Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

1 Two and Three dimensional problems

We now turn to the study of the Schrodinger equation in two and three dimensions. Let us start
by considering a central potential in two dimensions. A central potential will depend only on |�x|,
which in polar coordinates is r; thus V (�x) = V (r).
In general, the TISE will be written as
� �
h̄2 � 2
− ∇ + V (�x) ψ(�x) = Eψ(�x)
2m x

� x is the gradient with respect to �x.


Here ∇
In this course, we shall only consider central potentials, i.e. where the potential is only a function
of |�x|. A standard approach is the separation of variables.
In some special cases, the above equation can be separated by x, y, z. An example is the 3d
harmonic oscillator, for which we have
1 1
V (|�x|) = mω 2 |�x|2 = mω 2 (x2 + y 2 + z 2 )
2 2

The gradient operator is


2 ∂2 ∂2 ∂2
∇ =
� + +
∂x2 ∂y 2 ∂z 2
from which it follows that the Hamiltonian splits into

H = Hx + Hy + Hz

where
h̄2 ∂ 2 1
Hx = − + mω 2 x2
2m ∂x 2 2
is the Hamiltonian for a 1d SHO in the x direction, and similarly for Hy and Hz .
Now we simply assume that the stationary states also factorise as

ψ(�x) = X(x)Y (y)Z(z)

and each of the X, Y, Z are the stationary states of the 1d problem for that coordinate. Thus the
complete basis of stationary states for the 3d harmonic oscillator is

ψn1 n2 n3 = ψn1 (x)ψn2 (y)ψn3 (z)


where ψn1 (x) is the n1 th excited state of the 1d harmonic oscillator.
The energy of the above state is given by
� �
3
E n1 n2 n3 = n1 + n 2 + n 3 + h̄ω
2

1
Thus the ground state energy of the 3d harmonic oscillator is E000 = 32 h̄ω.
For the excited states, we encounter the phenomenon of degeneracy. The next higher energy value,
the first excited value is 52 h̄ω, but this can be achieved in three different ways, by taking (n1 , n2 , n3 )
to be (1, 0, 0) or (0, 1, 0) or (0, 0, 1). The three corresponding wavefunctions ψ100 , ψ010 , ψ001
are completely different functions having the same energy eigenvalue. We say that there is a
“threefold” degeneracy.

1.1 Plane polar coordinates


Let us begin by studying plane polar coordinates in 2d as an exercise. These results can be used
for cases where the particle is confined to a plane, such as in surface physics. To switch to plane
polar coordinates, we use the following relations

x = r cos φ , y = r sin φ
where
r≥0 , 0 ≤ φ < 2π
The inverse transformation is given by
� y
r= x2 + y 2 , φ = tan−1
x
(note that we must be careful in defining the branch for tan− 1 such that (x, y) and (−x, −y) are
mapped to different values of φ). Using the chain rule, we can calculate
2 2
�2= ∂ + ∂

∂x2 ∂y 2
as follows.
∂ ∂r ∂ ∂φ ∂
= +
∂x ∂x ∂r ∂x ∂φ
∂ sin φ ∂
= cos φ −
∂r r ∂φ
Similarly we find
∂ ∂r ∂ ∂φ ∂
= +
∂y ∂y ∂r ∂y ∂φ
∂ cos φ ∂
= sin φ +
∂r r ∂φ
Summing and squaring we get

�2= ∂ +1 ∂ + 1 ∂
2 2

∂r2 r ∂r r2 ∂φ2

This is almost separated among the radial and angular variables, and it is now useful to consider
the TISE for a central force.
� �
h̄2 � 2
− ∇ + V (r) ψ(r, φ) = Eψ(r, φ)
2m

2
Let us propose that the wavefunction can be factorised into a radial part R and an angular part Φ

ψ(r, φ) = R(r)Φ(φ)

Inserting the gradient operator and the factorised wavefunction into the TISE, we get
� �
h̄2 ∂2 1 ∂ 1 ∂2
− + + + V (r) R(r)Φ(φ) = ERΦ
2m ∂r2 r ∂r r2 ∂φ2
This is simplified as
� �
h̄2 ∂2 1 ∂ h̄2 R ∂ 2
− Φ + R − Φ + V (r)R(r)Φ(φ) = ERΦ
2m ∂r2 r ∂r 2m r2 ∂φ2

Dividing both sides by R(r)Φ(φ) and using the simplified notation R� (r) = ∂R
∂r
and Φ� (φ) = ∂Φ
∂φ

� �
h̄2 1 1 h̄2 1 ��
− R�� + R� − Φ + V (r) = E
2m R r 2mr2 Φ

Here now the first and third term depends on R and the second on Φ, while the right hand side is
just a number.
Now let us see if we can solve the angular part

Φ�� (φ) = λΦ(φ)

This will allow us to insert the solution into the above equation and obtain a purely radial equation.
Now solving this angular part is very straightforward, we have

Φ(φ) = einφ

for some unknown number n, and thus λ = −n2 . Now since φ is a periodic variable, the solution
to this equation must come back to itself as φ → φ + 2π. Thus

einφ → ein(φ+2π) = e2πin einφ

For the right hand side to be same as the left hand side, we see that n must be a real integer;
n = . . . , −2, −1, 0, 1, 2, . . .
Inserting this solution into the TISE above, we find an ordinary differential equation
� � � �
h̄2 1 � h̄2 n2
− R + R + V (r) +
��
R = ER
2m r 2mr2
Thus we have now reduced the problem to a one dimensional problem. There are however three
differences: (i) the differential operator is changed, having a first derivative together with the
second derivative, (ii) the effective potential has picked up an extra term

h̄2 n2
Ṽ (r) = V (r) +
2mr2
and (iii) the range of r is of course r ≥ 0.

3
Anticipating our need to normalise the wave function, we can choose to normalise the radial and
angular parts separately. The normalisation of Φ is given by
� �
dφ|einφ |2 = dφ = 2π

and the correctly normalised angular wave function is


1
Φ(φ) = √ einφ

The extra term that appears in the potential can be associated orbital angular momentum. This we
can anticipate of course from classical physics. In quantum mechanics, the angular momentum
will become an operator. Since L � = �x × p�, and both �x and p� lie in a plane, we can have angular
momentum only along the z-axis
L̂z = x̂p̂y − ŷ p̂x
. In the position basis,
� � � �
∂ ∂
L̂z → x −ih̄ − y −ih̄ = −ih̄(x∂y − y∂x )
∂y ∂x

Changing to polar coordinates we find that



L̂z → −ih̄
∂φ
which was the operator for the which the angular wavefunction was an eigenfunction. The angular
wave functions are eigenfunctions of the angular momentum

L̂z einφ = nh̄einφ

and we see that orbital angular momentum is quantised in integer units of h̄. This is also true in 3
dimensions, where the details are more complicated.
The effective potential in the Schrodinger equation can be written as

L2z
Ṽ (r) = V (r) +
2mr2

where by L2z we really mean the eigenvalue of the operator L̂2z .


So far, all we know about the potential is that it is central. For a central force, angular momentum
is conserved. In such cases, it is useful to reformulate the problem in terms of angular momentum
(which is completely solved) leaving only the radial problem. The radial problem, of course,
depends on the details of the potential and can only be solved only in special cases.

4
1.2 Spherical polar coordinates in 3d
Let us begin by using the relations to switch from cartesian to spherical polar coordinates

x = r sin θ cos φ
y = r sin θ sin φ
z = r cos θ

where we have r ≥ 0, 0 ≤ θ < π, and 0 ≤ φ < 2π. The inverse transformations are

r= x2 + y 2 + z 2
z
θ = cos−1 2
x + y2 + z2
y
φ = tan−1
x
where we must be careful in defining the branch for tan−1 such that (x, y) and (−x, −y) are
mapped to different values of φ.
Using the chain rule, we can now calculate

2 ∂2 ∂2 ∂2
∇ =
� + +
∂x2 ∂y 2 ∂z 2
as follows. For example, we have
∂ ∂r ∂ ∂θ ∂ ∂φ ∂
= + +
∂x ∂x ∂r ∂x ∂θ ∂x ∂φ
Proceeding for ∂
∂y
and ∂
∂z
, and summing and squaring we finally get
� �
� = 1 ∂ r+ 1 1 ∂ 1 ∂2
2
2 ∂2
∇ + + (1)
r ∂r2 r2 ∂θ2 tan θ ∂θ sin2 θ ∂φ2
This is not yet completely separated among the three variables, but we can notice that the first term
depends on r, while the factor in brackets depends only on θ and φ.
The factor depending on the angular part θ and φ has a useful physical interpretation, connected to
� = �x × p�.
the angular momentum. To see this, let us start from: L
We can write down the operator form by replacing the �x and p� with �xˆ and p�ˆ on the right hand side.
Using the representation for p�ˆ in the x-basis: p�ˆ = −ih̄∇
� we get
� �
∂ ∂
L̂x = −ih̄ y −z
∂z ∂y
� �
∂ ∂
L̂y = −ih̄ z −x
∂x ∂z
� �
∂ ∂
L̂z = −ih̄ x −y
∂y ∂x

5
These different components obey commutation relations which are essential in all studies of angu-
lar momentum in quantum mechanics. We can show that

[L̂x , L̂y ] = ih̄L̂z


[L̂y , L̂z ] = ih̄L̂x
[L̂z , L̂x ] = ih̄L̂y

Let us now transform to the spherical polar coordinates


� �
∂ cos φ ∂
L̂x = ih̄ sin φ +
∂θ tan θ ∂φ
� �
∂ sin φ ∂
L̂y = ih̄ − cos φ +
∂θ tan θ ∂φ

L̂z = −ih̄
∂φ

which as expected depend only on the angular derivatives ∂


∂θ
and ∂
∂φ
.

�ˆ 2 as follows
Now we can construct the operator L

�ˆ 2 = L̂2x + L̂2y + L̂2z


L
� �
2 ∂2 1 ∂ 1 ∂2
= −h̄ + +
∂θ2 tan θ ∂θ sin2 θ ∂φ2

As we see, this expression is precisely the one that appears in the brackets in Eq. 1. Putting this
into the Hamiltonian for our class of central problems, we have

�ˆ 2
� �
h̄2 1 ∂2 L
H=− r + + V (r)
2M r ∂r2 2M r2
where we have now denoted the mass by M . (**Note that from here on out, mass will be M ).
�ˆ 2 is a Hermitian operator and commutes with each the individual components:
The operator L

[L �ˆ 2 , L̂y ] = [L
�ˆ 2 , L̂x ] = [L �ˆ 2 , L̂z ] = 0

Now we have a strategy.


�ˆ 2 has its own eigenfunctions and eigenvalues:
L

�ˆ 2 χ(θ, φ) = λχ(θ, φ)
L

We will consider solutions of the full Schrodinger equation of the form

ψ(r, θ, φ) = R(r)χ(θ, φ)

6
Putting this back into the eigenvalue equation

Ĥψ(r, θ, φ) = E ψ(r, θ, φ)
 
�ˆ 2
2
� �
h̄ 1 ∂2 L
− r + + V (r) R(r)χ(θ, φ) = E R(r)χ(θ, φ)
2M r ∂r2 2M r2
� � � �
h̄2 1 ∂2 λ
− r + + V (r) R(r)χ(θ, φ) = E R(r)χ(θ, φ)
2M r ∂r 2 2M r2

The χ cancels from both sides to get just the radial problem
� � � �
h̄2 1 ∂2 λ
− r + + V (r) R(r) = E R(r) (2)
2M r ∂r 2 2M r2

1.2.1 Angular momentum eigenfunctions and eigenvalues


Let us first deal with the angular part.
�ˆ 2 has its own eigenfunctions and eigenvalues:
L

�ˆ 2 χ(θ, φ) = λχ(θ, φ)
L

This is a mathematical problem, and rather than solve it here we will state the results1

λ = l(l + 1)h̄2

for all integers l = 0, 1, 2.


Each of these eigenvalues l, is 2l + 1-fold degenerate - i.e. there are 2l + 1 different eigenfunctions
for each eigenvalue.

We can understand the degeneracy as follows. Since L�ˆ 2 commutes with each of L̂x , L̂y , L̂z , the χ
can also be chosen to be simultaneous eigenfunctions of one of those.
Conventionally these are chosen to be eigenfunctions of L̂z . On solving, we get the eigenfunctions
and eigenvalues to be

�ˆ 2 χ(θ, φ) = l(l + 1)h̄2 χ(θ, φ)


L
L̂z χ(θ, φ) = mh̄ χ(θ, φ)

Here l is a positive integer,


and m ranges from −l to l in integer steps (m = l, l − 1, l − 2, · · · , −l + 1, −l).
�ˆ 2 .
The 2l + 1 possible values of m give rise to the 2l + 1-fold degeneracy in eigenfunctions of L

1
The solutions are readily available in textbooks

7
Now let us see what these χ are. We have seen that L̂z = −ih̄∂/∂φ, and have solved this in the 2D
case. We thus know that the φ dependence is eimφ with m being an integer.
So we have χ(θ, φ) = Θ(θ)eimφ .
The Θ solution is not very straightforward, so I will just quote results. The solution is

Θ(θ) ∼ Plm (cos θ)

where Plm is the associated Legendre function, defined as


� �|m|
2 |m|/2 d
Plm (x) = (1 − x ) Pl (x)
dx

where Pl is the Legendre polynomial, given by the Rodrigues formula


� �l
1 d
Pl (x) = l (x2 − 1)l
2 l! dx

These angular solutions are very useful and famous, and altogether we have

χ(θ, φ) ∼ Plm (cos θ) eimφ ∼ Ylm (θ, φ)

(where I am using ∼ since there will be a normalization constant which I am not explicitly men-
tioning here).
The Ylm are known as the spherical harmonics, and the first few ones are
� �
1 3 3
Y00 = √ , Y10 = cos θ , Y1±1 = ∓ sin θ e±iφ (3)
4π 4π 8π

Notice that:
1. The Y00 state is spherically symmetric and we have L̂2 Y00 = 0 since Y00 is a constant.
2. Yl0 is always independent of φ.
3. The Ylm are also orthogonal to each other
� π � 2π
(Ylm
� ) Yl sin θ dθ dφ = δll� δmm�

∗ m
0 0

We can given an expression for all the spherical harmonics with m = l



(−1)l (2l + 1)!
Yl l = l (sin θ)l eilφ
2 l! 4π

8
Now we can return to the radial part. We have so far that (for central potentials)

ψ(r, θ, φ) = R(r) Ylm (θ, φ)

and that the radial part of the Schrodinger equation is


� � � �
h̄2 1 d2 l(l + 1)h̄2
− r + + V (r) R(r) = E R(r) (4)
2M r dr2 2M r2

Here, as in the 2D case, we see


1 d2
• the “kinetic” term is r dr 2
r instead of just d2 /dx2 .
• the potential has an additional term, thus the effective potential is

l(l + 1)h̄2
Ṽ (r) = V (r) +
2M r2

• the radial coordinate goes from 0 to ∞.


We can do one easy simplification by defining

U (r)
R(r) =
r
and thus reduce the radial problem to
� �
h̄2 d2
− + Ṽ (r) U (r) = E U (r)
2M dr2

which is now in a more familiar form to us.


We have to of course remember that the solution we obtain from this equation will not be the radial
wave function, we will have to multiply it by 1/r to get the radial wave function.
There is just one important part about this. Since r starts at r = 0, we have to pay special attention
to the behavior of our solution at that point. The solutions we get must lead to a R(r) that is non-
divergent at r = 0, so that our probability remains finite. This then means that U (r) must vanish
at r = 0, at least as fast as r, i.e. U (r) ∼ r α as r → 0, with α ≥ 1. This will be important for us
in selecting the acceptable wavefunctions.

9
1.3 Particle in a Spherical Cavity
Let us solve the following problem. A free particle of mass M in three dimensions is confined
inside a spherical cavity with an infinitely hard wall. Find its ground state wave function and
energy.
We’ll not worry about the normalisation for this wave function.
The potential is given as �
0, if 0 ≤ r ≤ a
V (r, θ, φ) =
∞, otherwise

Recall that our Hamiltonian for central potentials is

�ˆ 2
� �
h̄2 1 ∂2 L
H=− r + + V (r)
2M r ∂r2 2M r2

The TISE is written as


� �
−h̄2 d2 L̂2
r + + V (r) ψ(r, θ, φ) = Eψ(r, θ, φ)
2M r dr2 2M r2

We are asked to find the ground state wave function and energy. Since inside the sphere V (r) = 0,
and since in the spherically symmetric ground state l = 0, both the L̂2 term and the V (r) term will
vanish from the TISE. We will have the radial part left.
Let us first factorise ψ as ψ(r, θ, φ) = R(r)Ylm (θ, φ) and write the Schrodinger equation
� �
−h̄2 1 d2 l(l + 1)h̄2
r + R(r)Ylm (θ, φ) = ER(r)Ylm (θ, φ)
2M r dr2 2M r2

Here the angular dependence is written as the spherical harmonics Ylm , since they are eigenfunc-
tions of L̂2 , i.e. L̂2 Ylm (θ, φ) = l(l + 1)h̄2 Ylm (θ, φ).
Now putting l = 0, and dividing both sides by Ylm (θ, φ)

−h̄2 1 d2
rR(r) = ER(r)
2M r dr2
Putting in the substitution R(r) = U (r)/r

h̄2 d2 U
− = EU
2M dr2
Thus

d2 U 2M E 2 2M E
2
= − 2 U = −k U where k =
dr h̄ h̄

The solutions to this equation are U (r) = Aeikr + Be−ikr

10
Applying the boundary conditions
• The wavefunction R(r) should be finite at r = 0. This implies that U (r)/r → 0 as r → 0.
This is only true when U (r) → 0 as r → 0.
This means that 0 = Aeik0 + Be−ik0 = A + B, thus B = −A.
The solution is thus U (r) = A(eikr − e−ikr ) = A sin (kr), where the constants have been
absorbed into A.
• The wavefunction should vanish at the boundary of the sphere, i.e R(a) = 0. This gives us
U (a)/a = A sin (ka) = 0.
This gives kn = nπ/a where n = 1, 2, 3, ...,
and thus ground state energy (n = 1) is

k 2 h̄2 π 2 h̄2
E0 = =
2M 2M a2
and ground state � �
A πr
R(r) = U (r)/r = sin
r a

For the ground state Y00 = √1 , and the ground state ψ0 (r, θ, φ) = R0 (r)Y00 (θ, φ)

Thus the ground state wave function and energy are


� �
A πr π 2 h̄2
ψ0 = √ sin ; E0 =
r 4π a 2M a2
and the constant A will be determined from the normalization condition. (Note that the ground
state Ylm (θ, φ) is a constant, and so can be absorbed in the normalization constant in the above
answer.)

11

You might also like