Вы находитесь на странице: 1из 46

Chapter 4:

Harmonic Oscillator
1. Importance in physics
2. Algebraic method
3. Analytic method

2-1
MSE 227 – Applied Quantum Mechanics – Stefaan De Wolf
Harmonic oscillator
Let’s first quickly review the classic harmonic oscillator from chapter 1.

Classically We attach a mass 𝑚 to a spring with force constant 𝑘.


Let’s ignore friction.

d 2x
 The motion is governed by Hooke’s law F  kx  m 2
dt
 Remember, there is a negative sign as it is a restoring force,
proportional to the displacement 𝑥 from the equilibrium position.

Solution: x  t   A sin t   B cos t 

k
with  the (angular) frequency of oscillation.
m

The potential we get from the fact we are dealing here with a conservative force:

V 1 2 1
F   V  x  kx  m 2 x 2
x 2 2
 it’s graph is a parabola.
Harmonic oscillator

 Extremely important potential!

 Because practically any potential is


approximately parabolic, in the neighborhood
of a local minimum.

How do we see this?

 Formally, expand any 𝑉 𝑥 in a Taylor series about its minimum at 𝑥 = 𝑥0 :

 x  x0   n 
n

1
V  x   V  x0   V  x0   V '  x0  x  x0   V "  x0  x  x0   ...
2
Brook Taylor
n 0 n! 2 (1685-1731)

 n
V  x
V    x0  
n
we use here the notation
x n x  x
0

 Subtract now from this now 𝑉 𝑥0 (you can always add/subtract a constant to the
potential, as this does not change the force).
 Next, note that 𝑉′ 𝑥0 = 0 (𝑥0 is the minimum!);
 Finally, drop the higher-order terms (which are negligible as long as 𝑥 − 𝑥0 is small)
3
Harmonic oscillator
1
V  x   V "  x0  x  x0 
2
From this, we get
2
 This describes a simple harmonic oscillation around 𝑥0 , with an
effective spring constant 𝑘 = 𝑉" 𝑥0

 This is why the simple harmonic oscillator is so important:

Virtually any oscillatory motion is approximately


simple harmonic, as long as the amplitude is small.

Examples of systems that behave like harmonic oscillators on a microscopic scale are:
The vibrations of nuclei within diatomic molecules.
The vibrations of atoms in a crystalline solid about their equilibrium positions.

Applications of the harmonic oscillator can be found back in molecular spectroscopy, solid state
physics, nuclear structure, quantum field theory, quantum optics, quantum statistical
mechanics, etc.

4
Harmonic oscillator
Quantum mechanically
1
One needs to solve the Schrödinger equation for the potential V  x  m 2 x 2
2
As we have seen, it basically suffices to solve the time-independent Schrodinger equation:
2
d 2 1
 2
 m x   E
2 2

2 m dx 2
What is the solution of this eigenvalue problem?
 Two possible, but entirely different ways to solve this equation:
1. Algebraic method
 Developed by P.A.M. Dirac in 1925, based on earlier work by M. Born and N. Wiener.
 Clever algebraic manipulation
 Also referred to as the ladder operator method

2. Analytic method
 Also referred to as the power series method
 “Brute” force approach P.A.M. Dirac
(1902-1984)
 Advantage is that it can also be applied to many other types of potentials, 1933

including the Coulomb potential (see hydrogen calculation)

 This will yield then the stationary states, from which we then can construct the
general, time-dependent solution, 𝛹 𝑥, 𝑡 , as discussed in the previous chapter. 5
Harmonic oscillator – Algebraic method
2
d 2 1
 2
 m x   E
2 2

2 m dx 2

Let’s first rewrite this in a more suggestive form using operator notation:

1  2
Ĥ  E Hˆ  pˆ   m xˆ  
2
with
2m  
ℏ 𝑑
with, of course, 𝑝 ≡ and 𝑥 ≡ 𝑥, the momentum and position operators
𝑖 𝑑𝑥

The basic idea now is:


Let’s factorize the Hamiltonian operator 𝐻 (i.e. we want to write it as a product).

Remember, for numbers, factorization is straight forward, we can then write:


u 2  v 2   iu  v  iu  v 
Proof:  iu  v  iu  v   i 2u 2  iuv  ivu  v 2
(Numbers do commute: 𝑢𝑣 = 𝑣𝑢)
 u 2  iuv  iuv  v 2
 u 2  v2 6
Harmonic oscillator – Algebraic method

But here, 𝑥 and 𝑝 are not numbers! They are operators...


 An extremely important property of operators is that they generally do not commute:

𝑥 𝑝 is not equal to 𝑝𝑥

The term 𝑥 𝑝 − 𝑝𝑥 is called the commutator of 𝑥 and 𝑝.

 In general we write this as: 𝐴, 𝐵 ≡ 𝐴𝐵 − 𝐵𝐴

Let’s find out what is 𝑥, 𝑝 ….


 xˆ  x

For this we need to do the canonical substitutions:  
 ˆ
p 
i x

Next, let’s ‘tack’ a test function 𝑓 𝑥 onto 𝑥, 𝑝 to avoid making errors:


 xf 
df d
 xˆ, pˆ  f  x    xp ˆ ˆ  f  x  x
ˆˆ  px 
 i dx i dx 
7
Harmonic oscillator – Algebraic method

 df d 
  xˆ , pˆ  f  x    xp ˆ ˆ f  x  x
ˆˆ  px   
xf
 i dx i dx 

 df df 
  x  f  x   i f  x
i  dx dx 

  xˆ , pˆ   i This is known as the canonical commutation relation

The fact that position and momentum do not commutate is extremely important;
it is the root cause for all quantum ‘weirdness’!

8
Harmonic oscillator – Algebraic method

Let’s return to our goal to factorize the Hamiltonian of the Harmonic Oscillator:

1  2
Hˆ  pˆ   m xˆ  
2

2m  

Can we find an equivalent, factorized form of 𝑝2 + 𝑚𝜔𝑥 2

 For this, let’s evaluate the following operators:

 ipˆ  m x̂  and  ipˆ  m xˆ 

Actually, it will turn out to be advantageous to introduce a prefactor; let’s


call these new operators 𝑎+ and 𝑎− :

1
aˆ   ipˆ  m xˆ 
2 m 1
Or, compactly aˆ   ipˆ  m xˆ 
1 2 m
aˆ   ipˆ  m xˆ 
2 m 9
Harmonic oscillator – Algebraic method
Let’s find out their products 𝑎+ 𝑎− and 𝑎− 𝑎+
1
aˆ aˆ   ipˆ  m xˆ  ipˆ  m xˆ 
2 m
1  2
 pˆ   m xˆ   im  xp ˆ ˆ 
ˆˆ  px
2

2 m  
1  2 i
 pˆ   m xˆ     xˆ , pˆ 
2

2 m   2

Remember, the Hamiltonian for the harmonic oscillator was


1  2
Hˆ  pˆ   m xˆ  
2
1 ˆ i
 H   xˆ , pˆ  2m  
 2

We also had  xˆ, pˆ   i


1 ˆ i
 H i
 2

1 ˆ 1
 aˆ aˆ  H
 2 10
Harmonic oscillator – Algebraic method
1 ˆ 1
aˆ aˆ  H
 2
We can make the same calculation for the product 𝑎+ 𝑎−, where we finally obtain
1 ˆ 1
aˆ aˆ  H
 2
From these two products, 𝑎− 𝑎+ and 𝑎+ 𝑎−, we immediately obtain the following three identities

 aˆ , aˆ   1  1
Hˆ    aˆ aˆ   ˆ  1
H    a a  
ˆ ˆ
 2  2

 1
Compactly, Hˆ    aˆ aˆ  
 2

With this, we obtain two alternative forms of the Schrödinger equation 𝐻𝜓 = 𝐸𝜓 for the
harmonic oscillator:

 1  1
  a a    E
ˆ ˆ and   a a    E
ˆ ˆ
 2  2 11
Harmonic oscillator – Algebraic method
Having these expressions for 𝐻, 𝑎+ , and 𝑎− , we now calculate their commutation relations,
which will turn out to be very useful for later use.
 1
Hˆ    aˆ aˆ   ˆ ˆ    aˆ aˆ  1  aˆ    aˆ aˆ aˆ  1 aˆ 
 Ha
We had           
 2  2   2 

 1  1  1 
We also had Hˆ    aˆ aˆ    aˆ Hˆ   aˆ  aˆ aˆ      aˆ aˆ aˆ  aˆ 
 2  2  2 

  Hˆ , aˆ   Ha
ˆ ˆ  aˆ Hˆ   aˆ
  

Similarly, ˆ ˆ    aˆ aˆ  1  aˆ    aˆ aˆ aˆ  1 aˆ 
Ha           
 2  2 
 1  1 
and aˆ Hˆ   aˆ  aˆ aˆ      aˆ aˆ aˆ  aˆ 
 2  2 

  Hˆ , aˆ   Ha
ˆ ˆ  aˆ Hˆ    aˆ
  

We can write both relations again compactly as  Hˆ , aˆ     aˆ


  12
Harmonic oscillator – Algebraic method
We will now reveal the function of the operators 𝑎+ and 𝑎−

Firstly, if 𝜓 is an eigenfunction of our Harmonic Oscillator Schrödinger (HOS) equation,


we can always create a new function from this, such as 𝑎+ 𝜓.

Crucial step:

If 𝜓 is an eigenfunction of the HOS equation with eigenvalue (i.e. energy) 𝐸,


then this new function 𝑎+ 𝜓 actually also satisfies the Schrödinger equation,
now with eigenvalue (𝐸 + ℏ𝜔):

Ĥ  E  Hˆ  aˆ    E    aˆ

Proof: We had already that  Hˆ , aˆ   Ha


ˆ ˆ  aˆ Hˆ   aˆ
    

ˆ ˆ   aˆ Hˆ    aˆ   aˆ E   aˆ    E    aˆ 
 Ha Q.E.D.
     

13
Harmonic oscillator – Algebraic method
We have a similar relations for 𝑎− :
If 𝜓 is a wave function of our Harmonic Oscillator Schrödinger equation, we can always
create a new function from this, such as 𝑎− 𝜓.

Now again, what is particular is that this new function 𝑎− 𝜓 actually also
satisfies the Schrödinger equation, now with energy (𝐸 − ℏ𝜔):

Ĥ  E  Hˆ  aˆ    E    aˆ

 Hˆ , aˆ   Ha
ˆ ˆ  aˆ Hˆ    aˆ
Proof: We had already that     

ˆ ˆ   aˆ Hˆ    aˆ   aˆ E   aˆ    E    aˆ 
 Ha Q.E.D.
     

14
Harmonic oscillator – Algebraic method
Let us summarize this: so if 𝐻𝜓 = 𝐸𝜓,

then 𝐻 𝑎+ 𝜓 = 𝐸 + ℏ𝜔 𝑎+ 𝜓  𝑎+ is called the raising operator


then 𝐻 𝑎− 𝜓 = 𝐸 − ℏ𝜔 𝑎− 𝜓  𝑎− is called the lowering operator

 Sometimes they are also called the creation and destruction operators.

 Together, they are called the ladder operators

Note that it will turn out that 𝑎+ 𝜓𝑛 = 𝑐𝑛+1 𝜓𝑛+1

while 𝑎− 𝜓𝑛 = 𝑐𝑛−1 𝜓𝑛−1

Ground state

What if I keep on applying the lowering operator 𝑎− to 𝜓?


 Will we reach a state with negative energy? NO!

 There is actually a ‘lowest rung’ on the ladder,


for which
𝑎− 𝜓0 = 0
15
Harmonic oscillator – Algebraic method

We use 𝑎− 𝜓0 = 0 to determine the ground state 𝜓0

1  d 
 aˆ 0    m x  0  0
2 m  dx 

d 0 m
  x 0
dx
d 0 m
   xdx
0
m 2
 ln  0    x C
2

 m 2 
  0  A0 exp   x 
 2 

We need to normalize this Gaussian integral to obtain 𝐴0 .

16
Harmonic oscillator – Algebraic method Remember, Gaussian integral:

 dx
Let’s normalize this
F   2
exp ax
 
 m 2  
1  0 dx  A  exp  
2 2  
x dx

0
    F2    
exp   a  x 2
 y 2
dxdy
 
 x  r cos 
 
 1  A02 2 
 y  r sin 
m
F  2
  exp   ar 2
 rd dr
0 0

2
exp   ar 

 m   m 2 
14

 0  x    2
 exp  x 
  2a 0
 2 

F
a

17
Harmonic oscillator – Algebraic method
 What is the corresponding energy eigen state?

 Plug into Schrödinger equation: 𝐻𝜓0 = 𝐸0 𝜓0

 Use one of the H.O. operator forms of 𝐻, and realize that 𝑎− 𝜓0 = 0


(we can not get lower on the ladder than 𝜓0 )

 1
   aˆ aˆ   0  E0 0
 2

 1
   0   0  E0 0
 2
1
 E0  
2

18
Harmonic oscillator – Algebraic method
Excited states How do we get these?
1
 Apply repeatedly the raising operator 𝑎+ = −𝑖 𝑝 + 𝑚𝜔𝑥 on 𝜓0
2ℏ𝑚𝜔

 n  x   An  aˆ   0  x 
n
This yields

and
 1
En   n   
 2
Some remarks
1. By applying repeatedly the raising operator on 𝜓0 we can – in principle – construct all
the stationary states of the harmonic oscillator.
2. Each stationary state 𝜓𝑛 has it’s own normalization constant 𝐴𝑛 .
 Each time we need to normalize in principle 𝜓𝑛 .
 Note that this is different than for the infinite square well, where
each 𝜓𝑛 had the same normalization constant.

3. Note that this is Planck’s famous equidistant energy-level scheme (with energy spacings
equal to ℏ𝜔) to explain the black-body emission spectrum, which gave birth to quantum
19
mechanics! For this reason, some books talk about the ‘Planck oscillator’.
Harmonic oscillator – Algebraic method

Example 1.

Let’s calculate the state just above ground state

 m   m 2 
14
A1  d
  1  x   A1aˆ 0     m x   exp  x 
2 m  dx     2 

 m   m  m 2 
14
A1 
    2 x  m x  exp  x 
2 m    2   2 

 m  2m  m 2 
14

 A1   x exp   x 
   2 
Normalization:
 
m 2m  m 2 
1   1 dx  A   
2 2 2 2
x exp  x dx A

1
  
1

 A1  1
20
Harmonic oscillator – Algebraic method
We can even find an algebraic expression for the normalization.
We have aˆ n  cn 1 n 1  What is 𝑐𝑛+1 , if all states are normalized ?

Let’s normalize the above expression:

 cn 1 n 1 dx    aˆ n  aˆ n dx


2 *

1 𝑑
Let’s write this out: 𝑎+ = −ℏ 𝑑𝑥 + 𝑚𝜔𝑥
2ℏ𝑚𝜔
1  d n 
    aˆ n  dx  m   aˆ n  x n dx 
* *

2 m  dx 
(integration by parts)
 d  aˆ n  
*
1
   dx  n dx  m  x  aˆ n   n dx 
*

2 m  
1 𝑑
You can recognize 𝑎− = −ℏ 𝑑𝑥 + 𝑚𝜔𝑥
2ℏ𝑚𝜔
  aˆ  aˆ n   n dx   aˆ aˆ n* n dx
*

1 1
1  ˆ  * Remember: 𝑎− 𝑎+ = ℏ𝜔 𝐻 + 2

   H   n n dx
2 
1   * 1  1 1
          n n
        
* 2
 n
E  n n dx  n dx n 1 n dx
  2   2 2 21
Harmonic oscillator – Algebraic method

cn21   n 1 dx   n  1   n dx
2 2
So,

But, as said, all stationary states are normalized  cn 1  n  1

So: aˆ n  n  1 n 1 Similarly, one can find aˆ n  n n 1

As a consequence:  1  â 0
1 1
2  aˆ 1   aˆ   0
2

2 2
1 1 1
3  aˆ 2     0
3
ˆ
a
3 3 2
1 1 1 1
4  aˆ 3   aˆ   0
4
et cetera...
4 4 3 2

1  The normalization factor is 𝐴𝑛 = 1


n   aˆ   0 𝑛!
n
Clearly,
n! (in particular 𝐴1 = 1, as indeed obtained before)
22
Harmonic oscillator – Algebraic method

Just as in the case of the infinite square well, the stationary states of the harmonic oscillator
are orthogonal:

 m n dx   nm
 *



This can be proven, see Griffiths for those interested.

Note, from aˆ n  n  1 n 1 and aˆ n  n n 1

we derive the following expressions:


 So applying first 𝑎− and then 𝑎+ on any
aˆ aˆ n  naˆ n 1  n n stationary state reveals the value of its
quantum number 𝑛.
aˆ aˆ n  n  1aˆ n 1   n  1 n

From this, we get 𝑎− 𝑎+ − 𝑎+ 𝑎− = 1,


which is the commutation relation we found already before:

 aˆ , aˆ   1 23
Harmonic oscillator – Algebraic method
From the relations of 𝑎+ and 𝑎− we can derive a useful algebraic upward recursion relation
for the normalized stationary states:
1
By definition: aˆ   ipˆ  m xˆ  and aˆ 
1
 ipˆ  m xˆ 
2 m 2 m
1  d 
From normalization: aˆ n  n  1 n 1     m x  n  n  1 n 1
2 m  dx 
and
1  d 
aˆ n  n n 1    m x  n  n n 1
2 m  dx 
2m
 x n  n  1 n 1  n n 1 (eliminate the derivatives)

2m n 1
n  x n 1   n2
n n
Starting functions:

 m   m 2   m  2m  m 2 
14 14

0    exp   x  1    x exp   x 
    2     2 
The algebraic recursion relation is a very convenient way to compute numerical 24
values of the normalized energy eigen functions (no derivates to deal with!)
Harmonic oscillator – Algebraic method

Example 2.
Let’s find the expectation value of the potential energy of the 𝑛𝑡ℎ state in the
harmonic oscillator

1 1
Solution: V  m 2 xˆ 2  m 2   n* xˆ 2 n dx
2 2 

 We will solve this in a clever way using the ladder operators!

1 1
aˆ   ipˆ  m xˆ  aˆ   ipˆ  m xˆ 
2 m 2 m
Let’s rewrite this to express 𝑥 and 𝑝 as function of the ladder operators:

m
xˆ   aˆ  aˆ  pˆ  i  aˆ  aˆ 
2m 2

Here, we are interested in 𝑥 2 :

xˆ 2 
2m
 aˆ
2
  aˆ aˆ  aˆ aˆ  aˆ2  25
Harmonic oscillator – Algebraic method


 V 
4   *
n  ˆ
a 2
  ˆ
a 
ˆ
a   ˆ
a 
ˆ
a   ˆ
a 2
  n dx


2
But in this, we know that 𝑎+ 𝜓𝑛 is (apart from normalization) 𝜓𝑛+2 , which is orthogonal to 𝜓𝑛 .
2
Similarly, we know that 𝑎− 𝜓𝑛 is (apart from normalization) 𝜓𝑛−2 , which is orthogonal to 𝜓𝑛 .


 V    aˆ aˆ  aˆ aˆ  n dx
 *
n
4 
aˆ aˆ n  naˆ n 1  n n
Remember:
aˆ aˆ n  n  1aˆ n 1   n  1 n


  2n  1   n* n dx
4 

1  1
 n 
2  2

 The expectation value of the potential energy is exactly half the total energy.
 The other half is of course kinetic.
 This is a peculiarity of the harmonic oscillator 26
Harmonic oscillator – Analytic method
Let’s now see if we can come to the same results via the analytical way
A big advantage of the analytic way is that the strategy we will follow can also be applied to
many other types of potentials, including the Coulomb potential (see hydrogen calculation).

General strategy:
1. Clean up notation
2. Get rid of asymptotic terms
3. Solve the ‘reduced’ Schrodinger equation by the so-called ‘series’ method of Frobenius.
4. Construct the stationary states from the product of the asymptotic terms and the
Frobenius solution

Let’s first remind us the Schrödinger equation again for the H.O.:
2
d 2 1
 2
 m x   E
2 2

2 m dx 2
𝑥
Let’s clean up notation by defining the dimensionless position variable 𝑄 = ,
𝐿

with 𝐿 = , a characteristic length.
𝑚𝜔

d 2

dQ 2
  Q 2
 K 
1 2E
Here, 𝐾 is the energy, in units ℏ𝜔: K 27
2 
Harmonic oscillator – Analytic method

d 2
dQ 2
  Q 2
 K 

Our problem is to solve this equation, and in the process obtain the ‘allowed’ values of 𝐾.
 Let’s first look at the asymptotic values

Note that this is equivalent to looking at the boundary conditions as we did before.
Problem is, this is a ‘soft’ potential, so the boundary values are not ‘defined’.
 We look here rather at the asymptotic behavior.

Note that at very large 𝑄 (which is to say, at very large 𝑥), the 𝑄2 term completely dominates
over 𝐾, so in this regime we have.
d 2
 2
 Q 2

dQ
This equation has as approximate solution (verify this!)

  Q   A exp  Q 2 2   B exp  Q 2 2 
The 𝐵 term clearly blows up at |𝑥| → ∞, so the asymptotic form is

  Q   ... exp  Q 2 2  at large 𝑄 28


Harmonic oscillator – Analytic method
This suggests we ‘peel off’ the asymptotic exponential part in the general equation:

  Q   h  Q  exp  Q 2 2 
Note that this is an exact solution!

Now we hope that the equation simplifies for ℎ 𝑄 , compared to that for 𝜓 𝑄 !

d  dh 
Let’s differentiate this equation:   Qh  exp  Q 2 2 
dQ  dQ 

d 2  d 2 h 
2
  2
 2Q
dh
  Q 2
 1 h  exp  Q 2 2 
dQ  dQ dQ 

So the Schrödinger equation becomes:

d 2h
dQ 2
 2Q
dh
dQ
  Q 2
 1 h   Q 2
 K h

d 2h dh
 2
 2Q   K  1 h  0
dQ dQ 29
Harmonic oscillator – Analytic method

Let’s try solutions in the form of a polynomial:



h  Q   A0  A1Q  A2Q  A3Q  ...   A j Q j
2 3
Ferdinand Georg Frobenius
(1849-1917)
j 0

 This is the so-called Frobenius method for solving a differential equation.

We differentiate this term by term:



dh
 A1  2 A2Q  3 A3Q  ...   jA j Q j 1
2

dQ j 0


d 2h
dQ 2
 2 A2  2  3 A3Q  3  4 A4 Q 2
...  
j 0
 j  1 j  2  A j  2 Q j

We plug these in our ‘reduced’ Schrodinger equation:

d 2h dh
 2Q   K  1 h  0
dQ 2 dQ

   j  1 j  2  Aj  2  2 jA j   K  1 A j Q j  0
30
j 0
Harmonic oscillator – Analytic method
In any power series, each coefficient is unique.
Hence, the coefficient of each power of 𝑄 must vanish in the above equation.

 j  1 j  2  Aj  2  2 jAj   K  1 Aj  0
We can rewrite this as a recursion formula:

2 j  1  K 
Aj  2  Aj
 j  1 j  2 
 Note that this formula is actually entirely equivalent to the Schrödinger equation!

Starting with 𝐴0 , it generates all even-numbered coefficients:

1 K 

 5 K

 5  K 1  K  etc.
A2  A 0
A4 A 2 A 0
2 12 24
Starting with 𝐴1 , it generates all odd-numbered coefficients:

3  K  A A5 
7  K  A 
 7  K  3  K  A etc.
A3  1 3 1
6 20 120 31
Harmonic oscillator – Analytic method

We write the complete solutions as h  Q   heven  Q   hodd  Q 

where heven  Q   A0  A2Q 2  A4Q 4  ...

is an even function, built on 𝐴0

and hodd  Q   A1Q  A3Q 3  A3Q 3  ...


is an odd function, built on 𝐴1

So this recursion formula determines ℎ 𝑄 fully in terms of two arbitrary constants,


𝐴0 and 𝐴1 , which is just what one would expect from a second-order differential equation.

Not all solutions so obtained are normalizable, however!

32
Harmonic oscillator – Analytic method
2𝑗+ 1−𝐾
We see this by taking another look at the recursion formula, 𝐴𝑗+2 = 𝐴𝑗 :
𝑗+1 𝑗+2
we find that at very large 𝑗,

2 4 1
 Aj  2  Aj  Aj 2  Aj  2  ...
j j  j  2  j 2  j 2  1
C
 Aj 
 j 2 !
This means that towards large 𝑄, where high powers dominate, the power series becomes

Q  C exp  Q 2 
1 1 2j
h Q   C  Q j  C
 j 2 !  j !
However, this implies that for large 𝑄

  Q   h  Q  exp  Q 2 2   C exp   Q 2 2  Q 2   C exp   Q 2 2 

So the wave function explodes for |𝑄| → ∞ and, hence |𝑥| → ∞


How is this possible??! 33
Harmonic oscillator – Analytic method
There is only one way out of this:
the power series must terminate somewhere to yield a normalizable solution.

 There must be ‘highest’ 𝑗-value (let’s call it 𝑛), for which the recursion formula
starts to spit out 𝐴𝑛+2 = 0:

2n  1  K 
An  2  An  0
 
n  1 n  2 
 This will truncate either the series ℎ𝑒𝑣𝑒𝑛 or ℎ𝑜𝑑𝑑

The other series must be zero from the start: 𝐴1 = 0, if 𝑛 is even; 𝐴0 = 0, if 𝑛 is odd.

 𝐾 = 2𝑛 + 1 with 𝑛 = 0,1,2, ….

2𝐸
Remember, 𝐾 ≡ , so
ℏ𝜔

 1
 E  n   Just as we found with the algebraic method!
 2 34
Harmonic oscillator – Analytic method
Let’s now find the stationary states of the harmonic oscillator.
2 j  1  K  2  j  n
With 𝐾 = 2𝑛 + 1, the recursion formula reads: Aj  2  Aj  Aj
 j  1 j  2   j  1 j  2 
2j
Ground state: 𝑛 = 0  Aj  2  Aj
 j  1 j  2 
 To obtain 𝐴2 , we set 𝑗 = 0, which yields 𝐴2 = 0
 With this all higher order even terms 𝐴𝑖 will also be zero.
 We must pick 𝐴1 = 0 to kill ℎ𝑜𝑑𝑑 .

 h0  Q   A0   0  Q   A0 exp  Q 2 2 

2  j  1
First excited state: 𝑛 = 1  Aj  2  Aj
 j  1 j  2 
 To obtain 𝐴3 , we set 𝑗 = 1, which yields 𝐴3 = 0
 With this all higher order odd terms 𝐴𝑖 will also be zero.
 We must pick now 𝐴0 = 0 to kill ℎ𝑒𝑣𝑒𝑛 .

 h1  Q   A1Q   1  Q   A1 Q exp  Q 2 2  35
Harmonic oscillator – Analytic method
2  j  2
Second excited state: 𝑛 = 2,  Aj  2  Aj
 j  1 j  2 
 To obtain 𝐴2 , we set 𝑗 = 0, which yields now 𝐴2 = −2𝐴0
 To obtain 𝐴4 , we set 𝑗 = 2, which yields 𝐴4 = 0
 With this all higher order even terms 𝐴𝑖 will also be zero.
 We take again 𝐴1 = 0 to kill ℎ𝑜𝑑𝑑 .

 h2  Q   A0 1  2Q 2 

  2  Q   A0 1  2Q 2  Q exp  Q 2 2 

and so on…

 In general, ℎn 𝑄 will be a polynomial of degree 𝑛 in 𝑄, involving even powers only if


𝑛 is even, and odd powers only if 𝑛 is odd.

Apart from the overall factor (𝐴0 or 𝐴1 ), they are the so-called Hermite polynomials, 𝐻n 𝑄 .

36
Harmonic oscillator – Analytic method
The first few Hermite polynomials are H0 Q   1
H1  Q   2Q
H 2  Q   4Q 2  2
H 3  Q   8Q 3  12Q Charles Hermite
(1822-1901)

H 4  Q   16Q 4  48Q 2  12
H 5  Q   32Q 5  160Q 3  120Q

We note that these polynomials can be obtained in different ways, here is


one via upward algebraic recursion (without proof):
H n  Q   2Q  H n 1  Q   2  n  1 H n  2  Q 
By tradition, the arbitrary multiplicative factor is chosen so that the coefficient of the highest
power of 𝑄 is 2𝑛 .
With this convention, the normalized stationary states for the harmonic oscillator are

 m   Q2 
14
1
 n  x    H n  Q  exp   
𝑥
with 𝑄 = 𝐿 and 𝐿 =

   n
2 n!  2  𝑚𝜔

37
(without proof for the normalization).
Harmonic oscillator – Classic vs. Quantum case
Let’s briefly compare the classic and quantum harmonic oscillator
The right figure shows a superposition of the
energy levels onto the parabolic potential

The points 𝑥𝑛 , where 𝑉 𝑥 = 𝐸𝑛 are the points


where, classically, the particle motion would turn
around; they are called classical turning points
1  1
V  m 2 x 2 and En   n   
2  2

 xn  2n  1L

𝐿 is the same characteristic length 𝐿 = we introduced earlier for the harmonic oscillator.
𝑚𝜔

𝐿 represents the spatial size of state 𝜓0 of the system and may therefor be regarded as the
natural unit of length for the harmonic oscillator.

Lets’ develop some intuition for this:

For electrons bound sufficiently tightly that their oscillation frequency is in the visible light
range, say 𝑓 = 1015 𝐻𝑧, we find that 𝐿 = 0.135 𝑛𝑚, roughly the size of a small atom.
38
Harmonic oscillator – Classic vs. Quantum case

7
E3   odd
2

5
E2   even
2

3
E1   odd
2

1
E0   even
2

Fig. Lowest 4 normalized eigen functions and their associated probability densities

In classical mechanics, a particle with energy 𝐸𝑛 cannot travel beyond the classical turning points.
In quantum mechanics, the wave function becomes evanescent there, representing a tunneling
of the particle into the potential barrier.
 We have a certain probability of finding a particle outside the classically forbidden zones!
Also, for all odd functions, the probability of finding the particle at the origin is zero! 39
Harmonic oscillator – Classic vs. Quantum case

Classical turning points The probability density distribution, averaged


over the rapid spatial oscillations, approaches
the time-averaged classical probability density
distribution.

The time-averaged classical probability


distribution for finding an oscillating particle at
position 𝑥 is inversely proportional to the
magnitude of the velocity 𝑉 𝑥 at that point;

  x 
 v  x
Consider a classical sinusoidal oscillation: x  a sin t
dx
v   a cos t  a 1  sin 2 t   a 2  x 2
dt
1
   x   Dashed line in figure
 a x
2 2

 Again, an example of the correspondence principle of Bohr. 40


Harmonic oscillator – Stationary states: a reminder

As for the infinite square well, we have again the following properties of the stationary states:

1. States are alternatively even and odd, with respect to center of the well:

𝜓1, 𝜓3, 𝜓5,… are even; 𝜓2, 𝜓4, 𝜓6,… are odd
 Valid for every symmetric potential
2. With increasing energy, number of nodes goes up by 1.
 Valid for every potential

3. All states are orthonormal.  Valid for every potential

4. All states 𝜓𝑛 for a complete set, any other function 𝑓(𝑥) can be written as a
linear combination of them:  Valid for every potential

 
 m 
14
1  m   m 2 
f  x    cn n  x   cn   H n  x  exp   x 
n 1 n 1   n
2 n!    2 

41
Harmonic oscillator – Stationary States
How to calculate 𝑐𝑛 ?  Same as before, use ‘Fourier’s trick’


 Multiply both sides by 𝜓𝑚 , followed by integration

 

  x  f  x dx   c   x    x dx  c  cm


* *
m n m n n nm
n 1 n 1

cn   n  x  f  x dx
*

 1
En   n   
 2
Putting things together for the harmonic oscillator:

Stationary states:
 m 
14
1  m   m 2    1 
 n  x, t     H n  x  exp   x  exp  i  n   t 
  2n n !    2    2 
General solution:

 m 
14
1  m   m 2    1 
  x, t    cn   H n  x  exp   x  exp  i  n   t 
n 0   2n n !    2    2 
42
Harmonic oscillator – Oscillating states
As with all stationary states, the probability density distributions for the harmonic oscillator
bound states are time-independent.

 therefore, they do not fit the classical picture of a particle oscillating back and forth.

To construct oscillating states, we must look at linear super positions of bound states with
different energies, with time-dependent coefficients chosen in such a way that the
superposition satisfies the Schrödinger wave equation.

Simplest case: a linear superposition of the states 𝑛 = 0 and 𝑛 = 1 for the harmonic oscillator.

 i 
Time-dependent wave functions:  0  x, t    0  x  exp   t 
 2 
 3i 
1  x, t    1  x  exp   t 
 2 
Any linear superposition of these wave functions will also satisfy the Schrödinger wave
equation:
  x, t   c0  0  x, t   c11  x, t 

Let’s assume the chosen coefficients 𝑐0 and 𝑐1 to be real. 43


Harmonic oscillator – Oscillating states
The probability density is easily found:
  x, t   c02 02  x   c12 12  x   2c0c1 0  x  1  x  cos t
2

The wave packet will be normalized if: 1  c02  c12


(remember, for normalization we integrate all over space and the states are orthogonal)

𝑡=0 The left figure shows the probability density distribution


for a wave packet with 𝑐0 = 𝑐1 = 1 2, at 𝑡 = 0, after
a quarter oscillation cycle and half an oscillation cycle

Oscillation amplitude
 calculation of the center of probability
t   2 (= expectation value of the position)

x     x  xdx


Note: 𝜓0 is an even function, while 𝜓1 is odd


t    they will not contribute

 
x  2c0 c1    0  x  1  x  xdx  cos t
   44
Harmonic oscillator – Oscillating states

Let’s now plug in the actual expressions now for the stationary states:

Remember:  0  Q   A0 exp   Q 2 2 

  0  x   A0 exp   x 2 2 L2 

 1  Q   A0 2Q exp   Q 2 2 

  1  x   A0 2 x L exp   x 2 2 L2 

  1  x   2 x L 0  x 

 
 x  2c0 c1    0  x  1  x  xdx  cos t
  
2c0 c1 L  

    1  x  dx  cos t
2

2   
L
 cos t
2 45
Harmonic oscillator – Oscillating states
We can generalize to linear super positions of the generalized form, as already discussed:


  1 
  x, t    cn n  x  exp  i  n   t 
n 0   2 

The coefficients are subject to the constraint that the overall superposition state has to
be normalized.

In this way, an infinite variety of oscillating states can be constructed, with a far greater
richness of oscillation behavior than in classical mechanics.

46

Вам также может понравиться