You are on page 1of 11

QM notes

Mann Malviya
November 2021

The Wave Function


In classical mechanics we need to know only two variables, i.e., the momentum(⃗
p)
and the position(⃗r) any other quantity of interest can be calculated by using
p2 ⃗
these two parameter’s, e.g., KE = 2m , L = ⃗r × p⃗ etc. In quantum mechanics
we describe a ”quantum system” by a wave function ψ which is in some ways
is very different from classical mechanics because we are replacing two values
by infinite values(i.e. a function). In general the wave function is a function
of both position and time, we use Ψ(⃗r, t) for the time dependent wave function
and use ψ(⃗r) for the time independent wave function. Although often we use Ψ
and ψ interchangeably and it is understood by the context whether it is time
dependent or independent.
The main equation used in classical mechanics is newtons second law F⃗ = m⃗a
in quantum mechanics this is replaced by the Schrodinger equation:

∂Ψ −ℏ2 ∂ 2 Ψ
iℏ (x, t) = (x, t) + V (x)Ψ(x, t)
∂t 2m ∂x2

Interpretation of the Wave Function


Schrodinger had incorrectly interpreted the wave function as electric charge
density. The correct interpretation was given by Max Born. Born’s statistical
interpretation states that |Ψ(⃗r, t)|2 represents the probability density, the prob-
ability of finding a particle in a volume dV is hence given by |Ψ(⃗r, t)|2 dV . The
probability density is written as the norm-squared of the wave function because
in general the wave function is a complex valued function Ψ(⃗r, t) ∈ C. Hence
the probability density is strictly real |Ψ(⃗r, t)|2 = Ψ∗ Ψ.
In 1-D the probability of finding the particle in [a, b] is given as:
ˆ b
P(a, b) = |Ψ(⃗r, t)|2 dx
a

The statistical interpretation has introduced a kind of indeterminacy into QM.


This indeterminacy is unlike the one seen in thermodynamics and statistical

1
mechanics where the probabilities arise due to our ignorance, QM tells us that
there is something fundamentally probabilistic about the nature of reality.If
given all the information that one can get of a system he/she still cannot predict
with 100% accuracy the outcome of the experiment. There are two entirely
distinct kinds of physical processes:”ordinary” ones, in which the wave function
evolves under the Schrodinger equation, and ”measurements”, in which the wave
function suddenly and discontinuously collapses.

Probability
let the number of total events be N .then the probability of an event happening
is:
N (j)
P(j) =
N
hence the sum of the probability of each event happening is:

X
P(j) = 1
j=0

In general the average value of a function of j is given by:



X
⟨f (x)⟩ = f (j)P(j)
j=0

In statistics we call ⟨f (x)⟩ the average value, but in QM we call it the


expectation value. we need a numerical measure of the amount of ”spread” in
a distribution, with respect to the average. The most obvious way to do this
would be to find out how far each member is from the average,

∆j = j − ⟨j⟩

and compute the average of ∆j. The trouble is that you get zero.
X X X
⟨∆j⟩ = (j − ⟨j⟩)P(j) = jP(j) − ⟨j⟩ P(j)

= ⟨j⟩ − ⟨j⟩ = 0
We get around this problem by squaring before averaging:

σ 2 = ⟨(∆j)2 ⟩

σ 2 is called the variance of the distribution and σ is called the standard devia-
tion.
A theorem on variances:
X X 2
σ 2 = ⟨(∆j)2 ⟩ = (∆j)2 P(j) = (j − ⟨j⟩) P(j)

2
X 2
 X 2
X X
= j 2 + ⟨j⟩ − 2j⟨j⟩ P(j) = j 2 P(j) + ⟨j⟩ P(j) − 2⟨j⟩ jP(j)
2 2 2
= ⟨j 2 ⟩ + ⟨j⟩ − 2⟨j⟩ = ⟨j 2 ⟩ − ⟨j⟩
hence the standard deviation σ can be written as:
q
2
σ = ⟨j 2 ⟩ − ⟨j⟩

2 2
In general ⟨j 2 ⟩ is not equal to ⟨j⟩ . since σ 2 is plainly non-negative ⟨j 2 ⟩ ≥ ⟨j⟩ .
and the two are equal only when σ = 0, which is the case for distribution with
no spread at all, i.e., each member has the same value.
we shall now deal with continuous variables:
The probability of a member(chosen at random) lies between x and (x + dx)

P(x, x + dx) = ρ(x)dx

ρ(x) is called the probability density. The probability that x lies between a and
b (a finite interval) is given by:
ˆ b
P(a, b) = ρ(x)dx
a

The rules we deduced for discrete distributions translate in the obvious way:
ˆ +∞
ρ(x)dx = 1
−∞
ˆ +∞
⟨x⟩ = xρ(x)dx
−∞
ˆ +∞
⟨f (x)⟩ = f (x)ρ(x)dx
−∞
2
σ 2 = ⟨(∆x)2 ⟩ = ⟨x2 ⟩ − ⟨x⟩

Normalisation
The statistical interpretation says that |Ψ(x, t)|2 is the probability density.
Hence it follows that the integral of the probability density over all space should
be 1(as the particle should be present somewhere).
ˆ +∞
|Ψ(x, t)|2 dx = 1
−∞

3
Any wave function that satisfies the Schrodinger equation must also be normal-
isable. For a wave function to be normalisable its must tend to 0 as |x| tends to
∞. To normalize a wave function ψ(x) we simply add a constant N in front of
it such that its satisfies the normalisation condition. We shall now prove that
once the wave function is normalised at t = 0, it remains normalised under it’s
evolution by the Schrodinger equation.
To prove: ˆ +∞ 
d 2
|Ψ(x, t)| dx = 0
dt −∞

By the Leibniz rule the derivative becomes a partial derivative when taken inside
the integral and then we apply the product rule:
ˆ +∞ ˆ +∞ ˆ +∞ 
∂|Ψ|2 ∂(Ψ∗ Ψ) ∂Ψ∗

∗ ∂Ψ
dx = dx = Ψ +Ψ dx
−∞ ∂t −∞ ∂t −∞ ∂t ∂t

The Schrodinger equation reads:

∂Ψ ℏ2 ∂ 2 Ψ
iℏ =− +VΨ
∂t 2m ∂x2
∂Ψ iℏ ∂ 2 Ψ iV
= − Ψ
∂t 2m ∂x2 ℏ
the complex conjugate of this equation is:

∂Ψ∗ iℏ ∂ 2 Ψ∗ iV ∗
=− 2
+ Ψ
∂t 2m ∂x ℏ
ˆ +∞  
iℏ ∂ 2 Ψ iV iℏ ∂ 2 Ψ∗
  
iV ∗
= Ψ∗ − Ψ + Ψ − + Ψ dx
−∞ 2m ∂x2 ℏ 2m ∂x2 ℏ
ˆ +∞ 
∂2Ψ ∂ 2 Ψ∗

iℏ
= Ψ∗ 2 − Ψ dx
2m −∞ ∂x ∂x2
we now apply integration by parts on the second term:
ˆ +∞  ∗ +∞ ˆ +∞
∂ 2 Ψ∗ ∂Ψ ∂Ψ∗ ∂Ψ
=⇒ Ψ 2
dx = Ψ − dx
−∞ ∂x ∂x −∞ −∞ ∂x ∂x

as |x| → ∞, ψ(x) → 0 and dx (x) →0
ˆ +∞ ˆ +∞
∂ 2 Ψ ∂Ψ∗ ∂Ψ
   
iℏ iℏ ∂ ∂Ψ
= Ψ∗ 2 + dx = Ψ∗ dx
2m −∞ ∂x ∂x ∂x 2m −∞ ∂x ∂x
 +∞
iℏ ∗ ∂Ψ
= Ψ =0
2m ∂x −∞
hence we gave proved once the wave function is normalised it always remains
normalised.

4
we can also prove the conservation of Probability in an alternative way: Let’s
look at how the probability densityP = |Ψ(⃗r, t)|2 changes with time,

∂P ∂|Ψ|2 ∂Ψ ∂Ψ∗
= = Ψ∗ +Ψ
∂t ∂t ∂t ∂t
the Schrodinger equation reads:

∂Ψ ℏ2 2
iℏ =− ∇ Ψ+VΨ
∂t 2m
∂Ψ iℏ 2 iV
= ∇ Ψ− Ψ
∂t 2m ℏ
and

∂Ψ iℏ 2 ∗ iV ∗
=− ∇ Ψ + Ψ
∂t 2m ℏ
if we substitute this in our original equation we have:
   
∂P ∗ iℏ 2 iV iℏ 2 ∗ iV ∗
=Ψ ∇ Ψ− Ψ +Ψ − ∇ Ψ + Ψ
∂t 2m ℏ 2m ℏ

iℏ
Ψ∗ ∇2 Ψ − Ψ∇2 Ψ∗

=
2m
iℏ
= ∇ · (Ψ∗ ∇ψ − Ψ∇Ψ∗ )
2m
we finally end up with a continuity equation, these equation’s arise whenever
there is a conserved quantity:
∂P
+∇·J =0
∂t
where J is the Probability current:
iℏ
J =− (Ψ∗ ∇Ψ − Ψ∇Ψ∗ )
2m
To see that the continuity equation implies a conservation law, consider some
region V ⊂ R3 with boundary S. We then integrate to get the probability PV
that the particle lies somewhere in this region.
ˆ
PV (t) = d3 rP (⃗r, t)
V

The continuity equation tells us that the Probability changes as:


ˆ ˆ
∂PV 3 ∂P
= d r (⃗r, t) = − d3 r∇ · J
∂t V ∂t V

5
´ ¸
we can now apply the Divergence theorem V (∇ · ⃗v )dV = S ⃗v · dS
˛
∂PV
= − J · dS
∂t S

we see that the probability that the particle lies in V can change, but only if
there is a flow of probability through the surface S that bounds V . If we know
for sure that J = 0 everywhere on the surface S, then the probability that the
particle is in the region V doesn’t change: ∂P∂t = 0
V

The importance of the continuity equation is that it doesn’t just tell us that
some quantity is conserved: it tells us that the object is conserved locally. In
the present context, this quantity is probability. The probability density can’t
just vanish in one region of space, to reappear in some far flung region of the
universe. Instead, the evolution of the probability density is local. If it does
change in some region of space then it’s because it has moved into a neighbouring
region. The probability current J tells you how it does this.
As a special case, If we take the region V to be all of space, so V = R3 , then
ˆ
Panywhere = d3 r|Ψ(⃗r, t)|2

as we’ve already seen Panywhere = 1. The change in the Probability is, using
ˆ
∂Panywhere
=− dS · J
∂t 2
S∞

2
where S∞ is the asymptotic 2-sphere at the infinity of R3 . But any normalised
wave function must have ψ → ∞ as |x| → ∞ and J → 0as |x| → ∞. Further-
more, you can check that J falls off quickly enough so that
∂Panywhere
=0
∂t
This is the the statement that if the wave function is normalised at some time,
then it remains normalised for all time.

The Momentum Operator


The expectation value of an operator Q̂ is
ˆ +∞ ˆ +∞
⟨Q⟩ = Q|Ψ(x, t)|2 = Ψ∗ Q̂Ψdx
−∞ −∞

The expectation value is the average of measurements on an ensemble of


identically-prepared systems, not the average of repeated measurements on one
and the same system.
The position operator x̂ tells us to multiply by x .The expectation value of
x is ˆ +∞
⟨x⟩ = x|Ψ(x, t)|2 dx
−∞

6
⟨x⟩ does not mean if you measure the position of the particle again and again,
´ +∞
−∞
x|Ψ|2 dx is the average of the result you will get. To keep things from
getting cluttered I will omit writing the limits of integration.
ˆ ˆ
d⟨x⟩ d ∂|Ψ|2
= x|Ψ|2 dx = x dx
dt dt ∂t
ˆ 
∂Ψ∗

∗ ∂Ψ
= x Ψ +Ψ dx
∂t ∂t
from the Schrodinger equation we get:

∂Ψ iℏ ∂ 2 Ψ iV
= − Ψ
∂t 2m ∂x2 ℏ
and

∂Ψ iℏ ∂ 2 Ψ∗ iV ∗
=− 2
+ Ψ
∂t 2m ∂x ℏ
on substituting this in our original equation we get:
ˆ  2
∂ 2 Ψ∗

iℏ ∗∂ Ψ
= x Ψ −Ψ
2m ∂x2 ∂x2

using integration by parts we get:


"  +∞ ˆ  #
∂Ψ∗ ∂Ψ∗

iℏ ∗ ∂Ψ dx ∗ ∂Ψ
= x Ψ −Ψ − Ψ −Ψ dx
2m ∂x ∂x −∞ dx ∂x ∂x
ˆ 
∂Ψ∗

iℏ ∗ ∂Ψ
=− Ψ −Ψ dx
2m ∂x ∂x
integrating the second term by parts gives:
ˆ ˆ
∂Ψ∗ +∞ ∂Ψ
Ψ dx = Ψ [Ψ∗ ]−∞ − Ψ∗ dx
∂x ∂x
now substituting this back gives us:
ˆ  
iℏ ∗ ∂Ψ ∗ ∂Ψ
= Ψ +Ψ dx
2m ∂x ∂x
ˆ +∞
d⟨x⟩ iℏ ∂Ψ
= Ψ∗ dx
dt m −∞ ∂x
multiplying both sides by m
ˆ +∞
d⟨x⟩ ∂
m = ⟨p⟩ = Ψ∗ (−iℏ )Ψdx
dt −∞ ∂x

7
so we finally get our momentum operator:

p̂x → −iℏ
∂x

p̂ → −iℏ∇
this was in some sense just a proof that the momentum operator is equal to −iℏ∇ ⃗
as we used the Schrodinger equation which already contains the momentum
operator in the Kinetic energy term of the Hamiltonian(the total energy of the
system). The Schrodinger equation cannot be derived from first principles just
like F = ma cannot be derived. The Schrodinger equation is a model that fits
practical data with very high accuracy. Therefore even the momentum operator
i
can only be motivated using the De Broglie plane wave Ae ℏ (px±Et)
all classical observable’s can be replaced with operators in quantum mechan-
ics and all classical observable can be represented by some combination of ⃗x and
p⃗ therefore we write a general operator as Q̂(x̂, p̂)
ˆ +∞
⟨Q(x̂, p̂)⟩ = Ψ∗ [Q(x̂, p̂)]Ψdx
−∞

Note: d⟨p⟩ ∂V
dt = ⟨− ∂x ⟩ which is an ´ instance of the Ehrenfest’s theorem, can also
+∞
be proved by replacing ⟨p⟩ with −∞ Ψ∗ (−iℏ∂x )Ψdx taking the ∂t inside the
integral(by Leibniz rule) and then simplifying using the Schrodinger equation.
To prove this explicitly is left as an exercise for the reader.

Operators
The most common operators used in QM are:
• The position operator: x̂ → x
• The momentum operator: p̂ → −iℏ∇

• The Energy operator: Ê → iℏ∂t


we shall now motivate the momentum operator p̂x and the energy operator Ê
using the De Broglie plane wave.

ψ(x, t) = Aei(kx−ωt)

E = ℏω
p = ℏk
i
ψ(x, t) = Ae ℏ (px−Et)
if we differentiate ψ(x, t) wrt to x,
∂ψ i i
(x, t) = pAe ℏ (px−Et)
∂x ℏ

8
∂ψ
−iℏ (x, t) = pψ(x, t)
∂x
from which we get the momentum operator p̂x and if we consider a 3-D plane
i
wave ψ(⃗r, t) = Ae ℏ (⃗p·⃗r−Et) we can calculate p̂y and p̂z in a similar manner.

∂ ∂ ∂
p̂x = −iℏ p̂y = −iℏ p̂z = −iℏ
∂x ∂y ∂z


p̂ = −iℏ∇
i
to derive the energy operator Ê we differentiate the plane wave(Ae ℏ (⃗p·⃗r−Et) )
wrt t.
∂ψ i i
(⃗r, t) = − EAe ℏ (⃗p·⃗r−Et)
∂t ℏ
∂ψ
iℏ (⃗r, t) = Eψ(⃗r, t)
∂t

Ê = iℏ
∂t
We shall now define a commutator and check if the position operator x̂ and
momentum operator p̂ commute.

[Â, B̂] = ÂB̂ − B̂ Â

If two operators commute, i.e., [Â, B̂] = 0 then both quantities can be known
for a quantum system and if the do not commute both the quantities cannot be
know simultaneously to an arbitrary accuracy(a common example is the position
and momentum operator where [x̂, p̂x ] = iℏ).
checking if x̂ and p̂x commute: dealing with operators is kind of slippery so
as to not make any mistakes we shall use a test function f which we will frop
after finding [x̂, p̂].

ℏ ∂f ℏ ∂(xf )
[x̂, p̂]f = x̂p̂f − p̂x̂f = x −
i ∂x i ∂x
 
ℏ ∂f ℏ ∂f ℏ
=x − x + f = − f = iℏf
i ∂x i ∂x i
[x̂, p̂] = iℏ

The Schrodinger Equation


As seen earlier the energy operator is Ê = iℏ∂t if we act the energy of operator
on a wave function ψ(x, t) we get

∂ψ
Êψ(x, t) = iℏ (x, t)
∂t

9
the total energy of system is the kinetic energy plus the potential energy, since
the Hamiltonian for a system is its total energy it is common to write Ĥ for the
energy operator instead of Ê

p̂2
Ĥ = T̂ + V̂ = + V̂
2m

if we now substitute p̂ = −iℏ ∂x we get:

ℏ2 ∂
Ĥ = − + V̂
2m ∂x2
if we multiply both sides of the equation by ψ(x, t) we get,

ℏ2 ∂ 2 ψ
Ĥψ(x, t) = − (x, t) + V (x)ψ(x, t)
2m ∂x2

if we now replace Ĥ → iℏ ∂t we get the time dependent Schrodinger equation:

∂ψ ℏ2 ∂ 2 ψ
iℏ (x, t) = − (x, t) + V (x)ψ(x, t)
∂t 2m ∂x2

Solving The Schrodinger Equation


The Schrodinger equation is a Linear Partial differential equation, as with most
PDE’s we shall try using the variable separation method, we will hence end
up with two ordinary differential equations. The time-dependent Schrodinger
equation reads:

∂Ψ ℏ2 ∂ 2 Ψ
iℏ (x, t) = − (x, t) + V (x)Ψ(x, t)
∂t 2m ∂x2
if we assume Ψ(x, t) = ψ(x)ϕ(t)

∂ψ(x)ϕ(t) ℏ2 ∂ 2 ψ(x)ϕ(t)
iℏ =− + V (x)ψ(x)ϕ(t)
∂t 2m ∂x2
∂ϕ ℏ2 ϕ(t) ∂ 2 ψ
iℏψ(x) (t) = − (x) + V (x)ψ(x)ϕ(t)
∂t 2m ∂x2
On dividing both sides by ψ(x)ϕ(t)

iℏ ∂ϕ ℏ2 1 ∂ 2 ψ
(t) = − (x) + V (x)
ϕ(t) ∂t 2m ψ(x) ∂x2

we can now notice that the LHS depends only on t and the RHS depends only
on x. Hence both sides must be constant(E).

iℏ dϕ ℏ2 1 d2 ψ
E= (t) E=− (x) + V (x)
ϕ(t) dt 2m ψ(x) dx2

10
dϕ ℏ2 d 2 ψ
Eϕ(t) = iℏ (t) Eψ(x) = − (x) + V (x)Ψ(x)
dt 2m dx2
we can easily solve the first differential equation for ϕ(t)


Eϕ(t) = iℏ (t)
dt
ˆ ˆ
−iE dϕ iEt
dt = =⇒ ln|ϕ(t)| = − +C
ℏ ϕ ℏ
iEt iEt
ϕ(t) = e− ℏ +C = Ae− ℏ

The second differential equation that we obtained is called the time-independent


Schrodinger equation. We can solve for ψ(x) only if we are given a specific
potential V (x). It should be noted that we could not have used the method of
variable separation if the potential is time dependent V (x, t). Time dependent
potentials are covered under Quantum Dynamics. Strictly speaking ψ(x) that
has appeared in the time-independent Schrodinger equation is not called the
wave function,ψ(x) is called a ”stationary state” as it is time-independent.
The wave function is therefore:
iEt
Ψ(x, t) = ψ(x)e− ℏ

Note that we Have dropped off the A when finally writing the wave function as
the A will get absorbed into the Normalisation constant that will be added so
that the wave function satisfies the normalisation condition.

11

You might also like