You are on page 1of 11

Quantum Mechanics SPH 313

2.1 The Four Postulates of Quantum Mechanics.


Before we learn about the applications of the time independent Schrodinger equation, let us look at
the postulates of quantum mechanics. A postulate is a hypothesis, if it agrees with experiment, then it
can be taken as an axiom, a truth which we cannot prove. Quantum Mechanics is based upon a number
of postulates; these are stated quite differently depending on the source, vary in actual number and
order etc. depending on what book one reads. In these notes, I have used a four‐postulate form.

Postulate 1 States and wavefunctions

The first postulate concerns the information we can know about a state: postulate 1: The state of a
system is fully described by a function denoted Ψ (r1, r2 , r3 …..t).In this statement, r1, r2,….. are the
spatial coordinates of particles 1, 2, ….that constitute the system and t is the time. The function Ψ
plays a central role in quantum mechanics, and is called the wavefunction of the system. When we
are not interested in how the way the system changes in time we shall denote the wavefunction Ψ (r1,
r2 ,…). The state of the system may also depend on some internal variable of the particles(their spin
states); we ignore that for now and return to it in advanced quantum mechanics in year 4 of your
studies. The wavefunction contains information about all the properties of the system that are open
to experimental determination. Every physically-realizable state of the system can be described in
quantum mechanics by a wave function (state function), Ψ(r, t) in 3-D or Ψ(x, t) in 1-D), which
contains all accessible physical information about the system, including its spatial and temporal
evolution. Ψ (r, t) will tell us everything!
Physically realizable states means states that can be studied in laboratory. Accessible information
means the information we can extract from the wave function.

Postulate 2 The interpretation of the wavefunction and outcome of measurements

According to Born interpretation; the probability of finding a particle in a volume element dV, at r, at
any one time,
PdV, is given by:
Pdx   x, t  dx    x, t  x, t  dx
2
in 1-D
PdV   r , t  dV    r , t  r , t  dV
2
in 3-D

for finding our particle in the interval dx at x and t, in 1D. This means that  r , t  r , t  or

 r , t  dV is a probability density function for our system or (  x, t  x, t  =  x, t  dx in


2 2

1D). The wavefunction itself is a probability amplitude,and has no direct physical meaning. Note that
the probability density is real andnon-negative, the wavefunction (amplitude) may be complex and
negative.. Since the particle must exist somewhere in the universe, we obtain the normalization
condition, namely:


 x, t  x, t dx  1



in 1-D


1
Maxwell Mageto, PhD
Quantum Mechanics SPH 313

The interpretation of the wavefunction itself, and is commonly called the Born interpretation. The
implication of the Born interpretation is that the wavefunction should be square integrable, that is
2
 dV  

because there must be a finite probability of finding the particle somewhere (and that probability is 1
for a normalized wavefunction). Limitations on the wave function: (1) Only normalizable functions
can represent a quantum state and these are called physically admissible functions (2) state function
must be continuous and single valued function and (3) state function must be continuous and have a
continuous derivative.

Postulate 3 Eigenfunctions & Eigenvalues and expectation values

An observable is any dynamical variable that can be measured (Examples include, position,
momentum and kinetic energy). In classical mechanics, the observables are represented by
functions such as position as a function of time whereas in quantum mechanics observables are
represented by mathematical operators. An operator is a symbol for an instruction to carry out some
action, an operation, on a function. Every observable (variable), q, is represented by an operator, Q̂ ,
and this operator is used to obtain information about the observable. e.g .momentum, p , operator
  2 2
ˆp   i , position, x, operator, X̂ = x , kinetic energy operator Kˆ  , and so on.
x 2m x 2
Combined with the wave function, they can give rise to eigenvalue equations of the form: Qˆ   q
where q is the eigenvalue and  is the eigen function.

Eigenfunctions and Eigenvalues


Suppose that the effect of operating on some function  with the operator A is simply to multiply
 by a certain constant a n

A  an  , (2.0)*

we then say that  is an eigenfunction of A with eigenvalue a n . For example e 2 x is an eigenfunction


of the operator d / dx with eigenvalue 2.

dx
 
d 2x
e  2 e2x

The energy eigenvalue equation


The time-independent Schrodinger eigen value equation is:

Hˆ  n ( x)  E n n ( x)

2
Maxwell Mageto, PhD
Quantum Mechanics SPH 313

where  n (x) is the energy eigenfunction (or eigenstate) and En is the energy eigenvalue Let the
 2x 
wavefunction,  n ( x)  A sin  . For a free particle ( V ( x )  0 ), the time independent Schrodinger
 L 
equation.

 2 d 2 n ( x)
  E n n ( x) (i)
2m dx 2

 2 d 2 n ( x)
Where Ĥ  
2m dx 2
Find the energy eigenvalue, E n

d n ( x ) n  nx 
 . A cos 
dx L  L 
d 2 n ( x)  n   nx   n 
2 2

   A sin      n ( x) (ii)
 L   L   L 
2
dx

Combining equations (i) and (ii) and cancelling  n (x) from both sides of the equation gives energy
eigenvalue

 n 
2
2
     n ( x)  En n ( x)
2m  L 
n 2 2  2
En  (energy eigenvalue)
2mL2
The momentum eigenvalue equation
pˆ  ( x)  p ( x)

where p is the momentum eigenvalue. (Notice that here p is a continuous not a discrete variable).

If  ( x)  exp(ikx) , find the momentum eigenvalue.

pˆ  ( x)   i
d ikx
dx
 
e   i(ik )e ikx   i 2 k ( x)

Hence pˆ  ( x)  k ( x) .  (x ) is an eigenstate of the momentum operator with with momentum


eigenvalue k .

The parity eigenvalue equation


Parity: a wave function has a define parity if  ( x)    ( x) ; this requires a symmetric potential,
V ( x)  V ( x)

3
Maxwell Mageto, PhD
Quantum Mechanics SPH 313

Pˆ  ( x)  P ( x)

Where P is the parity eigenvalue. This eigenvalue is discrete and satisfies P  1 . To prove this, note
that the parity operator sets x to  x and so a bouble application of P̂ leaves the system invariant.
Thus Pˆ Pˆ  ( x)  P2 ( x)   ( x) . Hence  ( x)    ( x) , where  ( x)  P ( x)

The equation for the wavefunction


The wavefunction  r1 , r2 ....t  evolves in time according to the equation

  2  2 
H   i or   V ( x)  i
t 2m x 2
t

This partial equation is the celebrated Schrodinger equation, which was introduced by Erwin
schrodinger in 1926. At this stage, we are treating the equation as unsubstantiated postulate. Here the
operator H is called the Hamiltonian operator. H  Kineticenergy  potential energy  T  V

 2 d 2 
The time independent Schrodinger equation is given as   2
 V  ( x)  En  ( x)
 2m dx 

2 d 2
where Hˆ    V . Hence the Schrodinger equation may be written as an eigen value
2m dx2
equation H ( x)  En ( x)

The wave function or state function of an isolated system develops in time according to the Time
Dependent Schrödinger Equation (TDSE), namely:
2 2  r , t 
   r , t   V r,t  r , t   i in 3-D
2m t 2
2    x, t 
2
   x, t   V  x, t   x, t   i in 1-D
2m x 2
t 2
Which may also be written as

H  i 
t
Since the wave functions in postulate 1 have to obey the TDSE in postulate 4, there are important
 x, t 
conditions which apply to  x, t  . Amongst others, both  x, t  and its derivative, , must
x
be continuous, single valued and finite.

Example

4
Maxwell Mageto, PhD
Quantum Mechanics SPH 313

2  nx 
Given that the Schrodinger eigen value equation is H ( x)  En ( x) and   sin  for a
L  L 
particle in a box with V=0. Use the given equation to determine the eigen value E n .

Hˆ   En
 h 2  2  2  nx  n 2  2 2  2  nx 
 sin    sin 
2m x 2  L  L  2mL2  L  L 
n 2  2 2
Hence the Eigen value E n 
2mL2

Postulate 3 continued. Expectation values

In quantum theory, the ensemble average of an observable for a particular state of the system is called
the expectation value of that observable. We can obtain the average or expectation value for any
observable, Q

 q    x, t .Qˆ . x, t  dx

For example, the expectation value of momentum, p and expectation value for momentum squared,
p2
 
 
 p     x, t .  i .  x, t  dx   i   dx

x 
x
 
   2
2

 p    x, t .  i  . x, t  dx   2   2 dx
2 

  x  
x
The other statistical quantity that one uses in Quantum mechanics is the standard deviation of an
observable – otherwise known as uncertainty. For a position measurement, the uncertainty in x
(denoted x ) answers the following question: In an ensemble measurement at time t of
the position of a particle in a state  ( x, t ) , what is the spread of the individual results around the
expectation value  x  ?
The uncertainty, or standard deviation is given by x   x 2    x 2 or the uncertainty equals the
square-root of the dispersion.

Examples :

Given that
2  2x 
 x   sin  ; 0 xL Find the expactation value of momentum  p
L  L 

and the expectation of square momentum  p 2  using relevant equations given above

5
Maxwell Mageto, PhD
Quantum Mechanics SPH 313

 2x   2x 
L
2 d 2
 p  
0
sin
L  L 
.  i . sin
dx L  L 

 i 2  2x  d  2x   i 2  2   2x   2x 


L L
 0 sin L . dx sin L  dx  L . L 0 sin L  cos L   0
L    
2  2x 
cos  
L  L 

 2x   2x 
L 2
2 2 d 2
 p2  
0
L
sin
 L 
.  
dx 2
.
L
sin
 L 

 2x  d  2x  2  2  2  2x  2  2 


L 2 2L 2
2 2 2
   sin . 2 sin  dx      sin   dx    
2

0
L  L  dxL
 L   L  L

0  L   L 
 2 
2
2  2x  1
  sin  
 L  L  L 

Where we took advantage of the normalization integral


2  2x 
L L
2
0  dx  1  L 0 sin  L  dx

Expectation value of kinetic energy, K


 
2 2 2   
2
 K    x, t .   
2m  x 2


. x, t dx    dx

2m x 2
Expectation value for position, x and expectation value for square position, x 2
 
 x     x, t .x.  x, t  dx   x 

dx
 
 
 x     x, t .x .  x, t  dx 
2  2
x  
2 
dx
 
Or, indeed, the uncertainty, q
q   q 2    q 2
Uncertainty in momentum
p   p 2    p 2

Example: Find expectation value of position x . Given that


2  2x 
 x   sin  ; 0 xL
L  L 
 2x   2x  2  2x 
L L
2 2 2 L
 x  
sin .x. sin  dx   x.sin   dx 
0
L  L  L  L  L0  L  2
sin ax x cos ax
Where we used the integral  x sin ax dx  
a2 a

6
Maxwell Mageto, PhD
Quantum Mechanics SPH 313

Postulate 4 The fundamental prescription

The postulate concerns the selection of operators: Observables are represented by operators chosed
to satisfy the commutation realtions:
 
q, p q /  i qq /  
q, q /  0 
pq , pq/  0 
where q, and q / each denote one of the coordinates x, y, z and p q , and p q / the corresponding linear
momentum.

The above basic commutation relation is a basic, unprovable, and underivable postulate. It is the basis
of the selection of the form of the operators in the position and momentum representations for all
observables that depend on the position and the momentum.

2.2 OPERATORS IN QUANTUM MECHANICS


An operator is a symbol for an instruction to carry out some action, an operation, on a function. An
operator is a rule that transforms a given function into another function. An observable is any
dynamical variable that can be measured. In Classical mechanics, observables are represented by
functions (such as position as a function of time), in quantum mechanics they are represented by
mathematical operators. Observables are represented by operators chosen to satisfy the commutation
relations x, p  i . In the Schrödinger approach to quantum mechanics, classical quantities are
replaced by operators chosen to be consistent with the position-momentum commutation relation,
x, p  i . There is an additional property that quantum mechanical operators satisfy, namely that
they are hermitian. This requirement has several important, useful consequences described below. The
most important of these is that thereby the eigenvalues are always real numbers.

Observable Classical Quantum Mechanics


Mechanics Algebraic representation
position x̂ x
Potential Vˆ = mgh Vˆ ( x) (1D) or (V(x,y,z) or V(r) (3D))
energy
Momentum pˆ  mv 
 i (1D) or (  i (3D))
x
Kinetic energy 1 2 p2 2 2 2 2
ˆ
T  mv   (1D) or (   (3D))
2 2m 2m x 2 2m
Hamiltonian ˆ
H  T V 2 2 2 2
  V (1D) or (    V (3D))
2m x 2 2m
Parity Pˆ : x   x Not applicable

2.1.5 Commutation and non-commutation relations

An important feature of the operators is that in general the outcome of successive operations (A
followed by B, which is denoted BA, or B followed by A, denoted AB) depends on the order in which

7
Maxwell Mageto, PhD
Quantum Mechanics SPH 313

the operations are carried out. That is in general BA  AB , we say that, in general, operators do not
commute. The quantity AB  BA is called the commutator of A and B and is denoted A, B :
A, B  AB  BA
If AB  BA , then A, B  0 , and we say that A and B commute.
 d d d d
3, dx   3 dx  dx 3  0 hence 3 and dx commute
 
d  d d d
 dx , x   dx x  x dx  1 hence the operators dx and x do not commute.
 

d 
Example: Evaluate  , z 3 
 dz 
Solution Applying an arbitrary function of z ; g(z) to the commutator
d 3 d 3 3 d  d 3 3 d
 dz , z  g ( z )   dz z  z dz  g ( z )  dz ( z g )  z dz g
   
 3z 2 g  z 3 g /  z 3 g /
Where we used product rule of differentiation and denoted g /  dg / dz . Deleting the arbitrary
function g(z) from both sides leads to
d 3
 dz , z   3 z
2

 
 d
Exercise Now show that  z 3 ,    3 z 2
 dz 
Example 2 Evaluate the commutator x , px  in the position representation
Solution Applying an arbtrary function of x i.e. f(x) to the commutator
x , px   ( xpx  px x) f ( x)  x(i f )  i ( xf )
x x
  ixf   ixf  if 
/ /

where we have used the product rule of differentiation and denoted f /  df / dx . Simplifying the
expression above we get
x , px   i
The right hand side should be interpreted as the operator ‘multiplied by the constant i
Since x , p x   0 , we conclude that x and p x do not commute and hence their associated observables
can not be measured simultaneously. If operators commute, then they share common eigenfunctions
and they can be measured simultaneously. In other words, observables corresponding to commuting
operators can be measured simultaneously, while those corresponding to non-commuting operators
have an uncertainty principle. Remember that operators A and B commute if A, B  AB  BA  0
or if
AB  BA . Examples of pair of operators that commute include momentum and energy as well as
energy and parity. Examples of non-commuting operators include momentum and position as well
as energy and time.

8
Maxwell Mageto, PhD
Quantum Mechanics SPH 313

Exercise Now convince yourself that  px , x   i

2.1.6 Linear Operators


It turns out that operators occurring in quantum mechanics are linear. Â is a linear operator if and
only if it has the following two properties:
Aˆ  f ( x)  g ( x)  Aˆ f ( x)  Aˆ g ( x) (2.1)*
Aˆ c f ( x)  c Aˆ f ( x) (2.2)*
where f and g are arbitrary functions and c is an arbitrary constant (not necessarily real).
Examples of linear operators include x̂ 2 , d / dx , and d 2 / dx2 . Non linear operators are cos , and ( ) 2
, where ( ) 2 squares the function it acts on.
Useful identities in linear – operator manipulations are
 
Aˆ  Bˆ Cˆ  Aˆ Cˆ  Bˆ Cˆ (2.3)*
 
Aˆ Bˆ  Cˆ  Aˆ Bˆ  Aˆ Cˆ (2.4)*

Example
Is d / dx a linear operator ? Is a linear operator ?
Solution
(d / dx) f ( x)  g ( x)  df / dx  dg / dx  (d / dx) f ( x)  (d / dx) g ( x)
d / dxcf ( x)  c df ( x) / dx
So d / dx obeys (2.1)* and (2.2)* and is a linear operator. However
 f ( x)  g ( x )   f ( x )  g ( x )
So does not obey (2.1)* and is non-linear.
Example
Find the square of the operator d / dx  xˆ
Solution
To find the effect of d / dx  xˆ  we apply this operator to any arbitrary function f(x); letting
2

Dˆ  d / dx
Dˆ  xˆ   Dˆ  xˆ Dˆ  xˆ  Dˆ Dˆ  xˆ   xˆDˆ  xˆ 
2

 Dˆ 2  Dˆ xˆ  xˆDˆ  xˆ 2
 Dˆ 2  xˆDˆ  1  xˆDˆ  xˆ 2  Dˆ 2  2 xˆDˆ  xˆ 2  1
Where we have used (2.3)*, (2.4)* and (2.5)*
Note: Dˆ xˆf ( x)  xf ( x)  f ( x)  xf / ( x)  (1  xˆDˆ ) f ( x)
d
dx
This shows that
Dˆ xˆ  (1  xˆDˆ ) (2.5)*

2.1.7 HERMITIAN OPERATORS

9
Maxwell Mageto, PhD
Quantum Mechanics SPH 313

Operators known as Hermitian operators play a very special role in quantum mechanics because
their their eigenvalues are real. Hence it is Hermitian operators that are used to represent observables,
because the outcome of the observation must be a real number. We shall meet non-hermitian
operators, but as their eigenvalues are not guaranteed to be real, they do not correspond to
observables.

Hermiticity Condition

An operator  is Hermitian if it satisfies the following relation


 

 m n d   n  m d 

(2.6)*
For any two wave functions  m and  n . An alternative version of this definition is

  n d    m   n d
 
m (2.7)*
This expression is obtained by taking the complex conjugate of each term on the right hand side of
eqn (2.6)*

To demonstrate how this definition works, let's show that the momentum operator, pˆ   id / dx
Is Hermitian. We need to show that the operator satisfy eqn (2.7)* for a general wavefunction.

Replacing  with  id / dx only in the LHS of eqn (2.7)* and rearranging
 d   d n
 m   i dx  n dx   i  m dx dx   i  m d n

Next we integrate by parts  udv  uv   vdu




  i  m d n   i  m  n |

    n d m


 i0   n
d 
dx
 m dx 
The first term vanishes if we assume the functions  m and  n are for bound states and so that they
vanish at the limits. We can then rearrange the remaining integral as

 d 
  n   i  m  dx
 dx 
and so show that the momentum operator satisfies

 d   d 
   i dx  n dx   n   i dx m  dx

m

which is the definition of hermiticity eqn (2.7)*.


There are three important consequences of an operator being hermitian: Its eigenvalues are real;
its eigenfunctions corresponding to different eigenvalues are orthogonal to on another; and the set
of all its eigenfunctions is complete.

Exercise
1. Show that the operator  id / dx is hermitian
2. Show that the operator d / dx is not hermitian

10
Maxwell Mageto, PhD
Quantum Mechanics SPH 313

3. Use the fact that the momentum operator is hermitian to show that the kinetic energy operator
is hermitian. Hint: Show that if an operator,  , is hermitian, then the operator  2 =  
is hermitian.
4. Use the fact that the operator for position is just "multiply by position" to show that the
potential energy operator is hermitian. Hint: Potential energy is a function of position.
5. Use the results of the previous two questions to show that the hamiltonian, H, is hermitian

11
Maxwell Mageto, PhD

You might also like