You are on page 1of 81

Chapter 3:

Time-independent Schrödinger equation


1. Stationary states
2. Infinite square well in 1-D
3. Infinite square well in 3-D
Density of States
4. Quantum dot, quantum wire, quantum well

3-1
MSE 227 – Applied Quantum Mechanics – Stefaan De Wolf
Time–independent Schrödinger equation – Stationary states
How do you actually get Ψ 𝑥, 𝑡 ?
 We need to solve the time-dependent Schrödinger equation, for a specific potential 𝑉 𝑥, 𝑡 .

 2
2
i   V  x, t  
t 2m x 2

We will usually assume that 𝑉(𝑥, 𝑡) is independent of 𝑡, so it becomes 𝑉(𝑥).


We now use the method of separation of variables, where we look for solutions that are simple
products.

Let’s propose the following solution:

𝜓 𝑥 (‘psi’) is only a function of 𝑥


  x, t     x    t 
𝜑 𝑡 (‘phi’) is only a function of 𝑡

Important remark: In principle, such separable solutions represent only a fraction of all
possible solutions to the Schrödinger equation…

However, we will see that we can write any solution to the Schrödinger
equation as a linear combination of separable solutions!
Time–independent Schrödinger equation – Stationary states
 2
2
We had i   V  x, t  
t 2m x 2

  x, t     x    t 

 d  2  d 2
  and  2
t dt x 2 dx
𝜕
Note: As 𝜓 𝑥 is only a function of 𝑥, the partial derivative ( )
𝜕𝑥
𝑑
becomes a total derivative ( ).
𝑑𝑥

d 2
d 2
i   2
  V
dt 2m dx
(divide by 𝜓𝜑)
1 d 2
1 d 2
i  V
 dt 2m  dx 2

The LHS is now only a function of 𝑡; the RHS is only a function of 𝑥.

 Both sides must be equal to a constant; let’s call this separation constant 𝐸.
Time–independent Schrödinger equation – Stationary states

1 d 2
1 d 2
i  V  E
 dt 2m  dx 2

 1 d
 i E
  dt
 2
 1 d 2

V  E
 2m  dx 2

 d iE
   
dt
 2 2
 d
 V  E
 2m dx 2

 Separation of variables has turned 1 partial differential equation into 2 ordinary differential
equations!

4
Time–independent Schrödinger equation – Stationary states
1. First equation:
d iE  iEt 
     t   C exp   
dt  
𝑖𝐸𝑡
Here, constant 𝐶 can be dropped, as it can be absorbed in 𝜓  𝜑 𝑡 = exp − ℏ
 𝜑 𝑡 is a time-dependent phase factor.
 Note that 𝜑 ∗ 𝜑 = 1

 Note that we could also write this as 𝜑 𝑡 = exp −𝑖𝜔𝑡 ,


𝐸
with 𝜔 = ℏ , taken from 𝐸 = ℏ𝜔, the Planck-Einstein relation.

2. The second equation is the time-independent Schrödinger equation:


2
d 2
  V  E
2m dx 2

To solve this equation, we need to know 𝑉 𝑥 , which will be problem-dependent.

Before starting to solve this equation, let’s make some important remarks about what is so
special about separable solutions.
5
Time–independent Schrödinger equation – Stationary states
Remark 1: Separable solutions are stationary states
Note that the wave function Ψ 𝑥, 𝑡 of a stationary state obviously does depend on 𝑡:
 iEt 
  x, t     x    t     x  exp   
 
However, its probability density does not:

 iEt   iEt 
  x, t     x, t    x, t     x  exp    x  exp       x
2 * * 2

   

The same is true for the expectation value of any dynamic variable 𝑄 of a stationary state:

Qˆ    t  Qˆ   t     *  x, t Qˆ   x, t  dx    *  x   *  t  Qˆ   x    t   dx

(any operator 𝑄 only depends on 𝑥 and/or 𝑝, so it will not change 𝜑 𝑡 ).


(we can bring any time-dependent term in front of the integral).

 Qˆ   *  t    t   *  x  Qˆ  x  dx

 Qˆ   *  x  Qˆ  x  dx 6
Time–independent Schrödinger equation – Stationary states
In Dirac notation, this statement would be

Qˆ    t  Qˆ   t    Qˆ 

Note: As mentioned, in Dirac notation we only write the essential information, usually independent
from a specific representation (such as position or momentum representations).

Please keep in mind this is only true for stationary states, not for their linear combinations (see
later)!

In short, nothing ever ‘happens’ in a stationary state, all expectation values are constant in time.

For example, 𝑥 is constant in time; 𝑝 = 0 (prove this!).

7
Time–independent Schrödinger equation – Stationary states
Remark 2: Separable solutions are states of definite total energy
Classic physics
Hamiltonian = total energy = kinetic energy + potential energy:
W.R. Hamilton
p2
H  x, p   V  x
(1805-1865)

2m
Quantum mechanics
We obtain the Hamiltonian operator 𝐻 via the following canonical substitutions:

x̂  x and p̂ 
i x
22
 Hˆ   V  x
2m x 2

We can write then the time-independent Schrödinger equation as 𝐻𝜓(𝑥) = 𝐸𝜓(𝑥)

In Dirac notation, we write this as Ĥ   E 


In this, 𝐻 operates on |𝜓 ; 𝐸 is a constant.
Algebraically, 𝐻 does not change the direction of ket |𝜓 , but only multiplies its length with 𝐸.
Obviously, this is an eigenvalue equation; |𝜓 are the eigenfunctions, 𝐸 are the eigenvalues.
8
Time–independent Schrödinger equation – Stationary states
Let’s calculate the expectation value of the total energy of state |𝜓 :

Hˆ   Hˆ 

(Use 𝐻 |𝜓 = 𝐸 |𝜓 )

 Ĥ   E 

(𝐸 is a constant)
 Ĥ  E  
(normalization: 𝜓|𝜓 = 1)
 Ĥ  E

 So the separation constant 𝐸 turns out to be the expectation value of the energy (hence its
name)!

Let’s also calculate Hˆ 2   Hˆ 2 

9
Time–independent Schrödinger equation – Stationary states
Let’s also calculate the expectation value of the square of the total energy of state |𝜓 :

Hˆ 2   Hˆ 2 

 Hˆ 2   HH
ˆˆ 

(Use 𝐻 |𝜓 = 𝐸 |𝜓 )
 Hˆ 2   HE
ˆ 
(𝐸 is a constant)
 Hˆ 2  E  Hˆ 
( 𝐻 = 𝐸)
 Ĥ 2
E 2

Combined, we find that the standard deviation of 𝐻 for stationary state is:

2
 Hˆ  Hˆ 2  Hˆ  E2  E2  0

 A separable solution has the property that every measurement of the total energy is certain
to return the value 𝐸.
10
Time–independent Schrödinger equation – Stationary states – Linear combinations
Remark 3: The general solution, 𝛹 𝑥, 𝑡 , of the time-dependent Schrödinger equation is a linear
combination of separable solutions.
2
d 2
Time-independent Schrödinger equation  2
 V  E
2m dx

 This yields a collection of solutions: 𝜓0 𝑥 , 𝜓1 𝑥 , 𝜓2 𝑥 ,…


Each of these solutions has an associated value of the separation constant: 𝐸0, 𝐸1, 𝐸2,…

𝜓0 𝑥 → 𝐸0,
𝜓1 𝑥 → 𝐸1,
𝜓2 𝑥 → 𝐸2,

We call these solutions 𝜓𝑖 𝑥 the energy eigenstates, with associated energy eigenvalues 𝐸𝑖
In general, there may be infinitely many eigenvalues and corresponding eigenstates.

If 𝑘 eigenstates correspond to the same energy eigenvalue 𝐸𝑙, then 𝐸𝑙 is said to be 𝒌-fold
degenerate.

11
Time–independent Schrödinger equation – Stationary states – Linear combinations

There is a different wave function for each allowed energy:

 iE t 
 0  x, t    0  x  exp   0 
 
 iE t 
1  x, t    1  x  exp   1 
 
etc…

These wave functions Ψ𝑛 𝑥, 𝑡 are the solutions of the time-dependent Schrödinger equation:
 2
2
i  V  x 
t 2m x 2

Now, any linear combination of these solutions of is also a solution of this equation:


 iE t 
  x, t    cn n  x  exp   n 
n 0  

The coefficients 𝑐𝑛 are called the expansion coefficients.


12
Time–independent Schrödinger equation – Stationary states – Linear combinations
Let’s prove this by substituting this sum in the time-dependent Schrodinger equation:

 i2 
 iEnt    iEnt 
LHS: i
t
 
n 0
c E 
n n n  x  exp 

   n n n


n 0
c E   x  exp 



2
2 
 2
 2 n   iEnt 
RHS:   V  x     cn    V  x  n  exp   
2m x 2 n 0  2 m  x 2
  

 iE t 
  cn En n  x  exp   n 
n 0  
 Both sides are indeed the same.

The coefficients 𝑐𝑛 are found by fitting to initial conditions for the problem at hand.

 Usually the wave function at 𝑡 = 0, Ψ 𝑥, 0 will be given as initial condition.

13
Time–independent Schrödinger equation – Stationary states – Linear combinations
Note that it is common practice to normalize all stationary states 𝜓𝑛, for all 𝑛:

  n* n dV  1
all _ space

Moreover, all stationary states are orthogonal.

Two states 𝜓𝑛 and 𝜓𝑚 are orthogonal when they satisfy


   r 0
* 3
m n d
with the integral going over all space

   r   nm
* 3
Orthogonality + normalization: m n d

0 if mn
Kronecker delta  nm  
1 if mn Leopold Kronecker
(1823-1891)

 We say that all 𝜓𝑖 ‘s orthonormal

14
Time–independent Schrödinger equation – Stationary states – Linear combinations
General orthogonality proof:
Two energy eigenfunctions 𝜓𝑛 and 𝜓𝑚 belonging to different energy eigenvalues 𝐸𝑛 and 𝐸𝑚
are necessarily orthogonal.

 Write Schrödinger equation for 𝜓𝑛, then multiply with complex conjugate of 𝜓𝑚:
2
  m*  2 n  m* V  r  n  En m* n
2m
 Similarly, write Schrödinger equation for complex conjugate of 𝜓𝑚, multiply with 𝜓𝑛:
2
  n  2 m*  nV  r  m*  Em n m*
2m
Let’s now subtract term-by-term both equations.

(LHS, the potential-energy terms cancel out)


2

2m
 *
m  2 n  n  2 m*    En  Em  m* n
(LHS, product rule derivatives)
2
    m*  n    m*  n     n  m*    m*  n    En  Em  m* n
2m  
15
Time–independent Schrödinger equation – Stationary states – Linear combinations
2
    m*  n  n  m*    En  Em  m* n
2m
(integrate both sides over space)
2 

 

  *
m  n   n  *
m   En  Em    m* n dV
2m 


(Remember, 𝜓 r → 0 when r → ∞)

 0   En  Em    m* n d 3r


 If 𝐸𝑚 ≠ 𝐸𝑛   m* n d 3 r  0 Q.E.D.

16
Time–independent Schrödinger equation – Stationary states – Linear combinations
In Dirac notation, we often write the stationary states 𝜓0 𝑥 , 𝜓1 𝑥 , 𝜓2 𝑥 ,… simply as |0 , |1 ,
|2 ,…

We write the orthonormality condition as then as

m n   mn

To keep notation clean, we could write the (time-independent) linear combinations then as

   cn n
n 0

We introduce time dependence as follows:



  t    cn  t  n
n 0

In this case, we assume that the time-dependent phase-factors are absorbed in each 𝑐𝑛 .

 iE t 
cn  t   cn exp   n 
 
17
Time–independent Schrödinger equation – Stationary states – Linear combinations
In a nutshell: General strategy to solve any quantum-mechanical problem
Given: a (time-independent) potential, 𝑉 𝑥 , and a starting wave function Ψ 𝑥, 0 .
Challenge: find the wave function Ψ 𝑥, 𝑡 .

Methodology:
1) Solve the time-independent Schrödinger equation:
2
d 2
 2
 V  E
2m dx
 This yields an infinite collection of solutions called the stationary states
[𝜓0 𝑥 , 𝜓1 𝑥 , 𝜓2 𝑥 , … ], each with its own associated energy (𝐸0, 𝐸1, 𝐸2,…).
Normalize all 𝜓𝑛 𝑥 .
2) Fit Ψ 𝑥, 0 (after normalization) by appropriate choice of coefficients (𝑐0, 𝑐1, 𝑐2,…).
Always possible! 
  x, 0    cn n  x 
n 0

3) Construct Ψ 𝑥, 𝑡 by ‘tacking’ to each term its characteristic time dependence,



 iEnt  
  x, t    cn n  x  exp      cn  n  x, t 
n 0   n 0 18
Time–independent Schrödinger equation – Stationary states – Linear combinations
Recall: The separable solutions Ψ𝑛 𝑥, 𝑡 are called the stationary states:
 iE t 
 n  x, t    n  x  exp   n 
 
 These solutions are obviously time dependent.
 However, all probabilities and expectation values of these states are independent
of time.
This is why we call them stationary states!

!!This is not the case for the general solution 𝛹 𝑥, 𝑡 !!



 iEnt  
  x, t    cn n  x  exp      cn  n  x, t 
n 0   n 0
 In the general solution, each stationary state 𝛹𝑛 has a different energy value 𝐸𝑛 .

 Hence, the exponentials usually don’t cancel out when calculating |𝛹 𝑥, 𝑡 |2

For the general solution, usually all probabilities and expectation


values of dynamic variables are time dependent.
19
Time–independent Schrödinger equation – Stationary states – Linear combinations
Example 1: A particle starts in a superposition of only two stationary states:

  x, 0   c1 1  x   c2 2  x 
• What is the wave function Ψ 𝑥, 𝑡 at subsequent times?
 iE t   iE t 
   x, t   c1 1  x  exp   1   c2 2  x  exp   2 
   
• What is the probability density of Ψ 𝑥, 𝑡 ?

Let’s assume 𝑐1, 𝑐2 and 𝜓1, 𝜓2 are real

20
 This is not a stationary state, but oscillating in time…
Time–independent Schrödinger equation – Stationary states – Linear combinations
For any linear superposition of stationary states we have that
the probability density depends only on the energy differences of the different states.

 The absolute energies never show up in the probability density Problem 2.1 Problem 2.2

Remark that indeed the probability density of a linear combination may fluctuate over time,
however, its normalization is a constant in time (as it of course it should)!
That is, by integrating the probability densities over space, their time-dependencies disappear.

 This is because of the orthogonality of the stationary states.

21
Time–independent Schrödinger equation – Stationary states – Linear combinations
Let’s prove that the normalization of any linear superposition is a constant in time:

 iEnt 
We had:   t    cn  t  n with cn  t   cn exp   
n 0  
 
   t    t    cm*  t  cn  t  m n
m0 n 0

  (Recall, 𝑚|𝑛 = 𝛿𝑚𝑛 )


   t    t    cm*  t  cn  t   mn
m0 n 0

   t    t    cn  t 
2

n 0


(absolute value of all phase factors is 1)
   t    t    cn
2

n 0

If |Ψ is normalized, then, moreover, at any time


 cn  1
2
Q.E.D.
n 0

22
Time–independent Schrödinger equation – Stationary states – Linear combinations
Exploiting orthonormality of the state kets, it is also easy to find the 𝑐𝑛 coefficients:

We had:   t    cn  t  n
n 0

To find for instance coefficient 𝑐𝑘 simply take the inner product of |Ψ 𝑡 with bra 𝑘|:
 
k   t    cn  t  k n   cn  t   kn  ck  t 
n 0 n 0

Generally: cn (t )  n  (t ) |𝑛

|𝑓
Note that 𝑛|𝛹(𝑡) is a scalar.

We can thus write :  t    n n  t  |𝑛 𝑛|𝑓
n 0

In this, |𝑛 𝑛| can be seen as a projection operator.

When |𝑛 𝑛| operates on a ket |𝑓 , it extracts from |𝑓 how much of |𝑛 it contains (figure).



From this, we obtain also the identity operator 1: n
n 0
n 1̂
23
Time–independent Schrödinger equation – Stationary states – Linear combinations

As an example, let’s assume we only have three stationary states:

 So, |𝑛 𝑛| indeed acts as projection operator on the vector |𝑓 .


It gives that part of |𝑓 lying in the |𝑛 direction.

Note: Is |𝑓 normalized in this 3-D example? 24


Time–independent Schrödinger equation – Stationary states – Linear combinations
To gain further insight in the meaning of the coefficients 𝑐𝑛 , let’s also calculate H of a linear
combination |Ψ 𝑡 :

  t    cn  t  n
n 0

 
 Hˆ    t  Hˆ   t    cm*  t  cn  t  m Hˆ n
m0 n 0

(Recall, all |𝑛 are stationary states, satisfying 𝐻 |𝑛 = 𝐸𝑛 |𝑛 )


 
 Hˆ   cm*  t  cn  t  m En n
m0 n 0

 
 Hˆ   cm*  t  cn  t En mn
m0 n 0


 Hˆ   cn En
2

n 0

 The time dependence disappeared due to the orthogonality of the basis kets |𝑛 .

25
Time–independent Schrödinger equation – Stationary states – Linear combinations
We can again also calculate H 2 for such a linear combination:

  t    cn  t  n
n 0
 
 Hˆ 2
   t  Hˆ   t    cm*  t  cn  t  m Hˆ 2 n
2

m0 n 0

(Recall, all |𝑛 are stationary states, satisfying 𝐻 |𝑛 = 𝐸𝑛 |𝑛 )


 
 Hˆ 2   cm*  t  cn  t  m HE
ˆ n
n
m0 n 0
   
 Hˆ 2   cm*  t  cn  t En2 m n   cm*  t  cn  t En2 mn
m0 n 0 m0 n 0


 Hˆ 2   cn En2
2

n 0

2

 
  cn En    cn En 
2
 Hˆ  Hˆ 2  Hˆ
2 2 2

n 0  n 0 
 Usually this is not equal to zero; each energy measurement does not always
give the same value.
26
Time–independent Schrödinger equation – Stationary states – Linear combinations
Example: consider a simple 2-state system with normalized state kets |1 and |2 and
corresponding energy eigenvalues 𝐸1 and 𝐸2 , respectively.
Let’s assume the starting wave function is an equal weight linear combination:
1
   1  2 
2
Time dependence:
1   iE1t   iE2t  
 t    exp    1  exp  2
2     
E1  E2
 Hˆ 
 2
E12  E22
 Hˆ 2

 2

E12  E22  E1  E2  E12  2 E1 E2  E22 E1  E2 E


2

 Hˆ      
2  2  4 2 2

27
Time–independent Schrödinger equation – Stationary states – Linear combinations
Let’s summarize these important results for linear combinations:
If |Ψ 𝑡 = ∞ 𝑛=0 𝑐𝑛 𝑡 |𝑛 , with
• |𝑛 the solutions (i.e., the stationary states) of the time-independent Schrodinger equation,
• all |𝑛 and also |Ψ 𝑡 normalized (i.e. 𝑛|𝑛 = 1 and Ψ 𝑡 |Ψ 𝑡 = 1),
then

 

 cn  1 Hˆ   cn En cn (t )  n  (t )
2 2

n 0 n 1

So |𝑐𝑛 |2 tells us the probability that a measurement of the energy would yield the value 𝐸𝑛 .

 A competent energy measurement will always return one of the allowed energy values, and
|𝑐𝑛 |2 is the probability of getting the particular value 𝐸𝑛 .

 Notice that the probability of getting a particular energy 𝐸𝑛 is independent of time: |𝑐𝑛 |2 .

 This is a manifestation of conservation of energy in quantum mechanics.

28
Time–independent Schrödinger equation – Stationary states – Measurements
In the context of measurements of the total energy we postulate the following:
1. The outcome of any measurement of H is always one of its eigenvalues 𝐸𝑛 .
2. The eigenkets |0 , |1 , |2 ,… form a complete set of states, i.e. they form a basis set that
spans the state vector space.
3. If the state of a system is described by the normalized state vector |Ψ 𝑡 , and if the states
|0 , |1 , |2 ,… are orthonormal, then the probability of finding the system in state |𝑛 at time 𝑡 is
given by |𝑐𝑛 |2 = | 𝑛|Ψ 𝑡 |2 .

4. The state of a system, immediately following a measurement of H that gave value 𝐸𝑛 ,


collapses to state |𝑛 .

 If 𝐸𝑛 is degenerate, the state collapses to the subspace spanned by the degenerate


states corresponding to 𝐸𝑛 .

Note these postulates hold for any observable A, with as eigenvalue equation A|𝜙𝑛 = 𝑎𝑛 |𝜙𝑛 .

29
Time–independent Schrödinger equation – Bound states
Let’s now make this discussion a bit more practical and discuss the case of a single object,
confined by a given external potential to a finite volume.
We call such cases bound-state problems.
We will find that boundary condition can only be satisfied for certain discrete values of the
energy E.
The allowed energy values are called the energy eigenvalues of the potential,
The states are called energy eigenstates (synonymous with stationary states).

30
Time–independent Schrödinger equation – The infinite square well

Practical example: particle in a box.

Inside the box, between 2 grids, ground potential.

As long as negatively charged particle is in


between grids, it experiences no electrical forces.

If particle passes through grid 𝐴, it experiences an


electrical field that exerts a backward force on it,

As long as 𝐸 < 𝑉0 , the particle will remain bound.

Regions 𝐴 − 𝐵 can be considered as soft walls,


from which the particle bounces.

Walls can be made ‘rigid’  put higher and higher


potential difference between 𝐴 and 𝐵.

Limiting case: 𝑉0 → ∞ “infinite square well”


31
TF, page 112
Time–independent Schrödinger equation – The infinite square well
1-D well with infinitely high barriers

0 for 0 xL
V  x  
Suppose
 elsewhere
 Outside the well, 𝜓 𝑥 = 0
 Inside the well, 𝑉 𝑥 = 0
d 2 2
So Schrödinger equation becomes:  2
 E
2m dx
d 2 2mE k2 2
Let’s rewrite this: 2
  k 2
 with k E
dx 2m
 This is the classical simple harmonic oscillator equation
 Inside the well, essentially this is a free-particle problem (no potential)
 Solution is a superposition of plane waves (one moving right, one moving left)

  x   A exp  ikx   B exp  ikx 

 These plane waves carry the same kinetic energy, but opposite momentum.
32
HK, page 54
Time–independent Schrödinger equation – The infinite square well
Let’s quickly verify the statement that the solution is a superposition of plane waves (one
moving right, one moving left):

We have   x   A exp  ikx   B exp  ikx 

We can ‘tack’ onto these solutions the standard time dependencies, exp −𝑖𝐸 𝑡 ℏ :

  E    E  ℏ2 𝑘 2
   x, t   A exp ik  x  t    B exp  ik  x  t   with 𝐸 = 2𝑚
  k    k 

Any function of 𝑥 and 𝑡 that depends on these variables in the special combination 𝑥 ± 𝑣𝑡 (for
some constant 𝑣), represents a wave of fixed profile, travelling in the ∓𝑥 direction, at speed 𝑣.

How do we see this?


A fixed point on the waveform (e.g. its maximum or minimum) corresponds to a fixed value of
the argument, and hence to 𝑥 and 𝑡 such that

𝑥 ± 𝑣𝑡 = constant, or 𝑥 = ∓𝑣𝑡 + constant

Since every point on the waveform is moving with the same velocity, its shape does not
change as it propagates.
33
HK, page 54
Time–independent Schrödinger equation – The infinite square well
In 𝛹 𝑥, 𝑡 , when considering the first term, when 𝑡 increases, 𝑥 has to increase proportionally
to keep 𝑥 − 𝑣𝑡 constant: This term represents a wave travelling to the right.
In 𝛹 𝑥, 𝑡 , when considering the second term, when 𝑡 increases, 𝑥 has to proportionally
decrease to keep 𝑥 + 𝑣𝑡 constant: This term represents a wave travelling to the left.

  E    E 
   x, t   A exp ik  x  t    B exp  ik  x  t  
  k    k 

Wave moving to the right Wave moving to the left

We get the wave velocity from the fact that the argument remains constant.
Therefore, it’s time-derivate must be zero.

Let’s e.g. look at the first term on the RHS

d E 
  x  t  0
dt  k 
dx E k
v  
dt k 2m 34
Time–independent Schrödinger equation – The infinite square well
Boundary conditions: 𝜓 𝑥 must remain continuous at 𝑥 = 0 and 𝑥 = 𝐿

x0   0  A  B  0  B  A

   x   A exp  ikx   exp  ikx  

 2iA sin  kx   C sin  kx  with C  2iA

 The two plane waves superpose into a standing wave!

xL   L   C sin kL
 𝐶 = 0 (trivial solution) or sin𝑘𝐿 = 0.

 kL  0,   , 2 , 3 ...


n
 Distinct solutions are kn  with n  1, 2,3,...
L
kn2  2 2 2
2  Energy is quantized!! Only these values are allowed!
En   2
n
2m 2mL  We call 𝑛 the quantum number of this system.
For bound-state problems, we have 1 quantum number per degree of freedom
35
Here, the only degree of freedom is the variable 𝑥.
Time–independent Schrödinger equation – The infinite square well
2 2
Lowest energy state: E1 
2mL2
Since the particle cannot have energy less than 𝐸1 , the energy of the particle in the box, even it
is temperature is reduced to 0 K, will still be 𝐸1 and will not be zero.
Hence we call 𝐸1 the zero-point energy.

 𝐸1 can be seen as the natural unit of energy for this system

 Lowest energy state has an energy above the minimum of the potential energy!

 Occurs in all quantum-mechanical bound-state problems

 Without it electrons in an atom would fall into the nuclei…

Note that when we superimpose the two plane waves to meet the boundary conditions, their
momentum contributions cancel (hence they are standing), but the energies not!

 We consider the zero-point energy to be internal kinetic energy, with no external


motion of the object through space.

This also implies that the particle is not at rest, even at 0 K

The result is, the position of the particle is not precisely known in the box.
36
Time–independent Schrödinger equation – The infinite square well
Let’s get some feeling for this:

Example 2:
• Typical atomic dimensions ~10-10 m • Macroscopic dimension ~1 cm
Let’s take L = 2x10-10 m and m = me  E1 = 3.8x10-15 eV
 E1 = 1.5x10-18 J = 9.4 eV  Very small, energy of macroscopic
 Large! Readily observable by experiments systems appears to be continuous!

ℏ = 1.05457x10-34 Js me = 9.10938x10-31 kg [Js] = [kg m2 /s]

37
Time–independent Schrödinger equation – The infinite square well
To find C, we normalize the wave function:
L

  kx  dx  1
2
2
C sin
0

 exp  ikx   exp  ikx  


L L 2

  C sin 2  kx  dx  C 0 
2 2
 dx
0
2i 
2 L
C
  exp  2ikx   exp  2ikx   2  dx
 2i 
2
0

C  exp  2ikx  exp  2ikx 


L

2

    2x
4  2ik 2ik 0
sin  2kx  
L
C 
2

 2 x  
4  k 0
2
C 
2

2  n 
 n  x  sin  x
L  L  38
DJG, page 32
Time–independent Schrödinger equation – The infinite square well

2  n 
 n  x  sin  x
L  L 

Note that negative quantum numbers 𝑛


lead to identical results…

 Physically they represent the same state


 Hence, we only consider positive 𝑛

HK, page 55 39
SB, page 113
Time–independent Schrödinger equation – The infinite square well
Interpretation of quantum amplitudes for bound states
𝒏=𝟐
2  n 
 n  x  sin  x
L  L 
2 2  n 
  n  x
2
 sin  x
L  L  𝒏 =18

In Newtonian mechanics, a particle of given energy,


confined between rigid walls but otherwise free,
would be equally likely to be found between any
point between 𝑥 = 0 and 𝑥 = 𝐿.

In quantum mechanics, there is a certain probability


to find particle, that is position dependent.

Here, for very large 𝑛, the wavelength in the well becomes very small.

 The result of averaging the probability over some Δ𝑥 will give a result
almost independently of 𝑥.
Niels Bohr
 Example of the so-called correspondence principle of Bohr: (1885-1962)
1922

Classical and quantum predictions agree in the limit of large quantum numbers 40
Time–independent Schrödinger equation – The infinite square well

2  n 
 n  x  sin  x
L  L 

n 1 ground state

n  2,3,... excited states

Properties: 1. States are alternatively even and odd, with respect to center of the well:
𝜓1, 𝜓3, 𝜓5,… are even; 𝜓2, 𝜓4, 𝜓6,… are odd
2. With increasing energy, number of nodes goes up by 1.

3. All states are mutually orthogonal:


 m  x  n  x  dx  0
 *

when 𝑚 ≠ 𝑛

41
HK, page 55
Time–independent Schrödinger equation – The infinite square well
 m   n 
L
2
   n   L 0  L
  
*
Explicit orthogonality proof: x
m x dx sin x  sin  x  dx
  L 
exp  iax   exp  iax  exp  ibx   exp  ibx 
sin  ax  sin  bx  
2i 2i
exp i  a  b  x   exp i  a  b  x   exp  i  a  b  x   exp  i  a  b  x 

4

cos  a  b  x   cos  a  b  x 



2

1  mn   n  m 
L
   x  n  x  dx   cos 
*
m  x   cos   x   dx
L0  L   L 
L
L  1 mn  1  m  n  
  sin   x  sin   x 
L   m  n    L   m  n    L   0

1  sin  m  n    sin  m  n    


   0 when 𝑚 ≠ 𝑛
  mn mn 
42
Time–independent Schrödinger equation – The infinite square well

Let’s summarize the general 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, i.e. any other function 𝑓(𝑥) can be written as a
linear combination of them.  Valid for every potential

Specific case of infinite square well:



2   n 
f  x    cn n  x    n  L
c sin x
n 1 L n 1  Joseph Fourier
(1768-1830)

 This is actually the Fourier series for 𝑓(𝑥) !


43
Time–independent Schrödinger equation – The infinite square well

2   n 
f  x    cn n  x    n  L
c sin x
n 1 L n 1 
How to calculate cn?  ‘Fourier’s trick’
 Multiply both sides by 𝜓𝑛∗ , followed by integration over all space
 

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


* *
k n k n n nk
n 1 n 1

ck   k*  x  f  x dx


Recall in Dirac notation: f   cn n  ck  k f
n 1


From this, we also found f  n n f
n 1

44
Time–independent Schrödinger equation – The infinite square well
Putting things together for the infinite square well
2  n   n 2 2 
Stationary states:  n  x, t   sin  x  exp  i t
 L 
2
L  2mL 


2  n   n 2 2 
 General solution:   x, t    cn sin  x  exp  i t
 L 
2
n 1 L  2 mL 
Any prescribed initial wave function 𝛹 𝑥, 0 can be fitted by appropriate choice of the
coefficients cn:

2   n 
  x, 0    cn n  x    n  L
c sin x
n 1 L n 1 

 n 
L
2
cn   n*  x    x, 0 dx  x    x, 0  dx
L 0
with sin 
 L 
Knowing the wave function, Ψ 𝑥, 𝑡 , the expectation values of all dynamical variables can be
calculated.

 This procedure is valid for any potential; the only thing that changes are
the function form of 𝜓𝑛’s and the equation for the allowed energies.
45
Time–independent Schrödinger equation – The infinite square well

Example 3:

A particle in the infinite square well has the


initial wave function

  x, 0   Ax  a  x  0  x  a for some 𝐴.

Outside the well, of course, Ψ = 0. Find Ψ 𝑥, 𝑡 .

Solution
First we need to determine 𝐴, by normalizing Ψ 𝑥, 0 :

a a
a5
1     x, 0  dx  A  x  a  x  dx  A
2 2 2 2 2

0 0
30

30
 A
a5

46
Note we used here 𝑎 instead of 𝐿 for the width of the well
Time–independent Schrödinger equation – The infinite square well
 n 
a
2
x    x, 0  dx
a 0
The 𝑛th coefficient is cn  sin 
 a 
 n  30
a
2
x  5 x  a  x  dx
a 0
 cn  sin 
 a  a

2 15   n   n  2 
a a
 3  a  sin  x  xdx   sin  x  x dx 
a  0  a  0  a  

Let’s solve these integral separately, each by integration by parts:

1 d  1 dx 
a a a

0        
c 0  dx 0
a (partial integration)
sin cx xdx   cos cx  xdx    cos cx x 0
 cos cx dx 
c dx 
a
1 1 
  cos  cx  x  sin  cx  
c c 0
1 1 
  cos  ca  a  sin  ca  
c c 

 n    n   n 
a
a2 a a3
 a  sin  x  xdx   cos  a aa  sin  a   cos  n 
0  a  n    n  a  n 47
Time–independent Schrödinger equation – The infinite square well
1 d  2
a a
also: 0 sin  cx  x 2
dx     cos  cx   x dx
c 0  dx 
1 
a
  cos  cx  x  2  cos  cx  xdx 
2 a

c 0
0 
1 2 d  
a
  cos  ca  a    sin  cx   xdx 
2

c c 0  dx  
1  
a
2
  cos  ca  a  sin  cx  x 0   sin  cx  dx  
2 a

c  c 0  
1  2 1  
a

  cos  ca  a  sin  cx  x  cos  cx   


2

c  c c  0 
a
 n  2 a  2a   n  a  n  

a

  sin  x  x dx   cos  n  a 2  sin  a x x  cos  x  


0  a  n  n    n  a   0 

 
3
 a  
     cos  n   2 
2
  2 n

 n 
48
Time–independent Schrödinger equation – The infinite square well
Let’s put things together:
2 15   n   n  2 
a a
 cn  3  a  sin  x  xdx   sin  x  x dx 
a  0  a  0  a  
2 15  a 3 
 
3
 a   
 3  cos  n      2   n  cos  n   2  
2

a  n  n  


4 15
1  cos  n  
 n 
3

 0 if 𝑛 is even,

  8 15 if 𝑛 is odd.
 n 3
 

 n  n 2 2 
3
2 30  1 
   x, t    
 

a n 1,3,5,... n3
sin 
 a
x  exp  i

2
t
 2 ma 

Loosely speaking, 𝑐𝑛 tells you the ‘amount of 𝜓𝑛 that is contained in Ψ 𝑥, 𝑡 .’ 49


Time–independent Schrödinger equation – The infinite square well

Example 4:
We note that in example 3, the starting wave function, Ψ 𝑥, 0 , actually closely resembles
the ground state 𝜓1.

Recall:

2  n 
 n  x  sin  x
L  L 

and

  x, 0  
30
5 
x a  x 0  x  a
a

 This suggests that |𝑐1 |2 should dominate in the


general solution, 𝛹 𝑥, 𝑡
50
Time–independent Schrödinger equation – The infinite square well

As general solution, we had :


 0 if 𝑛 is even,

 n   n 2 2  
  x, t    cn
2
sin  x  exp  i t cn   8 15
 L 
2
 n 3 if 𝑛 is odd.
 
n 1 L  2 mL 

In fact, we find indeed that


2
 8 15 
  3   0.998555...
2
c1
   2
  8 15  
1
 cn   3   1
2
The rest of the coefficients make up the difference:
  
6
n 1 n 1,3,5 ,... n

Next, the expectation value of the energy is


H   En cn
2

n 1

51
Time–independent Schrödinger equation – The infinite square well
kn2  2 2 2
2

The energy eigenvalues are En   2


n
2m 2mL

 0 if 𝑛 is even,

Recall, 𝛹 𝑥, 0 imposes that cn   8 15
 n 3 if 𝑛 is odd.
 
2
  8 15  
1  2 2n2
 H   En cn   3   6 2ma 2
2

n 1    n 1,3,5,... n

2
1  8 15  2 
1 2
  2 
2    ma 2

n 1,3,5,... n
4
5 2
ma
(such series can be looked up in
math tables)

𝜋 2 ℏ2
 As one might expect, it is very close to 𝐸1 = ; slightly larger, because of the admixture
2𝑚𝑎2
of excited states.

52
Time–independent Schrödinger equation – The infinite square well
Free-electron gas – Density Of States
With this simple, idealized case, we can already discuss also some important properties in solid
state materials such as the density of states function.
In the solid state, a few of the loosely bound outermost valence electrons in each atom become
detached.
They roam around throughout the material, no longer subject only to the Coulomb field of a
specific “parent” nucleus, but rather to the combined potential of the entire lattice.

The figure shows a qualitative indication of how an


approximation to a potential well results from
superimposing the potentials acting on a conduction
electron in a metal.

For a single ion, the electron-nucleus interaction is


purely Coulombic.

The potentials are due to the closely spaced positive


ions in the metal.

53
Time–independent Schrödinger equation – The infinite square well
We will examine two extremely primitive models in 3-D:
1. The electron gas theory of Sommerfeld
 Here all forces are ignored (except the confining boundaries).
 The wandering electrons are treated as free, non-interacting particles in a box Arnold Sommerfeld
(1868-1951)
 This is the 3-D analog to an infinite square well.
 We will look at the states of this system in the next few pages.

2. Bloch’s theory
 Here a periodic potential is introduced, representing the electron attraction
of the regularly spaced, positively charged, nuclei (but still ignoring the
electron-electron repulsion).

A typical idealization of the actual crystal Felix Bloch


potential (a) is the Kronig-Penney model (b). (1905-1983)
1952

 We will treat this in a chapter on carrier scattering, if time allows. 54


Time–independent Schrödinger equation – The infinite square well
Generally, in solid-state problems our goal is often to determine the number of allowed states
per unit of crystal volume as a function of energy.

We call this the density of states (DOS) function, 𝑔 𝐸 :

𝑔 𝐸 = 𝑑𝑁 𝐸 /𝑉𝑑𝐸 In this, 𝑉 is the volume of the crystal.

This is the number of states per unit energy interval, in the vicinity of 𝐸, and per unit volume

To get a first feeling for this, let’s briefly discuss the electron gas theory of Sommerfeld in 1-D.

55
Time–independent Schrödinger equation – The infinite square well
Electron gas theory of Sommerfeld in 1-D.
In this problem, we still work in the 1-D infinite square-potential well.
However, we assume the presence of a number of non-interacting electrons (say 𝑛𝑒 ).
The 1-D DOS function, 𝑔1𝐷 𝐸 is defined as:

𝑔1𝐷 𝐸 = 𝑑𝑁 𝐸 /𝑎𝑑𝐸 In this, 𝑎 is the length of the 1-D crystal.

This is the number of states per unit energy interval, in the vicinity of 𝐸, and per unit crystal
length

Before calculating 𝑔1𝐷 𝐸 , let’s analyze this problem qualitatively:


1. As the 𝑛𝑒 electrons do not interact (e.g. no Coulomb interaction), the potential profile 𝑉 in
the Schrodinger equation remains unchanged.
 Therefore, all electrons experience obey the same, identical 1-electron Schrodinger
equation!

2. If electrons were bosons, in the ground state, they all would occupy the lowest energy level 𝐸1 .
However, electrons are so-called fermions that obey the Pauli exclusion principle.
56
Time–independent Schrödinger equation – The infinite square well
Indeed, one particularly important property of fermions
is that they obey the Pauli exclusion principle:

Only two electrons can occupy any given state Enrico Fermi Wolfgang Pauli
(1901-1954) (1900-1958)
(that is, a given solution of the Schrodinger 1938 1945
equation): one with spin up, one with spin down.

Now, in the ground state of this 𝑛𝑒 -electron system, at a temperature of 0K, we have that:

 The electrons will fill all available energy states systematically, starting
with the lowest energy state (corresponding with 𝐸1 ).
 They will do so up to a given energy 𝐸 = 𝐸𝐹 , called the Fermi energy.

 𝐸𝐹 is the energy of the highest occupied energy level in the ground state of a multi-
electron system.
 In the ground state (at 0 𝐾), all electrons will have a energy lower (or equal) than 𝐸𝐹 .

Let’s now calculate 𝑔1𝐷 𝐸 , i.e. the number of states per unit energy interval, in the vicinity of
𝐸, and per unit crystal length

57
Time–independent Schrödinger equation – The infinite square well
ℏ2 𝑘 2
Recall that the energy of an electron within the quantum well is given by 𝐸 = (as the
2𝑚
solutions are plane waves).
𝑘
Let’s now try to find out the number of states 𝑁 that have an energy below
below a certain value 𝐸.
We can also express this as by finding the number of states with a 𝑘-value below
2𝑚𝐸
the 𝑘 corresponding to 𝐸, namely 𝑘 = . 𝜋
ℏ2
𝑎
Recall, each unique quantum state is characterized by its quantum number 𝑛.
In turn, each 𝑛 is uniquely associated with a 𝑘𝑛 -value: 𝑛 =1

𝑛𝜋
𝑘𝑛 = , with 𝑛 = 1,2,3, . . .
𝑎 𝑛=0

To assist in the counting process, we can plot these on an axis as shown in Fig.

Note: This axis has the dimension of 𝑥 −1 ; we call this reciprocal space (here in 1-D).

In the construct, the 𝑘𝑛 -values, each representing one Schrodinger equation solution, are
recorded as filled dots on a 1-D 𝑘-space plot.
Taking note of the lattice-like arrangement of the solution dots, one readily deduces that a 𝒌-
𝜋
space plot “unit cell” of volume 𝑎 contains one allowed solution. 58
Time–independent Schrödinger equation – The infinite square well
The total number of mathematical solutions, 𝑁𝑚𝑎𝑡ℎ , below such a constant-energy surface is:

𝑘 𝑘 𝑘𝑎
𝑁𝑚𝑎𝑡ℎ = =𝜋 =
(𝑣𝑜𝑙𝑢𝑚𝑒 𝑜𝑓 𝑢𝑛𝑖𝑡 𝑐𝑒𝑙𝑙) 𝜋
𝑎
Note: We excluded already the negative wave numbers from the solutions, as
these do not represent unique physical solutions.

For electrons, we still need to make an additional correction: namely, two allowed spin states
(spin up and spin down) must be associated with each independent solution.
2𝑘𝑎
𝑁 = 2𝑁𝑚𝑎𝑡ℎ =
𝜋

We are interested in the DOS as function of 𝐸.

So let’s write the number of states 𝑁 as function of 𝐸.

ℏ2 𝑘 2 2𝑚𝐸
For this, we make again use of 𝐸 = , (or 𝑘 2 = ):
2𝑚 ℏ2

2a
N 2mE
 59
Time–independent Schrödinger equation – The infinite square well
2a
N 2mE

dN a 2m
 
dE  E

1 dN
Finally, by definition g1D  E  
a dE

1 2m
hence, g1D  E  
 E

The DOS diverges as 𝐸 −1/2 , as 𝐸 → 0, which is a characteristic feature of 1 dimension.

60
Time–independent Schrödinger equation – The infinite square well in 3D
Let’s now go back to the case of the 1-electron infinite square well, but now in 3-D
Suppose the object in question is a rectangular solid, with dimensions 𝐿𝑥 , 𝐿𝑦 , and 𝐿𝑧 .
Imagine that an electron inside experiences no forces at all, except to the impenetrable walls:

0 if 0 < 𝑥 < 𝐿𝑥 , 0 < 𝑦 < 𝐿𝑦 , and 0 < 𝑧 < 𝐿𝑧


V  x, y , z   
 otherwise

 Let’s find first the stationary states of this problem!


2
The time-independent Schrodinger equation,   2  E , separates.
2m
2 f 2 f 2 f
In Cartesian coordinates:  f  2  2  2
2

x y z

Let’s propose the following solution:   x, y , z   X  x  Y  y  Z  z 

 d 2 X
2
  d 2Y   d 2Z  
   YZ   2  XZ   2  XY   EXYZ
2m  dx 2   dy   dz  
61
 Let’s divide this by 𝑋𝑌𝑍 on both sides.
Time–independent Schrödinger equation – The infinite square well in 3D

 d 2 X  1  d 2Y  1  d 2 Z  1 
2
  2    2    2    E
2m  dx  X  dy  Y  dz  Z 

 Each term only depends on one, different variable.


 Each term must be equal to a constant.

 Let’s introduce thus 𝐸𝑥 , 𝐸𝑦 , and 𝐸𝑧 , as separation variables

 2 d2X
 2m dx 2  Ex X

 2
d 2Y
We get  2
 E yY where Ex  E y  Ez  E
 2m dy
 2
d 2Z
  Ez Z
 2m dz 2

62
Time–independent Schrödinger equation – The infinite square well in 3D
d2X  2mEx  2 2
kx
  x  
x 0
2 k
 dx 2 k  x
E
 2   2m
d Y  

2 2
2mE y ky
  2  k y2  0 In this we defined k y   Ey 
 dy   2m
 d 2Z  2mEz  2 2
kz
 2  k z2  0  kz   Ez 
 dz   2m

2
The total particle energy is E  Ex  E y  Ez 
2m
 k 2
x  k y2  k z2 

2
k2
We can write this as E
2m

Here, 𝑘 = 𝑘𝑥2 + 𝑘𝑦2 + 𝑘𝑧2 is the magnitude of the wave vector, 𝐤 = (𝑘𝑥 , 𝑘𝑦 , 𝑘𝑧 ).

63
Time–independent Schrödinger equation – The infinite square well in 3D

Note that multiple vectors (𝑘𝑥 , 𝑘𝑦 , 𝑘𝑧 ) can have the same magnitude 𝑘 = 𝑘𝑥2 + 𝑘𝑦2 + 𝑘𝑧2 .
These vectors describe the surface of a sphere in 𝑘-space (figure).
ℏ2 𝑘 2
All these vectors represent states with the same energy 𝐸 = .
2𝑚

 The surface of this sphere is a constant-energy surface.

We call all states ending in the same constant-energy


surface degenerate states.

Note 1: This graph illustrates well how degeneracy only exists for
problems with more than 1 degree of freedom!

Note 2: Here, as we did not yet consider the boundary conditions of the
potential well, all values of (𝑘𝑥 , 𝑘𝑦 , 𝑘𝑧 ) possible.

64
Time–independent Schrödinger equation – The infinite square well in 3D
Solving now the differential equations, we obtain the following general solutions *:

 X  x   Ax sin  k x x   Bx cos  k x x 

Y  y   Ay sin  k y y   By cos  k y y 

 Z  z   Az sin  k z z   Bz cos  k z z 
The first boundary conditions requires that X (0)  Y  0   Z  0   0
 X  x   Ax sin  k x x 

 Bx  By  Bz  0  Y  y   Ay sin  k y y 

 Z  z   Az sin  k z z 

The second boundary conditions require that X (a)  Y  b   Z  c   0

 k x a  n x  nx  0, 1, 2, 3,...


 Mathematically, we find that each 
  k y b  n y n y  0, 1, 2, 3,...
k c  n  𝑛 must be an integer:  n  0, 1, 2, 3,...
 z z  z

* we directly wrote the solutions as sines and cosines, for convenience. We could of course also have written
65
the solutions as a superposition of two plane waves (as we did for the 1-D problem)
Time–independent Schrödinger equation – The infinite square well in 3D

 nx  1, 2,3,...
Physically, we only retain the 
n y  1, 2,3,...
positive, non-zero values:  n  1, 2,3,...
 z

 2 2
kx  nx
2 2 2

 x
E   2
 2 m 2 ma
 2 2
ky  ny
2 2 2
From this, we find that the energy is quantized in all directions:  E y   2
 2 m 2 mb
 2 2
kz  nz
2 2 2

 Ez   2
2 2  2 m 2 mc
k
As before, the total energy is E  Ex  E y  Ez 
2m

Again, 𝑘 = 𝑘𝑥2 + 𝑘𝑦2 + 𝑘𝑧2 is the magnitude of the wave vector, 𝐤 = (𝑘𝑥 , 𝑘𝑦 , 𝑘𝑧 ).

66
Time–independent Schrödinger equation – The infinite square well in 3D

However, due to the boundary solutions, only given


quantized solutions are possible (represented by dots
in a reciprocal space lattice).

Each rectangular unit cell features 8 dots, but each dot


is shared by 8 units.

Therefore, each mathematical distinct solution takes a


volume


a b c
 See also argument DOS calculation 1-D well

 3
Now, 
a b c V

In this, 𝑉 is the physical volume of the quantum well


𝜋3
 Each mathematically distinct solution takes a volume in 𝑘-space
𝑉
67
Time–independent Schrödinger equation – The infinite square well in 3D
For completeness, finally, we get the values for 𝐴𝑥 , 𝐴𝑦 , and 𝐴𝑧 from normalization:
a b b
1   X dx   Y dy   Z dz
2 2 2

0 0 0

The normalized stationary states are then:

8  n x   n y   n y 
 n ,n y , nz
 sin  x  sin  y  sin  y
x
abc  a   b   c 
 Each solution is now associated with three quantum numbers, one for each degree of
freedom (here the three dimensions, 𝑥, 𝑦, 𝑧).

(A fourth quantum number is associated with electron spin.)

2
 2  nx2 nz2 n y2
2 2
k
The allowed energies are Enx ,ny ,nz      
2m  a 2 b 2 c 2  2m

𝑘 is the magnitude of the wave vector, 𝐤 = (𝑘𝑥 , 𝑘𝑦 , 𝑘𝑧 ).


Recall, all 𝐤-vectors with the same length, and thus energy (i.e. ending on the same
spherical surface) are called these degenerate states. 68
Time–independent Schrödinger equation – The infinite square well in 3D: Degenerate states
Degenerate states
The higher the degree of symmetry of the potential function 𝑉, the higher the amount of
degenerate states.
Indeed, for a cubic well (𝑎 = 𝑏 = 𝑐 = 𝐿), we obtain

2
2
Enx ,ny ,nz 
2mL2
n 2
x  n y2  nz2   E1n 2

 most of the energy eigenvalues of this cubic well are now degenerate:

 If the three quantum numbers are different from each other, then each of the 3! = 6
possible permutations of the three numbers (𝑛𝑥 , 𝑛𝑦 , 𝑛𝑧 ) corresponds to a different state,
even though all have the same energy!

Example 5:  n , n , n   1, 2,3 ; 1,3, 2  ;  2,1,3 ;  2,3,1 ; 3,1, 2  ; 3, 2,1.


x y z

All have the same energy eigenvalue: E  1  2  3  E  14 E


nx , n y , nz
2 2 2
1 1

 This energy level is six-fold degenerate.

Note: as stated, degeneracy only exists for problems with more than 1 degree of freedom
69
Time–independent Schrödinger equation – The infinite square well in 3D: Degenerate states
In addition to these symmetry-induced degeneracies, there are numerous of additional
accidental degeneracies (simply as consequence of the 3-D nature of potential)
𝑬 𝑬𝟏
Example 6: E  27 E1
 Can be obtained in 2 different ways:
12  12  52  27
 Three different states have these quantum numbers
32  32  32  27
 One state has these quantum numbers

 Four different states lead to this energy level


 3 with quantum numbers (1,1,5),
(1,5,1) and (5,1,1)
 1 with (3,3,3)

The figure shows the degeneracy for the different energy levels in a cubic well.
 The number of additional degeneracies increases drastically with increasing energy.
 The degree of degeneracy does not follow a simple pattern
70
LP, page 99
Time–independent Schrödinger equation – The infinite square well in 3D: DOS
Density of states
To calculate this, let’s go back to the rectangular quantum well case.

As we argued, the solutions can be represented by


(𝑘𝑥 , 𝑘𝑦 , 𝑘𝑧 ), which forms a Cartesian space (see figure).

Recall, for the 3-D well, each distinct mathematical


𝜋3
solution (dots in figure) takes a volume in 𝑘-space.
𝑉

2
k2
We also had E
2m

We now want to determine the number of states 𝑁 with an energy below a certain energy value
𝐸.

In (𝑘𝑥 , 𝑘𝑦 , 𝑘𝑧 ) space, this corresponds the number of states inside a sphere of radius 𝑘 (see
figure).

Remember that 𝑘 = 𝑘𝑥2 + 𝑘𝑦2 + 𝑘𝑧2


71
HK, page 59
Time–independent Schrödinger equation – The infinite square well in 3D: DOS

k3  This is for the case if both positive and negative


Volume sphere: 4 values of the components of 𝑛𝑖 would be counted
3

However, the negative components of 𝑛𝑖 (and thus 𝑘𝑖 ) do not produce additional states that are
physically distinct

 only the positive octant of a sphere should be considered!

1 4 3
Volume octant sphere: k
8 3
3
Each state takes a volume
V
1 4 3
k
8 3 Vk 3
 The number of states in an octant of a sphere: N math   2
 3
6
V
For electrons (fermions), we still need to make an additional correction: namely, two allowed
spin states (spin up and spin down) must be associated with each independent solution:
Vk 3
N  2 N math  2
3 72
Time–independent Schrödinger equation – The infinite square well in 3D: DOS
2
k2 2mE
Let’s write this as a function of 𝐸, using E k 
2m

2Vm 2m 3 2
N E
3 2 3

dN 3 2Vm 2m 1/2 Vm 2mE


  E 
dE 2 3 2 3 2 3

With this, we find readily the density of states function 𝑔(𝐸), now in 3-D.

As defined earlier, this is the number of states per unit energy interval, in the vicinity of 𝐸, and
per unit crystal volume:

1 dN m 2mE
g E  
V dE 2 3
The square root is characteristic for three dimensions.
Note that 𝑔 𝐸 is a density in a two-fold sense: Both per unit energy and per unit volume!
Also note that the final result is volume-independent. 73
Time–independent Schrödinger equation – The infinite square well in 3D: DOS
Summarized

The singularity at the bottom of the energy scale is much weaker for the 3-D case, compared to
the 1-D case.
In general, the DOS shows a stronger feature at the origin of the energy scale in fewer
dimensions.

Generally, the DOS plays an important role in the quantum mechanics of gases and free
electrons in metals and semiconductors.
Let’s briefly discuss this, introducing also a few useful concepts from solid state physics.

74
Time–independent Schrödinger equation – The infinite square well in 3D: DOS
So far, did not say much about how these states are filled when electrons are present.
Suppose now that our sample contains 𝑀 atoms, and each atom contributes 𝑑 electrons
Note: In practice, 𝑀 will be enormous – on the order of Avogadro’s number, for
an object of macroscopic size, whereas 𝑑 is a small number: 1 or 2, typically.

As argued, at 0K, these 𝑀𝑑 electrons will start filling up the available states, starting with those
with the lowest energy.

They will fill up one octant of a sphere in 𝑘-space, called the Fermi sphere.
The radius 𝑘𝐹 of this sphere is determined by the
fact that each pair of electrons requires a volume
𝜋3
in 𝑘-space.
𝑉

𝑘𝐹 is called the Fermi wave vector.


The spherical surface it describes is the Fermi
surface.
It describes the boundary separating occupied and
unoccupied states at 𝑇 = 0 K, in 𝑘-space.
The states that are filled at T = 0 K are sometimes
called the Fermi sea.
75
Time–independent Schrödinger equation – The infinite square well in 3D: DOS
Let’s calculate 𝑘𝐹 .
𝑉 3
Using the same derivation as before, we get 𝑀𝑑 = 𝑘
3𝜋 2 𝐹
13
 Md 2 
  3 2 
13
Rewritten, kF   3  
 V 
𝑀𝑑
Here, 𝜌 = is the free electron density (the number of free electrons per unit volume).
𝑉

From this we get the Fermi energy (also called the Fermi level), 𝐸𝐹 , at 0 K, for a free electron
gas
 3 
2 23 2
2 2
k
EF   F

2m 2m
As stated, this is the highest energy level that is occupied at T = 0 K

Along with 𝑘𝐹 , we can also define a Fermi momentum, 𝑝𝐹 = ℏ𝑘𝐹 , and a Fermi velocity 𝑣𝐹 =
ℏ𝑘𝐹 /𝑚.

Frequently, one also defines a Fermi temperature, 𝑇𝐹 = 𝐸𝐹 /𝑘𝐵 , where 𝑘𝐵 is Boltzmann’s


constant.
76
Time–independent Schrödinger equation – The infinite square well in 3D: DOS
Coming back to the DOS equation, in the semiconductor case, the derived DOS functions
describe the DOS in allowed energy bands, such as the conduction band.
The DOS function needs to be slightly changed then:
• The electron mass must then be replaced by an effective electron mass (accounting for the
periodic potential profile induced by the lattice atoms).

• The lowest energy should be set at the bottom of the energy band.

Just as an illustration, the DOS for the conduction band electrons in a semiconductor is given by

𝑚𝑛∗ 2𝑚𝑛∗ (𝐸 − 𝐸𝑐 )
𝑔𝑐 (𝐸) = for 𝐸 ≥ 𝐸𝑐
𝜋 2 ℏ3

𝑚𝑛∗ is the electron (𝑒) effective mass.


𝐸𝑐 is the bottom of the conduction band

Optical properties such as absorption are strongly influenced by the DOS, and low-dimensional
systems are preferred for optoelectronic devices because their DOS is higher at the bottom of
the band.

77
Time–independent Schrödinger equation – Quantum dot, quantum wire, quantum well

Quantum dot
Quantum dots are structures which confine electrons in all three dimensions to regions typically
of order 10-1000 nm.
The 3-D Schrodinger equation and its solutions represent also the stationary states of the
quantum dot:
 The wave function represents an object (electron) enclosed in a small volume

For a cubic quantum dot, we simply have

8  n x   n y    nz  
 n ,n y , nz
 sin  x  sin  y  sin  z
x
L3  L   L   L 

2
2
Enx ,ny ,nz 
2mL2
n 2
x  n y2  nz2   E1n 2

 The problem is trickier when solving for a spherical quantum dot, where we need to
use the Laplacian in spherical coordinates

 Note how the energy spacings are increasing for smaller dots!
78
Time–independent Schrödinger equation – Quantum dot, quantum wire, quantum well

Quantum wire
In case the cubic quantum dot would not be confined in one direction (say in 𝑧-direction),
we still can use separation of spatial variables:
  x, y , z   X  x  Y  y  Z  z  and E  Ex  E y  Ez

n   n y 
with X
2
sin  x x and Y
2
sin  y but Z  A exp  ik z z 
L  L  L  L 
2mEz
with kz 
 The particle is confined in the 𝑥𝑦-plane, here it’s energy is quantized
 However, it can freely propagate in the 𝑧-direction (we don’t have a quantum number 𝑛𝑧 ).
 Here we show the case of a particle moving up along the 𝑧-direction.
Note that we can only normalize per unit length of quantum wire, but we cannot normalize
over all space!
 we come back to this issue when discussing the free particle
When we normalize over 1 particle per meter in 𝑧-direction, we have
2 n   n y 
  x, y, z   sin  x x  sin  y  exp  ik z z 
L  L   L  79
Time–independent Schrödinger equation – Quantum dot, quantum wire, quantum well

Quantum well
In case there is no confinement in two directions (say in 𝑥- and 𝑦- direction), we still use
separation of spatial variables:

  x, y , z   X  x  Y  y  Z  z  E  Ex  E y  Ez
2 n  2
2
with Z sin  z x Enz  nz2  E1nz2
L  L  2mL2

X  A exp  ik x x 
2
2mEx k x2
but kx   Ex 
2m

Y  B exp  ik y y 
2
and 2mE y k y2
ky   Ey 
2m
 The particle can freely propagate in the 𝑥𝑦-plane!
 However, it is confined in the 𝑧-direction, where its energy is quantized
2
k2 k   k1 , k2 
The total energy is En  Enz  where
2m
80
Time–independent Schrödinger equation – Quantum dot, quantum wire, quantum well

Graphically

Here, the quantum number 𝑛 (or 𝑛𝑧 ) is called the sub band index.
 For 𝑛𝑧 = 1,2,3, … the carrier is in the 𝑛𝑡ℎ sub band, for which the minimum energy, or “band
edge”, is 𝐸𝑛𝑧 .
81

You might also like