now the argument that ∇φ is a vector does not depend
on what scalar field we differentiated. As the trasformation as a vector is independent of
what scalar field we are differentiatinf we could just as well omit the φ.
∂ ∂ ∂
∇ = ~ex + ~ey + ~ez (B.3)
∂x ∂y ∂z
146
B.2.2 Divergence
∇ · ~v = ∇x vx + ∇y vy + ∇z vz
∂vx ∂vy ∂vz
= + + (B.4)
∂x ∂y ∂z
~v · ∇
This is a scalar operator and does appear in fluid mechanics, in particular the advective
term of the substantial derivative.
B.2.3 Laplacian
∇ · (∇φ) = (∇ · ∇)φ = ∇2 φ.
The ∇2 is a scalar operator that often appears in physics and is called the Laplacian,
2 ∂2 ∂2 ∂2
∇ = 2+ 2+ 2 (B.5)
∂x ∂y ∂z
How about acting ∇2 on a vector (we will at times need to do this as this terms appears
in the /navier-Stokes equations and the Stokes flow equations). In Cartesian coordinates
it is straightforward. For example the x-component of ∇2~v is simply
2 ∂2 ∂2 ∂2
(∇ ~v )x = + + vx = ∇2 vx . (B.6)
∂x2 ∂y 2 ∂z 2
The situation is not so simply in coordinates systems other than Cartesian coordinates.
We will later derive formula for spherical polar and cylindrical polar coordinates.
147
B.2.4 Curl
There is another object we can come uo with from the operator ∇ and a vector field ~v
using the cross product. The curl denoted ∇ × ~v can be defined via the cross product,
∂
[∇ × ~v]i = ǫijk v (B.8)
∂xj k
B.2.5 Identities
∂ ∂
∇ × ∇φ = ǫijk φ=0
∂xj ∂xk
∇ × ∇φ = 0. (B.9)
∂ ∂
∇ · (∇ × ~v ) = ǫijk ~v = 0
∂xi ∂xj k
∇ · (∇ × ~v) = 0. (B.10)
∇ × (∇ × ~v )
148
[∇ × (∇ × ~v )]i = ǫijk ∂j [∇ × ~v ]k
= ǫijk ∂j (ǫki′ j ′ ∂i′ vj ′ )
= ǫijk ǫki′ j ′ ∂j ∂i′ vj ′
= (δii′ δjj ′ − δij ′ δji′ )∂j ∂i′ vj ′
= ∂i (∂j vj ) − ∂j ∂j vi . (B.12)
The first type of integral will need to introduce is a line integral of a scalar function f ,
denoted
Z Q
f dl (B.14)
P
|{z}
along Γ
X
fn ∆ln (B.15)
n
where fn is the value of the fucntion at the nth segment. The integral (B.14) comes from
increasing the number of segments so that the largest ∆ln → 0
The particular type of line integral that we will be interested in is of the form
Z Q
~v · d~l (B.16)
P
|{z}
along Γ
149
Q
∆ln
∆l3
∆l1
P ∆l2
Figure B.1: .
where ~v is a vector field and d~l is an infintesimal displacement vector along the line Γ.
Hence it is of the form (B.14) where instead of f we have another scalar - the component
of ~v in the direction of d~l,
where ~t is the unit tangent vector to the curve (~t is the unit vector parallel to d~l). The
integral (B.16) is the sum of such terms.
~v
Q
d~l
P
Figure B.2: .
Theorem: If Γ is any curve joining point P and point Q, the following is true,
Z Q
ψ(Q) − ψ(P ) = (∇ψ) · d~l. (B.17)
P
|{z}
any curve f rom
P to Q
150
B.3.3 Definition of Flux and Statement Gauss’s Divergence The-
orem
Given a surface S with outward unit-normal vector field ~n and a vector field ~u, the flux
is defined as
Z
~u · ~ndS.
S
Gauss’s theorem pertains to a closed surface S that encloses a volume V , and relates the
flux through the surface to the behaviour of a vector field inside the surface.
Z I
~
∇ · AdV = ~ · ~ndS
A (B.18)
V S
Stokes’ theorem Let C be any simple closed path (i.e.,a path that starts and ends at
the same point and has no intersections), and consider any surface S of which C is the
boundary.Then stokes’theorem says that
Z I
~ · ~ndS =
∇×A ~ · d~l
A (B.19)
S C
A conservative force is a force with the property that the work done, W , in moving a
particle between two points
Z Q
W = F~ · d~l
P
151
is independent of the path taken. Equivalently, if a particle travels around a closed loop,
the net work done by a conservative force is zero. If a force F~ can be written in the form
F~ = −∇Φ
where Φ is a scalar field, the potential, the above theorem guarantees the force is conser-
vative. An obvious example of a conservative force is the gravitational force.
We now prove the theorem staed in section B.3.2. Consider the curve Γ split up into
segments as in fig (B1) where
∂ψ ∂ψ ∂ψ
(∇ψ) · ∆~ln = ∆xn + ∆yn + ∆z
∂x ∂y ∂y n
= (∆ψ)n
= ψ(~pn+1 ) − ψ(~pn ). (B.20)
Now with ψ(~p0 ) = ψ(P ) and ψ(~pN +1 ) = ψ(Q), the approximation to the line integral is
the summation
N
X
(∇ψ) · ∆~ln = [ψ(~p1 ) − ψ(~p0 )] + [ψ(~p2 ) − ψ(~p1 )] + [ψ(~p3 ) − ψ(~p2 )] + · · · +
n=0
+ · · · + [ψ(~pN ) − ψ(~pN −1 )] + [ψ(~pN +1 ) − ψ(~pN )]
= ψ(~pN +1 ) − ψ(~p0 )
= ψ(Q) − ψ(P ). (B.21)
where all but two terms cancel. Taking the continuum limit does not change the last line
on the RHS. Also it is obvious that the answer will be the same regardless which curve
we use between P and Q.
152
B.5 Proof of Gauss’s Divergence Theorem
Say we have a closed surface S that encloses a volume V . Separate the volume into two
parts by taking a slice resulting it two closed surfaces and volumes as in fig .. The volume
V1 is enclosed by the closed surface S1 , which is made up of part of the original surface S1′
and of the surface of the slice S12 . The volume V2 is enclosed by S2 , which is made up of
the rest of the original surface S2′ and the surface of the slice S12 . The flux through S1 is
~u
S1′ S2′
~n
~n2 ~n1
V1 S12 V2
Figure B.3: .
We then have S1 = S1′ + S12 and S2 = S2′ + S12 . The flux through S1 can be written as
the sum of two parts
Z Z
~ · ~ndS +
A ~ · ~n dS
A (B.22)
1
S1′ S12
Z Z
~ · ~ndS +
A ~ · ~n dS
A (B.23)
2
S2′ S12
Z Z
~ · ~ndS +
A ~ · ~ndS
A (B.24)
S1′ S2′
which is just to the flux through the original surface S = S1 + S2 . We can subdivide
the volume again and again and it will generally be true that the flux through the outer
surface will be equal to the sum of the fluxes out of all the smaller interior pieces.
We now consider the special case of the flux out of a small cube.
153
∂vx ∂v
vx + dx dydz − vx dydz = x dxdydz (B.25)
∂x ∂x
Similar contributions come from the other two pairs of faces, adding together their con-
tributions the total flux through all faces is
Z
∂v1 ∂v2 ∂v3
~v · ~ndS = + + dxdydz (B.26)
cube ∂x ∂y ∂z
or
Z
~v · ~ndS = (∇ · ~v )dxdydz (B.27)
cube
Splitting the volume V enclosed by a closed surface S into infintesimally small cubes,
summing the LHS of the above equation gives the total flux out of the closed surface and
summing over the RHS gives the volume integral
Z
∇ · ~vdV
V
resulting in
Z Z
~v · ~ndS = ∇ · ~v dV (B.28)
S V
I
∂vy ∂v
~v · d~l = vx dx + (vy + )dy − (vx + x )dx − vy dy
∂x ∂y
∂vy ∂v
= − x dxdy (B.29)
∂x ∂y
Now say we had a collection of such squares as in fig (B.2) and wished to add up the
circulation from each individual square. Interior paths are transversed in opposite di-
rections, thus their contributions to each line integral cancel pairwise. Therefore only
154
y
vy
dy v
v
(x, y) vx
dx
the outside edge contributes. This observation is an underlying principle in the proof of
Stoke’s theorem.
Figure B.5: When adding together the circulation of each loop the only remaing contri-
bution to the line integral comes from the outside edge.
(∇ × ~v)z dS (B.30)
Here the z-component is the normal to the surface. We can therefore write the circulation
around a infintesimal square in an invariant vector form,
I
~v · d~l = (∇ × ~v ) · ~ndS (B.31)
C
So we have that the circulation of any vector ~v around an infintesimal square is the
component of the curl of ~v normal to the surface, times the area of the square. This result
is independent on the orientation of the square.
155
Γ
Figure B.6: We have some surface bounded by the loop Γ. The surface is divided into
many small areas, each approximately a square.
Now suppose that we have a loop which is the boundary of some surface. There are
of course an infinite number of surfaces that have all have this loop as their boundary.
However the result does not depend on the particular surface choosen. Let the choosen
surface be divided into many small loops. If we take the loops small enough, we can
assume that each of the small loops enclose an area which is essentially flat. Also we
can choose our small loops so that each is very nearly a square. Combining (B.31) with
the fact that when you add up the circulation of each individual loop the only remaining
contribution comes from the outside edge, we have
Z I
∇ × ~v · ~ndS = ~v · d~l. (B.32)
S Γ
156
Appendix C
Curvilnear coordinates are a coordinate system for Euclidean space in which the coor-
dinate lines may be curved. We have already met such coordinates - cylindrical and
spherical polar coordinates.
We will use Guass’s theroem and Stoke’s theorem to find formula for the Divergence and
Curl respectively in orthogonal curvilinear coordinates. We have already derived formula
for the Gradient in the main text,combining this with formula for the Divergence we will
obtain a formula for the Laplacian of a scalar field.
The vector Laplacian and the advective term, both of which appear in the Navier-Stokes
equations, will be resolved in circular cylindrical and spherical polar coordinates respec-
tively in appendices D and E.
We have already derived formula for the Gradient in the main text in section 3.1. The
formuala for the grad of a scalar is
3
X 1 ∂φ
∇φ = ~ei . (C.1)
i=1
hi ∂qi
The hi is defined by
∂~r
hi =
∂qi
157
In cylindrical coordinates we have
q1 = ρ, q2 = ϕ, q3 = z
h1 = 1, h2 = ρ, h3 = 1 (C.2)
q1 = r, q2 = θ, q3 = ϕ
h1 = 1, h2 = r, h3 = r sin θ. (C.3)
From these examples it is clear that hi dqi (no summation over i implied) is the distance
change brought about by a coordinate change dqi .
We can use the Gauss’s Divergence theorem to find a formula for the divergence.
R
~
~v · dS
∇ · ~v (q1 , q2 , q3 ) = lim R (C.4)
dV →0 dV
Define
∂ ∂
v1 h2 h3 + (v1 h2 h3 )dq1 dq2 dq3 − v1 h2 h3 dq2 dq3 = (v h h )dq dq dq . (C.5)
∂q1 ∂q1 1 2 3 1 2 3
Adding in similar results for the other two pair of surfaces, we obtain
158
~e3
~e2 h3 dq3
h2 dq2
~e1 h1 dq1
Z
~ ∂(v1 h2 h3 ) ∂(v2 h1 h3 ) ∂(v3 h1 h2 )
~v (q1 , q2 , q3 ) · dS = + + dq1 dq2 dq3 . (C.6)
∂q1 ∂q2 ∂q3
1 ∂(v1 h2 h3 ) ∂(v2 h1 h3 ) ∂(v3 h1 h2 )
∇ · ~v (q1 , q2 , q3 ) = + + (C.7)
h1 h2 h3 ∂q1 ∂q2 ∂q3
Combining (C.1) and (C.7) we obtain the formula for the Laplacian of a scalar field,
2 1 ∂ h2 h3 ∂φ ∂ h1 h3 ∂φ ∂φ h1 h2 ∂φ
∇ φ(q1 , q2 , q3 ) = + +
h1 h2 h3 ∂q1 h1 ∂q1 ∂q2 h2 ∂q2 ∂q3 h3 ∂q3
(C.8)
Z I
~=
∇ × ~v · dS ~v · d~l
C
159
h3 dq3
h2 dq2
~e1
∂
~e1 · ∇ × ~v dq2 dq3 h2 h3 = h3 v3 + (h v )dq − h3 v3 dq3
∂q2 3 3 2
∂
− h2 v2 + (h v )dq − h2 v2 dq2 (C.9)
∂q3 2 2 3
and so
1 ∂ ∂
~e1 · ∇ × ~v = (h v ) − (h v ) (C.10)
h2 h3 ∂q2 3 3 ∂q3 2 2
Similar results for the other two rectangles, those orthogonal to ~e2 and ~e3 , we obtain
1 ∂ ∂
∇ × ~v = (h v ) − (h v ) ~e1
h2 h3 ∂q2 3 3 ∂q3 2 2
1 ∂ ∂
+ (h v ) − (h v ) ~e2
h1 h3 ∂q3 1 1 ∂q1 3 3
1 ∂ ∂
+ (h v ) − (h v ) ~e3 (C.11)
h1 h2 ∂q1 3 3 ∂q3 1 1
160
Appendix D
x = ρ cos ϕ
y = ρ sin ϕ
z = z (D.1)
p
ρ = x2 + y 2
ϕ = tan−1 (y/x)
z = z (D.2)
∂~eρ ∂~eϕ
= ~eϕ , = −~eρ . (D.3)
∂ϕ ∂ϕ
∂φ 1 ∂φ ∂φ
∇φ = ~eρ + ~eϕ + ~e (D.4)
∂r ρ ∂ϕ ∂z z
161
2 1 ∂ ∂ 1 ∂2 ∂2
∇ = ρ + 2 2 + 2. (D.5)
ρ ∂ρ ∂ρ ρ ∂ϕ ∂z
1 2 ∂uϕ
∇2~u ρ
= ∇2 u ρ − 2
uρ − 2
ρ ρ ∂ϕ
1 2 ∂uρ
∇2~u ϕ
= ∇2 u ϕ − 2 u ϕ + 2
ρ ρ ∂ϕ
∇2~u z
2
= ∇ uz (D.7)
where we have applied (D.5) to ~u = uρ~eρ + uϕ~eϕ + uz ~ez and used (D.3). These would
be used in writing the µ∇2~u term in the Navier-Stokes equations in cylindrical polar
coordinates.
If ~u is the velocity field and if we had incompressible flow, ∇ · ~u = 0. Then from the
identity ∇ × (∇ × ~u) = ∇(∇ · ~u) − ∇2~u and using ∇ · ~u = 0, then we can use the
alternative formula ∇2~u = −∇ × (∇ × ~u).
The non-linear term (~u · ∇)~u of the Navier-Stokes equations using (D.3) and (D.4) is
∂ 1 ∂ ∂
(~u · ∇)~u = (uρ + uϕ + uz )(uρ~eρ + uϕ~eϕ + uz~ez )
∂ρ ρ ∂ϕ ∂z
!
∂uρ 1 ∂uρ ∂uρ u2ϕ
= uρ + uϕ + uz − ~eρ
∂ρ ρ ∂ϕ ∂z ρ
∂uϕ 1 ∂uϕ ∂uϕ uρ uϕ
+ uρ + uϕ + uz + ~eϕ
∂ρ ρ ∂ϕ ∂z ρ
∂uz 1 ∂uz ∂uz
+ uρ + uϕ + uz ~ez (D.8)
∂ρ ρ ∂ϕ ∂z
Or
162
u2ϕ
(~u · ∇)~u|ρ = (~u · ∇)uρ −
ρ
uρ uϕ
(~u · ∇)~u|ϕ = (~u · ∇)uϕ +
ρ
(~u · ∇)~u|z = (~u · ∇)uz . (D.9)
!
1 ∂ui ∂uj
ǫij = + . (D.10)
2 ∂xj ∂xi
Take
∂ ∂ ∂
∇~u = ~ex + ~ey + ~ez ~u
∂x ∂y ∂z
where
Write
∂ ∂ ∂
∇ ⊗ ~u = ~ex + ~ey + ~ez ⊗ ~u
∂x ∂y ∂z
∂ ∂ ∂
= (ux~ex ) ⊗ ~ex + (ux~ex ) ⊗ ~ey + (ux~ex ) ⊗ ~ez
∂x ∂y ∂z
∂ ∂ ∂
(uy~ey ) ⊗ ~ex + (uy~ey ) ⊗ ~ey + (uy~ey ) ⊗ ~ez
∂x ∂y ∂z
∂ ∂ ∂
(uz~ez ) ⊗ ~ex + (uz~ez ) ⊗ ~ey + (uz ~ez ) ⊗ ~ez
∂x ∂y ∂z
X ∂u
i
= ~ei ⊗ ~ej (D.11)
i,j
∂x j
where we have used that the derivatives do not effect the basis vectors. This gives twice
the second part of (D.10). It can be seen that
163
T ∂ ∂ ∂
(∇ ⊗ ~u) = ~u ⊗ ~ex + ~ey + ~ez
∂x ∂y ∂z
∂ ∂ ∂
= (ux~ex ) ⊗ ~ex + (uy~ey ) ⊗ ~ex + (u ~e ) ⊗ ~ex
∂x ∂x ∂x z z
∂ ∂ ∂
(ux~ex ) ⊗ ~ey + (uy~ey ) ⊗ ~ey + (u ~e ) ⊗ ~ey
∂y ∂y ∂y z z
∂ ∂ ∂
(ux~ex ) ⊗ ~ez + (uy~ey ) ⊗ ~ez + (uz~ez ) ⊗ ~ez
∂z ∂z ∂z
X ∂uj
= ~ei ⊗ ~ej (D.12)
i,j
∂x i
So that
1
ǫ= (∇ ⊗ ~u)T + ∇ ⊗ ~u (D.13)
2
ǫxx ǫxy ǫxz
ǫ = ǫyx ǫyy ǫyz (D.14)
ǫzx ǫzy ǫzz
which are
∂ux ∂uy
ǫxx = ǫyy =
∂x ∂y
∂uz 1 ∂ux ∂uy
ǫzz = ǫxy = +
∂z 2 ∂y ∂x
1 ∂uy ∂uz 1 ∂ux ∂uz
ǫyz = + ǫxz = + (D.15)
2 ∂z ∂y 2 ∂z ∂x
We now consider the rate of strain tensor in cylindrical polar coordinates. Take
T ∂ 1 ∂ ∂
(∇ ⊗ ~u) = ~u ⊗ ~eρ + ~eϕ + ~ez
∂ρ ρ ∂ϕ ∂z
164
where
1 ∂ 1 ∂uρ uρ ∂~eρ
(uρ~eρ ) ⊗ ~eϕ = ~e ⊗ ~eϕ + ⊗ ~eϕ
ρ ∂ϕ ρ ∂ϕ ρ ρ ∂ϕ
1 ∂uρ uρ
= ~e ⊗ ~eϕ + ~e ⊗ ~eϕ (D.16)
ρ ∂ϕ ρ ρ ϕ
and
1 ∂ 1 ∂uϕ uϕ ∂~eϕ
(uϕ~eϕ ) ⊗ ~eϕ = ~e ⊗ ~eϕ + ⊗ ~eϕ
ρ ∂ϕ ρ ∂ϕ ϕ ρ ∂ϕ
1 ∂uϕ uϕ
= ~e ⊗ ~eϕ − ~e ⊗ ~eϕ (D.17)
ρ ∂ϕ ϕ ρ ρ
All the rest are trivial as the derivatives dont effect the basis vectors. For example
∂ ∂uρ 1 ∂ 1 ∂uz
(uρ~eρ ) ⊗ ~eρ = ~e ⊗ ~eρ , (uz~ez ) ⊗ ~eϕ = ~e ⊗ ~eϕ .
∂ρ ∂ρ ρ ρ ∂ϕ ρ ∂ϕ z
∂u
ρ 1 ∂uρ uϕ ∂uρ
∂ρ ρ ∂ϕ
− ρ ∂z
∂u 1 ∂uϕ uρ ∂uϕ
∂ρϕ ρ ∂ϕ
+ ρ ∂z
(D.19)
∂uz 1 ∂uz ∂uz
∂r ρ ∂ϕ ∂z
The rate of strain tensor is obtained by adding the transpose of the velocity gradient
tensor to the velocity gradient tensor and dividing by 2:
1
ǫ= (∇ ⊗ ~u)T + ∇ ⊗ ~u (D.20)
2
165
with components
ǫρρ ǫρϕ ǫρz
ǫϕρ ǫϕϕ ǫϕz (D.21)
ǫzρ ǫzϕ ǫzz
which are
∂uρ 1 ∂uϕ uρ
ǫρρ = ǫϕϕ = +
∂ρ ρ ∂ϕ ρ
∂u ρ ∂ u ϕ 1 ∂uρ
ǫzz = z ǫρϕ = +
∂z 2 ∂ρ ρ 2ρ ∂ϕ
1 1 ∂uz ∂uϕ 1 ∂uρ ∂uz
ǫϕz = + ǫρz = + . (D.22)
2 ρ ∂ϕ ∂z 2 ∂z ∂ρ
166
Appendix E
x = r sin θ cos ϕ
y = r sin θ sin ϕ
z = r cos θ (E.1)
p
r = x2 + y 2 + z 2
θ = cos−1 (z/(x2 + y 2 + z 2 )1/2 )
ϕ = tan−1 (y/x) (E.2)
167
~er = sin θ cos ϕ~ex + sin θ sin ϕ~ey + cos θ~ez
~eθ = cos θ cos ϕ~ex + cos θ sin ϕ~ey − sin θ~ez
~eϕ = − sin θ~er − cos θ~eθ (E.4)
∂φ 1 ∂φ 1 ∂φ
∇φ = ~er + ~eθ + ~e (E.5)
∂r r ∂θ r sin θ ∂ϕ ϕ
1 ∂ 2 1 ∂ 1 ∂uϕ
∇ · ~u = 2
(r ur ) + (sin θuθ ) + (E.6)
r ∂r r sin θ ∂θ r sin θ ∂ϕ
2 1 ∂ 2 ∂ 1 ∂ ∂ 1 ∂2
∇ = 2 r + 2 sin θ + 2 2 (E.7)
r ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂ϕ2
where we have applied (E.7) to ~u = ur~er +uθ~eθ +uϕ~eϕ and used (E.3). These would be used
in writing the µ∇2~u term in the Navier-Stokes equations in spherical polar coordinates.
168
If ~u is the velocity field and if we had incompressible flow, ∇ · ~u = 0. Then from the
identity ∇ × (∇ × ~u) = ∇(∇ · ~u) − ∇2~u and using ∇ · ~u = 0, then we can use the
alternative formula ∇2~u = −∇ × (∇ × ~u).
u2θ + u2ϕ
(~u · ∇)~u|r = (~u · ∇)ur −
r
uu u2ϕ cot θ
(~u · ∇)~u|θ = (~u · ∇)uθ + r θ −
r r
ur uϕ uθ uϕ cot θ
(~u · ∇)~u|ϕ = (~u · ∇)uϕ + − . (E.10)
r r
We now consider the rate of strain tensor in spherical polar coordinates. Take
T ∂ 1 ∂ 1 ∂
(∇ ⊗ ~u) = ~u ⊗ ~er + ~eθ + ~eφ
∂r r ∂θ r sin θ ∂φ
where
1 ∂ 1 ∂ur ur ∂~er
(ur~er ) ⊗ ~eθ = ~e ⊗ ~eθ + ⊗ ~eθ
r ∂θ r ∂θ r r ∂θ
1 ∂ur ur
= ~e ⊗ ~eθ + ~e ⊗ ~eθ (E.11)
r ∂θ r r θ
The rate of strain tensor is obtained by adding the transpose of the velocity gradient
tensor and the velocity gradient tensor and dividing by 2:
169
1
ǫ= (∇ ⊗ ~u)T + ∇ ⊗ ~u (E.13)
2
ǫrr ǫrθ ǫrφ
ǫθr ǫθθ ǫθφ (E.14)
ǫφr ǫφθ ǫφφ
which are
∂ur 1 ∂uθ ur
ǫrr = ǫθθ = +
∂r r ∂r r
1 ∂uφ ur uθ cot θ r ∂ uθ 1 ∂ur
ǫφφ = + + ǫrθ = +
r sin θ ∂φ r r 2 ∂r r 2r ∂θ
sin θ ∂ uφ 1 ∂uθ 1 ∂ur r ∂ uφ
ǫθφ = + ǫφr = + .(E.15)
2r ∂θ sin θ 2r sin θ ∂φ 2r sin θ ∂φ 2 ∂r r
ǫϕϕ
ǫθθ
ǫrr
170