Solutions PDF
Solutions PDF
We substitute this expression for ez into the second trajectory equation, this gives
dy e2t
=y z
dt e − 1 + e2t
0
y t
e2τ
Z Z
dη
⇔ = dτ
y0 η 0 ez0 − 1 + e2τ
e2t
d(e2τ )
y Z
1
⇔ log =
y0 2 1 ez0 − 1 + e2τ
ez0 − 1 + e2t
y 1
⇔ log = log
y0 2 ez0
ez0 − 1 + e2t 1/2
⇔ y = y0 .
ez0
Similarly we can show that
ez0 − 1 + e2t 1/2
x = x0 .
ez0
If x0 = 1, y0 = 1 and z0 = 0, then the parametric equations for the particle trajectory
are x = et , y = et and z = 2t.
To find the streamlines, we fix t, and solve the system of equations
dx dy dz
= (e2t ) xe−z , = (e2t ) ye−z , and = (2e2t ) e−z .
ds ds ds
Solving the third equation we get ez = ez0 + 2e2t s. Substituting this into the second
streamline equation implies
dy e2t
= z y
ds e 0 + 2e2t s
Z y Z s
dη 1
⇔ = e2t z0 + 2e2t σ
dσ
y0 η 0 e
ez0 + 2e2t s
y 1
⇔ log = log
y0 2 ez0
1/2
⇔ y = y0 1 + 2e2t−z0 s) .
1/2
Similarly we find x = x0 1 + 2e2t−z0 s) .
Introductory fluid mechanics: solutions 3
where we have assumed that at time t0 the particle is at position (x0 , y0 , z0 )T . Expo-
nentiating the first two equations and solving the last one for t, we get
x 1+t y (1 + 12 t)2
= , = and t = t0 + ln(z/z0 ).
x0 1 + t0 y0 (1 + 21 t0 )2
We can use the last equation to eliminate t so the particle path/trajectory through
(x0 , y0 , z0 )T is the curve in three dimensional space given by
2
1 + 21 t0 + 1
1 + t0 + ln(z/z0 ) 2 ln(z/z0 )
x = x0 · , and y = y0 · 1
.
(1 + t0 ) (1 + 2
2 t0 )
To find the streamlines, we fix time t. We must then solve the system of equations
dx dy dz
= u, =v and = w,
ds ds ds
with t fixed, and then eliminate s between them. Hence for streamlines we have
dx x dy y dz
= , = and = z.
ds 1+t ds 1 + 12 t ds
Assuming that we are interested in the streamline that passes through the point
(x0 , y0 , z0 )T , we again use the method of separation of variables and integrate with
respect to s from s0 to s, for each of the three equations. This gives
x s − s0 y s − s0 z
ln = , ln = and ln = s − s0 .
x0 1+t y0 1 + 12 t z0
Using the last equation, we can substitute for s − s0 into the first equations. If we then
multiply the first equation by 1 + t and the second by 1 + 12 t, and use the usual log
law ln ab = b ln a, then exponentiation reveals that
1+t 1+ 12 t
x y z
= = ,
x0 y0 z0
which are the equations for the streamline through (x0 , y0 , z0 )T .
4 Simon J.A. Malham
dx dr dθ
= cos θ − r sin θ ,
ds ds ds
dy dr dθ
= sin θ + r cos θ .
ds ds ds
Multiply the first equation above by −r sin θ and the second equation by r cos θ. Adding
the resulting two equations reveals that
dθ dx dy
r2 = −y +x .
ds ds ds
Now multiply the first equation above by −r cos θ and the second equation by r sin θ.
Adding these resulting two equations reveals that
dr dx dy
r =x +y .
ds ds ds
The equations for the streamlines are
dx dy dz
= α(t) x − y , = α(t) x + y and = 0.
ds ds ds
Hence z(s) = z0 for all s. The equation for r = r(s) is given by
dr dx dy
r =x +y
ds ds ds
= x · α(t) x − y + y · α(t) x + y
= α(t) x2 + y 2
= α(t)r2 .
Hence we see r = r(s) satisfies r0 = α(t)r which implies r(s) = r0 exp α(t)s . Similarly
dθ dx dy
r2 = −y +x
ds ds ds
= − y · α(t) x − y + x · α(t) x + y
= α(t) x2 + y 2
= α(t)r2 .
Hence we find θ = θ(s) satisfies θ0 = α(t), i.e. θ(s) = θ0 + α(t)s. Hence for any particle,
the θ = θ(s) coordinate increases linearly in time, while the r = r(s) coordinate
exponentially expands or contracts depending on the sign of α(t) and the z = z(s) is
constant. The streamlines thus look like horizontal exponential spirals.
Introductory fluid mechanics: solutions 5
dy dy/ds dx dx/ds
= or =
dx dx/ds dy dy/ds
We observe this is a first order linear differential equation for x = x(y), i.e. we are
treating y as the independent variable and x as the dependent variable and solution. We
can solve first order linear differential equations using the integrating factor technique,
where in this case the integrating factor is
Z y
sin η
exp − dη = exp log(cos y) = cos y.
0 cos η
Hence the integrating factor is cos y and we multiply the first order linear differential
equation above through by it. Hence we have
dx
cos y = 1 + (sin y) x
dy
d
⇔ (cos y) x = 1
dy
⇔ (cos y) x = y − y0 + (cos y0 ) x0
y − y0 + (cos y0 ) x0
⇔ x= .
cos y
This gives the equation for the streamline passing through (x0 , y0 )T .
Note that when x = 1 and y = −π/2 as well as when x = −1 and y = π/2, one
can check by direct inspection that dx/ds = 0 and dy/ds = 0 and thus both points are
stagnation points of the flow.
6 Simon J.A. Malham
A1 u1 = A2 u2 .
(b) Bernoulli’s Theorem implies the quantity H is the same in wide and narrow
regions on same streamline passing through the tube, i.e. we have
p1 p
1 2
2 u1 + = 12 u22 + 2 .
ρ ρ
p1 − p2 = 12 ρ(u22 − u21 ).
The quantity on the right is positive since u2 > u1 . Hence the pressure in the narrow
region is less!
(d) Examples are: a device for measuring flow speeds (measure p1 − p2 , A1 /A2
known =⇒ u1 , u2 ); a carburetor—the pressure drop draws in another fluid from a
side channel; or the lift of an aircraft wing.
10 Simon J.A. Malham
(Note that inside the container at the level at which water is pushed through the
orifice, the pressure is P = P0 + ρgz. The difference P − P0 accelerates the water
through the orifice.)
(b) Using incompressibility, we equate the volume flux through the orifice to the
rate of volume drop at the free surface (as it descends). We know S is the cross-sectional
area of the orifice and U is the velocity of the fluid through the orifice. The volume flux
through the orifice is equal to the cross section area S times the fluid velocity and is
thus SU . Similarly the cross-sectional area of the free surface at height z is say A, while
its velocity is dz/dt. Hence the drop in volume per unit time at the free surface—i.e.
the volume flux at the free surface—is A dz/dt. Hence incompressibility implies
dz
SU = A .
dt
(d) We want dz/dt to be constant, so combining the results of parts (b) and (c),
and using that A = πr2 as the Clepsydra is a volume of revolution, implies
dz
2gz = πr2
p
S
dt
π dz 2
⇔ z= √ r4 ,
2g S dt
as required.
Introductory fluid mechanics: solutions 11
u2θ 1 ∂p
− =− ,
r ρ ∂r
1 ∂p
0=− ,
ρr ∂θ
1 ∂p
0=− − g.
ρ ∂z
Note we have
1 ∂(rur ) 1 ∂uθ ∂uz
∇·u= + + =0
r ∂r r ∂θ ∂z
trivially. And u · ∇ = (uθ /r)(∂/∂θ) thus also acts trivially on the velocity field which
is independent of θ. The second equation in the system implies p = p(r, z).
(b) Using that uθ = Ωr—this is the rigid body rotation ansatz—implies
∂p
= ρΩ 2 r ⇔ p(r, z) = 12 ρΩ 2 r2 + G(z),
∂r
where G an arbitrary function. Substitute this into the other equation
1 ∂p
= −g ⇔ G0 (z) = −ρg.
ρ ∂z
p(r, z) = 12 ρΩ 2 r2 − ρgz + C.
P0 = 12 ρΩ 2 r2 − ρgz + C
⇔ z = (Ω 2 /2g) r2 − (C − P0 )/ρg.
Ω 2 a2
z0 + = d.
4g
12 Simon J.A. Malham
Ω 2 a2 Ω 2 a2 Ω 2 a2 Ω 2 a2
z = z0 + =d− + =d+ .
2g 4g 2g 4g
Spillage occurs when this exceeds h, i.e. when
4g(h − d)
z>h ⇔ Ω2 > .
a2
If the mug is initially less than half full then d < 12 h. Spillage occurs when
4g(h − d) Ω 2 a2 Ω 2 a2
Ω2 > ⇔ > (h − d) ⇔ − 6 −(h − d).
a2 4g 4g
Thus we observe that
Ω 2 a2
z0 = d − 6 d − (h − d) = 2d − h < 0.
4g
In other words the centre part of the free surface (the dip in the middle) will have hit
the bottom of the mug before spillage.
Introductory fluid mechanics: solutions 13
where we will ignore terms of degree x3 and higher. The quantities K, a1 and a2 are all
constants. Naturally since x = 0 is a local maximum for y = y(x) we know K > 0. We
substitute these expansions into the Bernoulli constraint (again see the general theory
in the notes)
U2 (U H)2
y= +H −h− .
2g 2gh2
−2
Note to expand h we use the Binomial expansion with X := a1 x + a2 x2 + · · · so
−2
h−2 = h−2
0 1 + a1 x + a2 x2 + · · ·
−2
= h−2
0 1+X
(−2)(−3) 2
= h−2
0 1 + (−2X) + X + ···
2!
= h−2 1 − 2(a1 x + a2 x2 + · · · ) + 3(a1 x + a2 x2 + · · · )2 + · · ·
0
= h−2 1 − 2a1 x + (3a21 − 2a2 )x2 + · · · .
0
14 Simon J.A. Malham
sub a 1<0
y0 y0
x=0 x x=0 x
Fig. 1 The cases for the two incident (upstream) Froude numbers are shown. The flow is
subcritical if a1 x > 0 and supercritical if a1 x < 0. For a real flow friction will force the
supercritical solution upstream in each case. Thus for example, in the case F > 1 upstream,
there will be a discontinuity in the gradient of h = h(x) at x = 0.
U2 (U H)2
y0 −Kx2 +· · · = +H−h0 1+a1 x+a2 x2 +· · · − 1−2a1 x+(3a21 −2a2 )x2 +· · · .
2g 2
2gh0
U2 (U H)2
x0 : y0 = + H − h0 − ,
2g 2gh20
(U H)2
x1 : 0 = −h0 a1 − (−2a1 ),
2gh20
(U H)2
x2 : − K = −h0 a2 − (3a21 − 2a2 ).
2gh20
Note that the relation between the coefficients at x0 simply reproduces the definition
for the maximum value y0 . Then for the relation between the coefficients at x1 , since
U H = uh0 and u2 = gh0 we see that the expression on the right is
(U H)2 u2 h20 u2
−a1 h0 − = −a1 h0 − = −a1 h0 − = 0,
gh20 gh20 g
u2 h20
K = h0 a2 + (3a21 − 2a2 )
2gh20
h
⇔ K = h0 a2 + 0 (3a21 − 2a2 )
2
2K 1/2
⇔ a1 = ± .
3h0
Around x = 0 where y = y0 = ymax , using the Binomial expansion we see that
u2 (U H)2 (U H)2
= 3
= (1 − 3a1 + · · · ).
gh gh gh30
Hence locally the flow is subcritical so F < 1 if a1 x > 0, and supercritical so F > 1 if
a1 x < 0; see Fig. 1. Friction forces the supercritical solution downstream in each case.
Introductory fluid mechanics: solutions 15
(b) Hence H = f (t) for some function f = f (t). Using the suggested redefined
potential for u given by
Z t Z t
V =ϕ− f (τ ) dτ ⇔ ϕ=V + f (τ ) dτ
0 0
Hence if we substitute the rigid body rotation ansatz (u, v)T = (−Ωy, Ωx)T , into the
Euler equations we find
−Ω 2 x
−Ω̇y 1 ∂p/∂x αx + βy
+ =− + .
Ω̇x −Ω 2 y ρ ∂p/∂y γx + δy
Consider the curl of these equations—the curl of a vector u = (u, v)T which only has
two components is the scalar −∂u/∂y + ∂v/∂x. By direct computation this gives
1
Ω̇ = (γ − β).
2
Now let us separate the equations above, i.e. those resulting from substituting the
rigid body rotation ansatz into the Euler equations. We consider them as two partial
differential equations for the pressure p as follows
1 ∂p
= Ω̇y + Ω 2 x + αx + βy,
ρ ∂x
1 ∂p
= −Ω̇x + Ω 2 y + γx + δy.
ρ ∂y
Recall to show that the rigid body rotation ansatz is a solution to the incompressible
Euler equations we must find a pressure field p such that (u, v)T = (−Ωy, Ωx)T and p
satisfy them. We integrate the first equation above with respect to x. This generates
the relation
p 1 1
= Ω̇yx + Ω 2 x2 + αx2 + βxy + F (y),
ρ 2 2
where F = F (y) is an arbitrary function. If we now substitute this expression into the
second equation above and use that Ω̇ = 12 (γ − β) we find
F 0 (y) = Ω 2 y + δy.
where ω is the scalar vorticity and H := p/ρ + 12 (u2 + v 2 ). Using the identity
1
∇ |u|2 = u · ∇u + u × (∇ × u),
2
we can write the steady Euler equations as (ρ is constant and we assume no body force)
1 p
∇ |u|2 − u × (∇ × u) = −∇
.
2 ρ
Substituting these expressions back into the equation above we find (ignoring the last
component)
1
p vω
∇ u2 + v 2 + ∇
= ,
2 ρ −uω
thus demonstrating the equivalence.
To show the scalar vorticity ω is constant along streamlines we observe that
∂ ∂
∇ × ∇H ≡ 0 ⇔ (vω) − (−uω) ≡ 0
∂y ∂x
∂v ∂u ∂ω ∂ω
⇔ + ω+u +v ≡ 0.
∂y ∂x ∂x ∂y
Incompressibility implies the first term on the left is zero. Hence the relation above is
equivalent to u · ∇ω ≡ 0 and thus ω is constant along streamlines.
18 Simon J.A. Malham
(b) We integrate the equation above with respect to r twice to find uz = uz (r) as
follows. Rearranging the equation and then integrating twice we find
∂ ∂uz C
r =− r
∂r ∂r ρν
∂uz C 2
⇔ r =− r +A
∂r 2ρν
∂uz C A
⇔ =− r+
∂r 2ρν r
C 2
⇔ uz = − r + A log r + B
4ρν
where A and B are arbitrary constants. We want the solution to be bounded for
0 6 r 6 a and therefore we must insist A = 0 because log r → −∞ as r → 0. The
no-slip boundary condition at the pipe wall r = a implies
C 2 Ca2
0=− a +B ⇔ B= .
4ρν 4ρν
Hence we see that
C 2 Ca2 C
uz = − r + = (a2 − r2 ).
4ρν 4ρν 4ρν
(c) Let S represent a disc cross-section of the pipe. Let dS represent a small patch
of area on S. Since ρ is the mass per unit volume and uz dS represents the volume of
fluid passing through dS per unit time, the mass flow rate across dS is ρuz dS. The
total mass flow rate across S is thus
Z Z a Z 2π
ρuz dS = ρuz (r) rdθdr
S 0
Z0 a
C
= 2πρ (a2 − r2 ) rdr
0 4ρν
πC 1 2 2 1 4 r=a
h i
= a r − r
2ν 2 4 r=0
πCa4
= .
8ν
Introductory fluid mechanics: solutions 19
∂p
0=− ,
∂x
∂p
0=− ,
∂y
∂p ∂2w ∂2w
0=− +ν 2
+ 2 ,
∂z ∂ x ∂ y
∂2w ∂2w G
2
+ 2 =− .
∂ x ∂ y ν
x2 y2 x2
2
+ 2 =1 ⇔ y 2 = b2 1 − ,
a b a2
generates the equation
x2 b2 2
Ax2 + Bb2 1 − +C =0 ⇔ A−B x + Bb2 + C = 0
a2 a2
which must hold for all x ∈ [−a, a]. Hence equating coefficients of x0 and x2 , we arrive
at the following three equations (including the result from part (b) above) for the three
unknowns A, B and C:
b2 G
A−B = 0, Bb2 + C = 0 and A+B =− .
a2 2ν
Solving the first equation for A in terms of B and substituting this into the third
equation we find
b2 G G a2
B 1+ =− ⇔ B=− · 2 .
a2 2ν 2ν a + b2
G a2 b2 G b2
A+ · 2 2
· 2 =0 ⇔ A=− · 2 .
2ν a + b a 2ν a + b2
20 Simon J.A. Malham
Finally substituting these two answers for A and B into the second equation implies
G a2 G a2 b2
− · 2 · b2 + C = 0 ⇔ C= · 2 .
2ν a + b2 2ν a + b2
(d) For a small patch of area dS of an elliptical cross section of the pipe, the volume
of fluid passing through dS per unit time is equivalent to the volume of the cylinder of
cross sectional area dS and length w (the orthogonal flow rate through dS), i.e. w · dS.
Summing over all such small patches of areas to make up the complete cross section
generates the integral Z
w dS
πa3 b3 G
Q := .
4ν(a2 + b2 )
We wish to find the maximum of Q subject to the constraint that the cross sectional
area of the elliptic pipe is given, say by K, i.e. πab = K with K fixed. Simply substitute
that b = K/πa into Q and differentiate with respect to a implies
∂Q ∂ π(K/π)3 G
=
∂a ∂a 4ν(a + K 2 π −2 a−2 )
2
2a − 2K 2 π −2 a−3 = 0
p
⇔ a= K/π.
p
Note that when a = K/π, then
where p is the pressure field, u the velocity field and µ is the viscosity. Since the flow
is uniform in the direction aligned with the imaginary line of intersection between the
plates—denote this the z-axis—there exists a stream function ψ = ψ(r, θ) such that in
cylindrical polar coordinates
∆(∆ψ) = 0.
To deduce this we simply follow the argument in the viscous corner flow example in
the Stokes flow section. Note the velocity components ur and uθ and stream function
are related by
1 ∂ψ ∂ψ
ur = and uθ = − .
r ∂θ ∂r
For the second part of the problem let us determine the appropriate boundary condi-
tions. We will assume a lower plate that is stationary and Ω is the angular velocity of
the upper plate relative to the lower plate. Using no-slip boundary conditions on the
plate along θ = 0 which is stationary we see that
1 ∂ψ ∂ψ
=0 and =0 on θ = 0.
r ∂θ ∂r
Also using the no-slip boundary conditions on the moving plate which at time t lies
along θ = β(t) > 0, we have
1 ∂ψ ∂ψ
=0 and = −Ωr on θ = β.
r ∂θ ∂r
Hence our goal now is to solve the biharmonic equation ∆(∆ψ) = 0 subject to the four
boundary conditions above. We thus look for a solution of the form
ψ = Ωr2 f (θ),
for some arbitrary function f = f (θ). This solution ansatz is suggested by the boundary
condition on the moving upper plate. We solve the biharmonic equation here in two
stages. First we observe that
∂2 1 ∂ 1 ∂2
∆ Ωr2 f (θ) = Ω r2 f (θ)
+ +
∂r2 r ∂r r2 ∂θ2
= Ω 4f (θ) + f 00 (θ) .
Hence our solution guess ψ = Ωr2 f (θ) satisfies the biharmonic equation if and only if
F 00 (θ) = 0. Hence we find F = 4A + 4Bθ for some constants A and B—the prefactor
‘4’ is chosen for convenience. Hence we find, combining the complementary function
and particular integral for f 00 + 4f = F , we have
f = A + Bθ + C cos(2θ) + D sin(2θ),
where C and D are two further constants. For our solution ansatz ψ = Ωr2 f (θ) the
four boundary conditions become
1
f (0) = f 0 (0) = f 0 (β) = 0 and f (β) = − .
2
The four constants are determined by the four boundary conditions above, and indeed
we find
sin θ cos(β − θ) − θ cos β
f (θ) = .
2(β − tan β) cos β
We finally focus on the issue of the distance within O within which this solution is
self-consistent. Our solution ansatz was inferred from the boundary condition uθ = Ωr
on the upper plate. Hence the inertia terms u · ∇u scale like Ω 2 r2 /r while the viscous
terms ν∆u scale like ν(Ωr)/r2 . Our Stokes flow assumption was
Ω 2 r2 /r Ωr2
1 ⇔ 1.
ν(Ωr)/r2 ν
The steady Stokes flow equations involving the velocity field u and pressure p are
µ∆u = ∇p,
where µ is the first coefficient of viscosity. Since there is only one non-zero component
of velocity uϕ , the Stokes flow equations in spherical polar coordinates reduce to
1 ∂p
µ∆uϕ = .
r sin θ ∂ϕ
Since uϕ = uϕ (r, θ) only, we deduce that p = p(r, θ) can only be linear in ϕ. However
p must be periodic in ϕ hence p must be constant, i.e. invariant to ϕ. Hence we have
∂p
= 0.
∂ϕ
Hence the Stokes flow equations reduce to ∆uϕ = 0 or ∆u = 0. We employ the strategy
suggested in the notes for such Stokes flow equations and write
∆u = 0
0
⇔ ∆ 0 = 0
uϕ
0
⇔ ∇×∇× 0 =0
uϕ
0
⇔ 0 = 0,
1 2
− r sin θ D (r sin θ) u ϕ
where
∂2 sin θ ∂ 1 ∂
D2 := + 2 .
∂r2 r ∂θ sin θ ∂θ
24 Simon J.A. Malham
Substituting the solution ansatz uϕ = Ωaf (r) sin θ into the equation above we find
1
D2 (r sin θ) uϕ = 0
−
r sin θ
D2 r sin θ Ωaf (r) sin θ = 0
⇔
D2 rf (r) sin2 θ = 0
⇔
∂2 f (r) ∂ 1 ∂
sin2 θ (sin2 θ) = 0
⇔ 2
rf (r) + sin θ
∂r r ∂θ sin θ ∂θ
∂ f (r) ∂
sin2 θ f (r) + rf 0 (r) + sin θ
⇔ (2 cos θ) = 0
∂r r ∂θ
f (r)
sin2 θ 2f 0 (r) + rf 00 (r) − 2 sin2 θ
⇔ =0
r
f (r)
⇔ rf 00 (r) + 2f 0 (r) − 2 =0
r
⇔ r2 f 00 (r) + 2rf 0 (r) − 2f (r) = 0.
This is a linear second order Cauchy–Euler type of equation. If we set ρ := log r and
suppose F (ρ) = f (r) then we see using the chain rule that
dF dρ dF 1
f 0 (r) = = = F 0 (ρ)
dr dr dρ r
and
d 1 0 1 1 dρ d 1 1
f 00 (r) = F (ρ) = − 2 F 0 (ρ) + F 0 (ρ) = − 2 F 0 (ρ) + 2 F 00 (ρ).
dr r r r dr dρ r r
Substituting these expressions into the Cauchy–Euler equation above we find
F 00 + F 0 − 2F = 0,
for F = F (ρ). This is a second order constant coefficient differential equation for F
whose independent solutions are exp(ρ) = r and exp(−2ρ) = r12 . Hence the general
solution to the Cauchy–Euler equation for f = f (r) is
B
f = Ar + ,
r2
for arbitrary constants A and B. Using the boundary condition that uϕ → 0 as r → ∞
implies A = 0. Then on the boundary r = a of the sphere we must have
uϕ (a, θ) = Ωa sin θ
⇔ Ωaf (a) sin θ = Ωa sin θ
⇔ f (a) = 1
⇔ B = a2 .
Hence the solution to the Stokes flow equations and boundary conditions is
Ωa3 sin θ
uϕ = .
r2
Introductory fluid mechanics: solutions 25
Having computed the velocity field, we can now compute the coupling exerted
on the fluid due to the sphere. Recall that for an incompressible Newtonian fluid
the deviatoric stress σ̂ is related to the deformation matrix D through the formula
σ̂ = 2µD. The components of the deformation matrix D are given the appendix of the
notes. Since the only non-zero component of the velocity field is uϕ = Ωa3 sin θ/r2 the
only non-zero component of the deformation matrix D is
−3Ωa3 sin θ
r ∂ uϕ
Drϕ = = .
2 ∂r r 2r3
The total couple exerted on the fluid at the sphere boundary is
Z π
a sin θR̂ × (σ̂n) 2πa2 sin θ dθ.
0
By analogy with the way we computed the force on a sphere in a uniform Stokes flow,
we split the sphere here into concentric rings about the axis of rotation and compute
the coupling at the sphere surface r = a. Hence in the above integral 2πa2 sin θ dθ
represents the area of the concentric ring at angle θ to the axis of rotation. Note we
have n = −r̂ and so σ̂n = −σ̂r̂ represents the stress at the angle θ on the surface
r = a of the sphere. The vector R̂ is a vector orthogonal to the axis of rotation and
a sin θ represents the distance from that axis of rotation to a point on the sphere. Thus
the coupling on the sphere at angle θ is −a sin θR̂ × (σ̂r̂). Now we compute the stress:
σ̂r̂ = 2µDr̂
Drr Drθ Drϕ 1
= 2µ Drθ Dθθ Dθϕ 0
Drϕ Dθϕ Dϕϕ 0
0
= 2µ 0
Drϕ
= 2µDrϕ ϕ̂.
Hence the coupling is −2µa sin θDrϕ R̂ × ϕ̂. We observe that R̂ × ϕ̂ = Ω̂ as it should
be and −2µa sin θDrϕ is the component we are after. We can now compute the integral
to obtain the total coupling
π π
−3Ωa3 sin θ
Z Z
(−2µa sin θDrϕ )(2πa2 sin θ) dθ = −4µπa3 sin2 θ dθ
0 2a3
Z0 π
3
= 6µπΩa sin3 θ dθ
0
Z π Z π
3
= 6µπΩa sin θ dθ − cos2 θ sin θ dθ
0 0
Z π
3 2
= 6µπΩa 2+ cos θ d(cos θ)
0
2
= 6µπΩa3 2 −
3
= 8µπΩa3 .
26 Simon J.A. Malham
∂p ∂2u
=µ 2,
∂x ∂z
where z represents the vertical coordinate, u = u(x, z) is the velocity in the x-direction
and µ is the viscosity. Integrating this equation with respect to z we find
1 ∂p
u=− z(h − z).
2µ ∂x
The volume flux through the layer is constant say Q, i.e. we have
Z h
u dz = Q.
0
1 ∂p h3 ∂p 12µQ
− =Q ⇔ =− 3 ,
2µ ∂x 6 ∂x h
as we required—so the constant A ≡ −12µQ.
Now suppose h = Ce−Bx ; this implies
∂p 12µQ 3Bx 4µQ 3Bx
=− e ⇔ p(x) = p(0) − e −1 .
∂x C3 BC 3
∂2u ∂h
µ = ρg
∂z 2 ∂x
∂u ∂h ∂Γ
⇔ µ = ρg (z − h) +
∂z ∂x ∂x
1 ∂h ∂Γ
⇔ µu = 2 ρg z(z − 2h) + z ,
∂x ∂x
and similarly
∂h ∂Γ
µv = 12 ρg z(z − 2h) + z ,
∂y ∂y
where in the first integration from z to h, we used the given conditions for the applied
surfaces stresses at z = h(x, y, t) and in the second integration from 0 to z, we used
the no-slip boundary conditions at z = 0.
Let us now restrict ourselves to the surface z = h(x, y, t). The acceleration of any
particle at the surface is given by
∂h ∂h ∂h
+ u(x, y, h, t) + v(x, y, h, t) .
∂t ∂x ∂y
28 Simon J.A. Malham
since w = 0 on z = 0 and z = h.
Integrating the shallow layer equations with respect to z twice we find
1 ∂p 1 ∂p
u=− z(h − z) and v=− z(h − z),
2µ ∂x 2µ ∂y
using the no-slip boundary conditions at z = 0 and z = h. Hence the scalar quantity
representing the curl of u = (u, v)T is given by
Z h Z h
∂ 1 ∂ 1
∇×u= u dz − v dz
∂y h 0 ∂x h 0
Z h
1 ∂u ∂v
= − dz
h 0 ∂y ∂x
h
z(h − z) ∂ 2 p ∂2p
Z
1
=− − dz
h 0 2µ ∂x∂y ∂y∂x
= 0.
Hence the vertically averaged velocity field is both incompressible and irrotational.
(d) Within a distance O(h) of the cylinder the solution above is no longer rep-
resentative as no-slip boundary conditions must hold on the cylinder boundary. The
solution further away from the cylinder is least disturbed if we suppose the boundary
condition for u on the cylinder to be u · n = 0.
30 Simon J.A. Malham
Second we compute the form of the vorticity for the flow, we have
1 ∂uz ∂uθ
r ∂θ − ∂z
∇×u= ∂ur ∂uz
∂z − ∂r
1 ∂ 1 ∂ur
r ∂r (ruθ ) − r ∂θ
0
= 0
1 ∂
r ∂r (ru θ ) − 0
0
= 0 ,
ω
1 ∂
where we set ω = ω(r, t) to be ω = r ∂r (ruθ ). The equation for the evolution of
vorticity ω = ω(x, t) is given by
∂ω
+ u · ∇ω = ν∆ω + ω · ∇u.
∂t
For the vorticity form above, in cylindrical polar coordinates, using for any vector v,
we have
∂ v ∂ ∂
v · ∇ = vr + θ + vz ,
∂r r ∂θ ∂z
and using the form for the Laplacian in cylindrical polar coordinates from the formulae
in the Appendix, the equations above for the evolution of vorticity collapse to
∂ω ∂ω 1 ∂ ∂ω ∂uz
+ ur =ν r +ω .
∂t ∂r r ∂r ∂r ∂z
(a) For the moment we assume the flow is inviscid and at time t = 0 the initial
data has the form ω = ω0 f (r) for some function f = f (r) and some constant ω0 . We
are required to verify that
ω = ω0 eαt f reαt/2
is a solution for all t > 0. Note that this ansatz satisfies the initial condition at t = 0.
Looking at the evolution equation for ω = ω(r, t) above with ν = 0, we see that we
need to compute
∂ω rα αt/2
= αω0 eαt f reαt/2 + ω0 eαt f 0 reαt/2 ·
e ,
∂t 2
∂ω
= ω0 eαt f 0 reαt/2 · eαt/2 .
∂r
Introductory fluid mechanics: solutions 31
Subsituting the ansatz for ω above and these two expressions for its partial derivatives
with respect to t and r into the acceleration terms in evolution equation for ω = ω(r, t),
we find
∂ω ∂ω rα αt/2
= αω0 eαt f reαt/2 + ω0 eαt f 0 reαt/2 ·
+ ur e
∂t ∂r 2
αr
ω0 eαt f 0 reαt/2 · eαt/2
−
2
= αω0 eαt f reαt/2
∂uz
=ω .
∂z
Hence we deduce that the ansatz does indeed give a solution to the evolutionary vor-
ticity equation in the inviscid case when ν = 0.
We can interpret this solution ansatz as follows. Consider a fluid particle whose
radial position is prescribed by r = r(t). Suppose that at t = 0 its radial position was
r(0) = r0 . Its radial trajectory is thus given by
dr α
=− r ⇔ r = r0 e−αt/2 .
dt 2
Thus a particle that at time t has radial coordinate r started with radial coordinate
reαt/2 at time t = 0. Thus in the ansatz ω = ω0 eαt f reαt/2 we see that as time evolves
the vorticity intensifies exponentially fast due to: (i) Stretching via the exponential
prefactor eαt , and (ii) Inwards advection via the term reαt/2 in the argument of f .
(b) Now consider steady viscous flow. The equation for the vorticity above becomes
αr ∂ω 1 ∂ ∂ω
− =ν r + αω.
2 ∂r r ∂r ∂r
2
We look for a solution ω = ω(r) with the ansatz ω = ω0 e−αr /4ν , where ω0 is a
constant. We observe that
∂ω αr 2
= −ω0 e−αr /4ν ,
∂r 2ν
and
∂
∂ω ω α ∂ 2 −αr2 /4ν
r =− 0 r e
∂r ∂r 2ν ∂r
ω α αr3 −αr2 /4ν
= − 0 2r − e .
2ν 2ν
Substituting these expressions into the vorticity equation above we find
1 ∂ ∂ω ω0 α αr2 −αr2 /4ν 2
ν r + αω = − 2− e + αω0 e−αr /4ν
r ∂r ∂r 2 2ν
ω0 α2 r2 −αr2 /4ν
= e
4ν
αr ω αr 2
=− − 0 e−αr /4ν
2 2ν
αr ∂ω
=− .
2 ∂r
2
Hence ω = ω0 e−αr /4ν is a solution to the steady viscous vorticity equation. A steady
solution is possible as there is a dominant balance between outwards diffusion, due to
the viscous term, and inwards advection and stretching.
32 Simon J.A. Malham
δ(∂/∂x)(U δ) (∂U/∂x)δ 2
α := and β := .
ν ν
Since f is independent of x, this last relation implies α and β are independent of x
also and are thus constants.
The product rule implies
∂ 2 ∂δ ∂
(δ U ) = δU + δ (δU )
∂x ∂x ∂x
∂ ∂U
=δ (U δ) − δ + να
∂x ∂x
= να − β + να
= ν(2α − β),
Acknowledgements The lecture notes were to a large extent grown out of a merging of, lec-
tures on Ideal Fluid Mechanics given by Dr. Frank Berkshire [1] in the Spring of 1989, lectures
on Viscous Fluid Mechanics given by Prof. Trevor Stuart [5] in the Autumn of 1989 (both at
Imperial College) and the style and content of the excellent text by Chorin and Marsden [2].
They also benefited from lecture notes by Prof. Frank Leppington [3] on Electromagnetism.
Some of these solutions are derived from their solutions to the problems in their lecture notes.
References
1. Berkshire, F. 1989 Lecture notes on ideal fluid dynamics, Imperial College Mathematics
Department.
2. Chorin, A.J. and Marsden, J.E. 1990 A mathematical introduction to fluid mechanics,
Third edition, Springer–Verlag, New York.
3. Leppington, F. 1989 Electromagnetism, Imperial College Mathematics Department.
4. Majda, A.J. and Bertozzi, A.L. 2002 Vorticity and incompressible flow, Cambridge Texts
in Applied Mathematics, Cambridge University Press.
5. Stuart, J.T. 1989 Lecture notes on Viscous Fluid Mechanics, Imperial College Mathematics
Department.