Mathematical Method For Physicists Ch. 1 & 2 Selected Solutions Webber and Arfken

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Physics 451 Fall 2004

Homework Assignment #3 — Solutions

Textbook problems: Ch. 1: 1.7.1, 1.8.11, 1.8.16, 1.9.12, 1.10.4, 1.12.9


Ch. 2: 2.4.8, 2.4.11

Chapter 1

1.7.1 For a particle moving in a circular orbit ~r = x̂ r cos ωt + ŷ r sin ωt


(a) evaluate ~r × ~r˙

Taking a time derivative of ~r, we obtain

~r˙ = −x̂ rω sin ωt + ŷ rω cos ωt (1)

Hence

~r × ~r˙ = (x̂ r cos ωt + ŷ r sin ωt) × (−x̂ rω sin ωt + ŷ rω cos ωt)


= (x̂ × ŷ)r2 ω cos2 ωt − (ŷ × x̂)r2 ω sin2 ωt
= ẑ r2 ω(sin2 ωt + cos2 ωt) = ẑ r2 ω

(b) Show that ~¨r + ω 2~r = 0

The acceleration is the time derivative of (1)

~¨r = −x̂ rω 2 cos ωt − ŷ rω 2 sin ωt = −ω 2 (x̂ r cos ωt + ŷ r sin ωt) = −ω 2~r

Hence ~¨r + ω 2~r = 0. This is of course the standard kinematics of uniform circular
motion.

1.8.11 Verify the vector identity

~ × (A
∇ ~ × B)
~ = (B
~ · ∇)
~ A~ − (A
~ · ∇)
~ B~ − B(
~ ∇~ · A)
~ + A(
~ ∇~ · B)
~

This looks like a good time for the BAC–CAB rule. However, we have to be
~ has both derivative and vector properties. As a derivative, it
careful since ∇
operates on both A~ and B.
~ Therefore, by the product rule of differentiation, we
can write
↓ ↓
~ ~ ~ ~ ~ ~ ~ ~
∇ × (A × B) = ∇ × (A × B) + ∇ × (A × B) ~
where the arrows indicate where the derivative is acting. Now that we have
~ as a vector. Using the
specified exactly where the derivative goes, we can treat ∇
BAC–CAB rule (once for each term) gives

↓ ↓ ↓ ↓
~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ · A)
∇ × (A × B) = A(∇ · B) − B(∇ · A) + A(∇ · B) − B(∇ ~ (2)

The first and last terms on the right hand side are ‘backwards’. However, we can
turn them around. For example
↓ ↓ ↓
~ ~ ~ ~ ~ ~ ~ ~
A(∇ · B) = A(B · ∇) = (B · ∇)A

With all the arrows in the right place [after flipping the first and last terms in
(2)], we find simply

~ × (A
∇ ~ × B)
~ = (B
~ · ∇)
~ A~ − B(
~ ∇~ · A)
~ + A(
~ ∇~ · B)
~ − (A
~ · ∇)
~ B~

which is what we set out to prove.

1.8.16 An electric dipole of moment p~ is located at the origin. The dipole creates an electric
potential at ~r given by
p~ · ~r
ψ(~r ) =
4πo r3
~ = −∇ψ
Find the electric field, E ~ at ~r.

We first use the quotient rule to write



p~ · ~r
 ~ p · ~r ) − (~
1 r3 ∇(~ ~ 3)
p · ~r )∇(r
E ~ =− 1 ∇
~ = −∇ψ ~ =−
4π0 r3 4π0 r6

Applying the chain rule to the second term in the numerator, we obtain
3~ ~
~ = − 1 r ∇(~p · ~r ) − 3r2 (~
p · ~r )∇(r)
E
4π0 r6

We now evaluate the two separate gradients

~ p · ~r ) = x̂i ∂ (pj xj ) = ~xi pj ∂xj = x̂i pj δij = x̂i pi = p~


∇(~
∂xi ∂xi

and

~ = x̂i ∂ 1 x̂i xi ~r
q
∇r x21 + x22 + x23 = x̂i p 2 2xi = = = r̂
∂xi 2 x1 + x22 + x23 r r
Hence
3 2
~ = − 1 r p~ − 3r (~
E
p · ~r )r̂
=−
1 p~ − 3(~p · r̂)r̂
4π0 r 6 4π0 r 3

Note that we have used the fact that p~ is a constant, although this was never
stated in the problem.

1.9.12 Show that any solution of the equation

~ × (∇
∇ ~ × A)
~ − k2 A
~=0

automatically satisfies the vector Helmholtz equation

~ + k2 A
∇2 A ~=0

and the solenoidal condition


~ ·A
∇ ~=0

We actually follow the hint and demonstrate the solenoidal condition first. Taking
the divergence of the first equation, we find

~ ·∇
∇ ~ × (∇
~ × A)
~ − k2 ∇
~ ·A
~=0

However, the divergence of a curl vanishes identically. Hence the first term is
~ ·A
automatically equal to zero, and we are left with k 2 ∇ ~ = 0 or (upon dividing
~ ·A
by the constant k) ∇ ~ = 0.

We now return to the first equation and simplify the double curl using the BAC–
~
CAB rule (taking into account the fact that all derivatives must act on A)

~ × (∇
∇ ~ × A) = ∇(
~ ∇~ · A)
~ − ∇2 A
~ (3)

As a result, the first equation becomes

~ ∇
∇( ~ · A)
~ − ∇2 A
~ − k2 A
~=0

~ ·A
However, we have shown above that ∇ ~ = 0 for this problem. Thus (3) reduces
to
~ + k2 A
∇2 A ~=0

which is what we wanted to show.


H
1.10.4 Evaluate ~r · d~r

We have evaluated this integral in class. For a line integral from point 1 to point
2, we have Z 2 Z 2
2
1
~r · d~r = 2 d(r2 ) = 12 r2 1 = 12 r22 − 21 r12
1 1

However for a closed path, point


H 1 and point 2 are the same. Thus the integral
along a closed loop vanishes, ~r · d~r = 0. Note that this vanishing of the line
integral around a closed loop is the sign of a conservative force.
Alternatively, we can apply Stokes’ theorem
I Z
~r · d~r = ~ × ~r · d~σ

S

It is easy to see that ~r is curl-free. Hence the surface integral on the right hand
side vanishes.

1.12.9 Prove that I I


~ · d~λ = −
u∇v ~ · d~λ
v ∇u

This is an application of Stokes’ theorem. Let us write


I Z
~ ~ ~
(u∇v + v ∇u) · dλ = ∇~ × (u∇v
~ + v ∇u)
~ · d~σ (4)
S

We now expand the curl using

~ × (u∇v)
∇ ~ = (∇u)
~ × (∇v)
~ + u∇
~ × ∇v
~ = (∇u)
~ × (∇v)
~

where we have also used the fact that the curl of a gradient vanishes. Returning
to (4), this indicates that
I Z
(u∇v ~ · d~λ =
~ + v ∇u) ~ × (∇v)
[(∇u) ~ + (∇v)
~ × (∇u)]
~ · d~σ = 0
S

where the vanishing of the right hand side is guaranteed by the antisymmetry of
~×B
the cross-product, A ~ = −B~ × A.
~
Chapter 2

2.4.8 Find the circular cylindrical components of the velocity and acceleration of a moving
particle

We first explore the time derivatives of the cylindrical coordinate basis vectors.
Since

ρ̂ = (cos ϕ, sin ϕ, 0), ϕ̂ = (− sin ϕ, cos ϕ, 0), ẑ = (0, 0, 1)

their derivatives are


∂ ρ̂ ∂ ϕ̂
= (− sin ϕ, cos ϕ, 0) = ϕ̂, = (− cos ϕ, − sin ϕ, 0) = −ρ̂
∂ϕ ∂ϕ

Using the chain rule, this indicates that

∂ ρ̂ ∂ ϕ̂
ρ̂˙ = ϕ̇ = ϕ̂ϕ̇, ϕ̂˙ = ϕ̇ = −ρ̂ϕ̇ (5)
∂ϕ ∂ϕ

Now, we note that the position vector is given by

~r = ρ̂ρ + ẑz

So all we have to do to find the velocity is to take a time derivative

~v = ~r˙ = ρ̂ρ̇ + ẑ ż + ρ̂ρ


˙ + ẑz
˙ = ρ̂ρ̇ + ẑ ż + ϕ̂ρϕ̇

Note that we have used the expression for ρ̂˙ in (5). Taking one more time deriva-
tive yields the acceleration

~a = ~v˙ = ρ̂ρ̈ + ẑ z̈ + ϕ̂(ρϕ̈ + ρ̇ϕ̇) + ρ̂˙ ρ̇ + ẑ˙ ż + ϕ̂ρ


˙ ϕ̇
= ρ̂ρ̈ + ẑ z̈ + ϕ̂(ρϕ̈ + ρ̇ϕ̇) + ϕ̂ρ̇ϕ̇ − ρ̂ρϕ̇2
= ρ̂(ρ̈ − ρϕ̇2 ) + ẑ z̈ + ϕ̂(ρϕ̈ + 2ρ̇ϕ̇)

2.4.11 For the flow of an incompressible viscous fluid the Navier-Stokes equations lead to

−∇ ~ × ~v )) = η ∇2 (∇
~ × (~v × (∇ ~ × ~v )
ρ0

Here η is the viscosity and ρ0 the density of the fluid. For axial flow in a cylindrical
pipe we take the velocity ~v to be
~v = ẑv(ρ)
From Example 2.4.1
~ × (~v × (∇
∇ ~ × ~v )) = 0

for this choice of ~v . Show that


~ × ~v ) = 0
∇2 (∇

leads to the differential equation

d2 v
 
1 d 1 dv
ρ 2 − =0
ρ dρ dρ ρ2 dρ

and that this is satisfied by


v = v0 + a2 ρ2

This problem is an exercise in applying the vector differential operators in cylin-


~ =∇
drical coordinates. Let us first compute V ~ × ~v

ρ̂ ρϕ̂ ẑ

~ × ~v = ∂
1 ∂ ∂ dv dv

~ =∇
V = −ϕ̂ ⇒ Vϕ = −
ρ ∂ρ ∂ϕ ∂z dρ dρ
0 0 v(ρ)

Note that, since v(ρ) is a function of a single variable, partial derivatives of v are
the same as ordinary derivatives. Next we need to compute the vector Laplacian
~ × ~v ) = ∇2 V
∇2 (∇ ~ . Using (2.35) in the textbook, and the fact that on the Vϕ
component is non-vanishing, we find

~ )ρ = − 2 ∂Vϕ
(∇2 V =0
ρ2 ∂ϕ
 
2~ 2 1 2 dv 1 dv
(∇ V )ϕ = ∇ (Vϕ ) − 2 Vϕ = −∇ + 2
ρ dρ ρ dρ
~ )z = 0
(∇2 V

This indicates that only the ϕ component of the vector Laplacian gives a non-
trivial equation. Finally, we evaluate the scalar Laplacian ∇2 (dv/dρ) to obtain

d2 v
 
2~ 1 d 1 dv
(∇ V )ϕ = − ρ 2 + (6)
ρ dρ dρ ρ2 dρ

Setting this equal to zero gives the equation that we were asked to prove.
To prove that v = v0 + a2 ρ2 satisfies the (third order!) differential equation, all
we have to do is substitute it in. However, it is more fun to go ahead and solve
the equation. First we notice that v only enters through its derivative f = dv/dρ.
Substituting this into (6), we find
 
1 d df 1
ρ − f =0
ρ dρ dρ ρ2

Expanding the derivatives in the first term yields

d2 f 1 df 1
2
+ − 2f = 0
dρ ρ dρ ρ

Since this is a homogeneous equation, we may substitute in f = ρα to obtain the


algebraic equation

α(α − 1) + α − 1 = 0 ⇒ α = ±1

This indicates that the general solution for f (ρ) is of the form

f = 2aρ + bρ−1

where the factor of 2 is chosen for later convenience. Integrating f once to obtain
v, we find Z
v= f dρ = v0 + aρ2 + b log ρ

which agrees with the given solution, except for the log term. However, now we
can appeal to physical boundary conditions for fluid flow in the cylindrical pipe.
The point ρ = 0 corresponds to the central axis of the pipe. At this point, the
fluid velocity should not be infinite. Hence we must throw away the log, or in
other words we must set b = 0, so that v = v0 + aρ2 .
Incidentally, the fluid flow boundary conditions should be that the velocity van-
ishes at the wall of the pipe. If we let R be the radius of the pipe, this means
that we can write the solution as

ρ2
 
v(ρ) = vmax 1 − 2
R

where the maximum velocity vmax is for the fluid along the central axis (with the
velocity going to zero quadratically as a function of the radius).

You might also like