0% found this document useful (0 votes)
33 views26 pages

Quantum Mechanics Notes

Uploaded by

xamoha8311
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
33 views26 pages

Quantum Mechanics Notes

Uploaded by

xamoha8311
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 26

1

The following notes on Quantum Mechanics are based on “Introduction to Quantum Mechan-
ics” by David J. Griffiths.

Let us consider a particle of mass m constrained to move along the x axis, subject to some
force F (x, t). In classical mechanics, the problem is to determine the position of the particle
at any given time x(t). Newton’s second law gives us F = ma, where a is the acceleration of

the particle. For conservative systems, F = − ∂V


∂x
, where V is the potential energy function.
Therefore, the equation F = ma can be rewritten as

d2 x ∂V
m = − . (1)
dt2 ∂x

Eq. 1, which is the time evolution equation of a particle in classical mechanics, along with the
initial conditions determine x(t).

In quantum mechanics the time evolution of a particle is determined by a second order partial
differential equation known as the time-dependent Schrödinger equation 2

∂Ψ ~2 ∂ 2 Ψ
i~ =− + V Ψ. (2)
∂t 2m ∂x2

Note that while we try to determine the position of the particle in classical mechanics, in
quantum mechanics we try to determine the particle’s wave function. We solve eq. 2 to find
out the wave function Ψ(x, t) of the particle for all future times, given the initial conditions

Ψ(x, 0). The obvious questions are: what is the wave function and how should it be related to
the position of the particle? The answer is provided by Max Born’s statistical interpretation
of the wave function. It states that

Z
|Ψ(x, t)|2 dx = probability of finding the particle between x and x + dx at time t

Note that while Ψ is complex, |Ψ|2 = ΨΨ∗ , where Ψ∗ is the complex conjugate of Ψ, is real and

non-negative. Fig. 1(a) shows the modulus square of the wave function Ψ plotted against x.
The probability of finding the particle between a and b is indicated in the figure. If the particle
is detected at some point c, by some measurement, the modulus square of the wave function
immediately thereafter is plotted in Fig. 1(b).
2

|(x,t)|2

(a)

a b x

|(x,t)|2

(b)

c x
Z b
2
Figure 1: (a) |Ψ(x, t)| plotted against x, the shaded area |Ψ(x, t)|2 dx is the probability
a
of finding the particle between a and b, (b) Collapse of the wave function: graph of |Ψ(x, t)|2
plotted against x, immediately after a measurement has found the particle at point c.

Digression: Let us discuss some concepts of probability which will be useful in our future

discussions of quantum mechanics.


Discrete variables: Example – Consider a room containing 14 people whose ages in years, are
as listed below
1 person aged 14, 1 person aged 15, 3 persons aged 16, 2 persons aged 22, 2 persons aged 24,

5 persons aged 25
Let N(j) represent the number of people of age j. Therefore,

N(14) = 1 N(15) = 1 N(16) = 3 N(22) = 2 N(24) = 2 N(25) = 5

The total number of people in the room is


X
N= N(j)
j=0

Select one individual at random from this group. If ρ(j) is the probability of getting age j,
3

then
N(j)
ρ(j) =
N

The sum of all the probabilities is 1.


X
ρ(j) = 1
j=0

Average value of j denoted by hji is


X
jN(j)

j=0 X
hji = = jρ(j)
N j=0

Average value of a random variable hji is called the expectation value (which is actually a

misnomer) of j in quantum mechanics. The average value of some function of j is given by


X
hf (j)i = f (j)ρ(j)
j=0

Note that two widely differing distributions can have the same median, same average, same
most probable value and the same number of elements. To differentiate between the two

distributions we need a numerical measure of the amount of spread in a distribution with


respect to the average. Now, deviation of individual value from average is

∆j = j − hji

Average of ∆j


X ∞
X ∞
X
h∆ji = (j − hji)ρ(j) = jρ(j) − hji ρ(j) = hji − hji = 0
j=0 j=0 j=0

So, h∆ji will not work as a measure of the spread of the distribution of the random variable j.

To get around the sign problem, square before averaging.

σ 2 = h(∆j)2 i
4

σ 2 is known as the variance of the distribution, while σ is called the standard deviation. Let
us study a theorem on variances.

σ 2 = h(∆j)2 i

X
= (∆j)2 ρ(j)
j=0

X
= (j − hji)2 ρ(j)
j=0
X∞
= (j 2 − 2jhji + hji2 )ρ(j)
j=0
X∞ ∞
X ∞
X
= j 2 ρ(j) − 2hji jρ(j) + hji2 ρ(j)
j=0 j=0 j=0

= hj 2 i − 2hjihji + hji2

= hj 2 i − hji2

σ 2 is non-negative by definition. Therefore, hj 2 i ≥ hji2 . The two are equal only when σ = 0.
Continuous variables: Example – Probability that the age of a person chosen at random lies

in some interval between x and x + dx is ρ(x) dx, where ρ(x) is the probability density. The
probability that x lies between a and b (a finite interval) is given by the integral of ρ(x)

Z b
Pab = ρ(x) dx
a

Some rules for continuous distributions

Z ∞
1= ρ(x) dx
Z−∞

hxi = xρ(x) dx
Z−∞

hf (x)i = f (x)ρ(x) dx
−∞

σ 2 = h(∆x)2 i = hx2 i − hxi2


5

Normalization of wave function:

Z ∞
|Ψ(x, t)|2 dx = 1. (3)
−∞

The particle has to be somewhere. If Ψ(x, t) is a solution of the Schrödinger equation, AΨ(x, t)
is also a solution where A is any complex constant. We must pick the undetermined multiplica-
tive factor so as to ensure that eq. 3 is satisfied. This process is called normalizing the wave

function. For some solutions to the Schrödinger equation the integral is infinite. Also, there
is the trivial solution Ψ = 0. Such non-normalizable solutions cannot represent particles and
must be rejected.
Suppose I have normalized the solution at t = 0. Does it stay normalized as time goes on and
Ψ evolves?

Schrödinger equation has the remarkable property that it automatically preserves the normal-
ization of the wave function. This crucial feature makes the Schrödinger equation compatible
with the statistical interpretation of the wave function.

Proof.

∞ ∞
d
Z Z
2 ∂
|Ψ(x, t)| dx = |Ψ(x, t)|2 dx
dt −∞ −∞ ∂t

By the product rule,

∂ 2 ∂ ∗ ∗ ∂Ψ ∂Ψ∗
|Ψ| = (Ψ Ψ) = Ψ + Ψ
∂t ∂t ∂t ∂t

From Schrödinger equation,

∂Ψ ~2 ∂ 2 Ψ
i~ =− +VΨ
∂t 2m ∂x2
∂Ψ ~2 ∂ 2 Ψ V Ψ
=− +
∂t i~2m ∂x2 i~
2
i~ ∂ Ψ iV Ψ
= −
2m ∂x2 ~
6

Taking the complex conjugate of the above equation,

∂Ψ∗ i~ ∂ 2 Ψ∗ iV Ψ∗
=− +
∂t 2m ∂x2 ~

So,

∂ ∂Ψ ∂Ψ∗
|Ψ|2 = Ψ∗ + Ψ
∂t ∂t ∂t
i~ ∂ 2 Ψ iV Ψ i~ ∂ 2 Ψ∗ iV Ψ∗
   

=Ψ − + − + Ψ
2m ∂x2 ~ 2m ∂x2 ~
2
∂ 2 Ψ∗
 
i~ ∗∂ Ψ
= Ψ − Ψ
2m ∂x2 ∂x2
∂Ψ∗
  
∂ i~ ∗ ∂Ψ
= Ψ − Ψ
∂x 2m ∂x ∂x
Z ∞ Z ∞
d ∂
∴ |Ψ(x, t)|2 dx = |Ψ(x, t)|2 dx
dt −∞ ∂t
Z−∞∞
∂Ψ∗
  
∂ i~ ∗ ∂Ψ
= Ψ − Ψ dx
−∞ ∂x 2m ∂x ∂x
∞
∂Ψ∗

i~ ∗ ∂Ψ
= Ψ − Ψ
2m ∂x ∂x −∞

∂Ψ ∂Ψ∗
Ψ(x, t), Ψ∗ (x, t) must go to zero and , must be finite as x goes to ±∞, otherwise the
∂x ∂x
wave function would not be normalizable.


d
Z
∴ |Ψ(x, t)|2 dx = 0
dt −∞

The integral is independent of time, if Ψ(x, t) is normalized at t = 0, it stays normalized for all
future times.

For a particle in state Ψ(x, t), the expectation value of x is

Z ∞
hxi = x|Ψ(x, t)|2 dx
−∞

What does it really mean? The expectation value is the average of repeated measurements on
an ensemble of identically prepared systems, not the average of repeated measurements on one
and the same system.
7

Let us now find out the expectation value of momentum. As time goes on hxi will change
because of the time dependence of Ψ(x, t), as determined by the Schrödinger equation. How

fast does the particle move?


dhxi
Z

= |Ψ(x, t)|2 dx
x
dt ∂t
Z−∞

∂Ψ∗
  
∂ i~ ∗ ∂Ψ
= x Ψ − Ψ dx
−∞ ∂x 2m ∂x ∂x

Use integration by parts

d dg df
(f g) = f +g
dx dx dx
Z b Z b b
dg df
∴ f dx = − g dx + f g
a dx a dx a

∂Ψ∗
 
∗ ∂Ψ
Here, f = x, g = Ψ − Ψ and f g → 0 as x → ±∞, then
∂x ∂x


∂Ψ∗
 
dhxi
Z
i~ ∗ ∂Ψ
=− Ψ − Ψ dx
dt 2m −∞ ∂x ∂x

Perform another integration by parts on the second term, note that ΨΨ∗ → 0 as x → ±∞

∞ ∞
∂Ψ∗
Z Z
∂Ψ
Ψ dx = − Ψ∗ dx
−∞ ∂x −∞ ∂x

Then,


dhxi
Z
i~ ∂Ψ
=− Ψ∗ dx
dt m −∞ ∂x

What does this mean? We do not have the rate of change of position of the particle, which
would be the velocity of the particle. Postulate: The expectation value of the velocity is equal to
the time derivative of the expectation value of position and the expectation value of momentum
is equal to mass times the expectation value of velocity.

dhxi
hvi =
dt
8

and

dhxi
hpi = mhvi = m
Z ∞ dt
∂Ψ
= −i~ Ψ∗ dx
−∞ ∂x

We rewrite x and p as

Z ∞
hxi = Ψ∗ (x)Ψ dx (4)
Z−∞
∞  
∗ ~ ∂
hpi = Ψ Ψ dx (5)
−∞ i ∂x

~ ∂
We say that the operator x represents position and the operator represents momentum.
i ∂x
All classical dynamical variables, like kinetic energy, angular momentum etc. can be expressed
in terms of position and momentum. Their expectation values can be written as

Z ∞ 
∗~ ∂
hQ(x, p)i = Ψ Q x, Ψ dx .
−∞ i ∂x

The Heisenberg uncertainty principle: One can construct an ensemble of states such that

repeated position measurements will be very close together but momentum measurements on
this state will be widely scattered. Quantitatively,

~
σx σp ≥ (6)
2

where σx and σp are the standard deviations of the distributions of position and momentum
measurements on an ensemble of identically prepared states. Eq. 6 sets the fundamental limit
of measurement dictated by the principles of quantum mechanics.

Solution of the Schrödinger equation: We propose to convert the second order partial differential
equation to ordinary differential equations using suitable substitutions. We rewrite the time

dependent Schrödinger equation

∂Ψ ~2 ∂ 2 Ψ
i~ =− +VΨ
∂t 2m ∂x2
9

Assume V is independent of t. We invoke the method of separation of variables, look for


solutions that are simple products of the form

Ψ(x, t) = ψ(x)φ(t)

and construct the most general solution later. Then,

∂Ψ dφ ∂2Ψ d2 ψ
=ψ and = φ .
∂t dt ∂x2 dx2

Rewrite eq. 2 as

dφ ~2 d2 ψ
i~ψ =− φ + V ψφ
dt 2m dx2

Divide both sides of the equation by ψφ, then

1 dφ ~2 1 d2 ψ
i~ =− +V
φ dt 2m ψ dx2

Left hand side of the above equation is a function of t alone while the right hand side is a
function of x alone. The only way this could be true is if both sides are equal to a constant.
Then,

1 dφ
i~ =E
φ dt
dφ iE
or =− φ
dt ~
dφ iE
or = − dt
φ ~

Integrating, we get


Z Z
iE
= − dt
φ ~
iE
ln φ = − t+k
~
φ = e−iEt/~ c
10

If we absorb the constant c in φ then,

φ = e−iEt/~ (7)

and,

~2 1 d2 ψ
− +V =E
2m ψ dx2
~2 d2 ψ
or − + V ψ = Eψ (8)
2m dx2

Eq. 8 is called the time independent Schrödinger equation and can be solved (analytically or
numerically) given the form of the function V . Let us now discuss about some of the properties
of the solutions of the time dependent Schrödinger equation.

1. The solutions are stationary states.

Ψ(x, t) = ψ(x)e−iEt/~ depends on t.

But,

|Ψ(x, t)|2 = Ψ∗ Ψ = ψ ∗ eiEt/~ ψe−iEt/~ = |ψ(x)|2 does not depend on t.

The same thing happens in calculating the expectation value of any dynamical variable.

Z ∞   Z ∞  
∗ ~ ∂ ∗ ~ ∂
hQ(x, p)i = Ψ Q x, Ψ dx −→ hQ(x, p)i = ψ Q x, ψ dx
−∞ i ∂x −∞ i ∂x

Every expectation value is constant in time.

2. The solutions are states of definite total energy. In classical mechanics, the total energy

(kinetic plus potential) is called the Hamiltonian

p2
H(x, p) = + V (x).
2m
11

The corresponding Hamiltonian operator, obtained by the canonical substitution p →


~ ∂
is
i ∂x
~2 ∂ 2
Ĥ = − + V (x).
2m ∂x2

Therefore, the time independent Schrödinger equation can be written as

Ĥψ = Eψ

and the expectation value of the total energy is

Z ∞ Z ∞

hHi = ψ Ĥψ dx = E |ψ|2 dx = E
−∞ −∞

Also

Ĥ 2 ψ = Ĥ(Ĥψ) = E Ĥψ = E 2 ψ
Z Z
2 ∗ 2 2
∴ hH i = ψ Ĥ ψ dx = E |ψ|2 dx = E 2

So the variance of H is

2
σH = hH 2 i − hHi2 = E 2 − E 2 = 0.

A separable solution has the property that every measurement of the total energy is cer-
tain to return the value E.

3. The general solution is a linear combination of the separable solutions. The time indepen-

dent Schrödinger equation yields an infinite collection of solutions (ψ1 (x), ψ2 (x), ...) each
with its associated value of the separation constant (E1 , E2 , ...), thus there is a different
wave function for each allowed energy

Ψ1 (x, t) = ψ1 (x)e−iE1 t/~

Ψ2 (x, t) = ψ2 (x)e−iE2 t/~


12

............

The time dependent Schrödinger equation has the property that any linear combination
of the solutions is itself a solution. So we can construct a general solution


X
Ψ(x, t) = cn ψn (x)e−iEn t/~ (9)
n=1

We can find out the constants cn , given the initial conditions.

Solution of the time independent Schrödinger equation for some given potentials:

The infinite square well

Suppose 

0,
 0 ≤ x ≤ a,
V (x) =

∞, otherwise.

A particle in this potential (see Fig. 2) is completely free except at the two ends (x = 0 and

V(x)

0 a x

Figure 2: The infinite square well.

x = a) where an infinite force prevents it from escaping. Outside the well ψ(x) = 0. Inside the

well where V = 0, the time independent Schrödinger equation reads

~2 d2 ψ
− = Eψ
2m dx2 √
d2 ψ 2mE
or = −k 2 ψ where k =
dx2 ~
13

The general solution to the above equation is

ψ(x) = A sin(kx) + B cos(kx)

where A and B are arbitrary constants, fixed by the boundary conditions of the problem.

Ordinarily, both ψ and are continuous, but where the potential goes to ∞ only the first of
dx
these applies. Continuity of ψ(x) requires that

ψ(0) = ψ(a) = 0

Now, ψ(0) = A sin 0 + B cos 0 = B = 0

∴ ψ(x) = A sin(kx)

Then, ψ(a) = A sin(ka) = 0

Since, A = 0 makes it a trivial solution,

∴ sin(ka) = 0

or, ka = 0, ±π, ±2π, ...

k = 0 is not acceptable, since ψ(x) = 0, and the negative solutions give nothing new since
sin(−θ) = − sin(θ) and we can absorb the negative sign into A. So the distinct solutions are


kn = with n = 1, 2, 3, ...
a

The boundary condition at x = a does not determine the constant A, but rather the constant
k and hence the possible values of E are

~2 kn2 n2 π 2 ~2
En = = with n = 1, 2, 3, ...
2m 2ma2

A quantum particle in the infinite square well cannot have any energy, it has to be one of these
14

special allowed values (quantization of energy). To find A, we normalize ψ.

Z a
|A|2 sin2 (kx) dx = 1
0

1 − cos 2θ
We know, cos 2θ = cos2 θ − sin2 θ and cos2 θ + sin2 θ = 1, ∴ sin2 θ = .
2

a
(1 − cos(2kx))
Z
∴ |A|2 dx = 1
0 2
a a
|A|2

sin(2kx)
or x − =1
2 0 2k 0
|A|2 a
or =1
2
2
∴ |A|2 =
a

The above equation


r only determines the magnitude of A. It is simplest to pick the positive real
2
root – A = , the phase of A carries no physical significance. Inside the well, the solutions
a
are r
2  nπx 
ψn (x) = sin
a a

The time independent Schrödinger equation has an infinite set of solutions (one for each positive
integer n). The first three wave functions are plotted in Fig. 3. ψ1 (x) which carries the lowest
energy is called the ground state, the others whose energies increase in proportion to n2 , are

called excited states.

1(x) 2(x) 3(x)

0 a x 0 a x 0 a x

Figure 3: The first three wave functions plotted against x.

Let us discuss some properties of the function ψn (x).

1. They are alternately even and odd with respect to the center of the well, ψ1 (x) is even,
ψ2 (x) is odd, ψ3 (x) is even and so on.
15

2. As you go up in energy, each successive state has one more node (zero crossing) – ψ1 (x)
has none (the end points don’t count), ψ2 (x) has one, ψ3 (x) has two and so on.

3. They are mutually orthogonal in the sense that

Z ∞

ψm (x)ψn (x) dx = 0 whenever m 6= n.
−∞

Proof.

∞ a
2
Z Z  mπx   nπx 

ψm (x)ψn (x) dx = sin sin dx
−∞ a 0 a a

Use the following trigonometric identities to transform the above integral, cos(A − B) =
cos A cos B + sin A sin B and cos(A + B) = cos A cos B − sin A sin B

1 a h  mπx nπx 
Z  mπx nπx i
= cos − − cos + dx
a 0 a a a a
  a
1 (m − n)πx 1 (m + n)πx
= sin − sin
(m − n)π a (m + n)π a 0
 
sin(m − n)π sin(m + n)π
= −
(m − n)π (m + n)π
=0

If m = n,

Z ∞
ψn∗ (x)ψn (x) dx = 1 normalization condition
Z −∞


∴ ψm (x)ψn (x) dx = δmn Kronecker delta
−∞



0, if m 6= n,

δmn =

1, if m = n.

ψn (x)’s are orthonormal.

4. They are complete, in the sense that any other function f (x) can be expressed as a linear
16

combination of them

∞ r ∞
X 2X  nπx 
f (x) = cn ψn (x) = cn sin
n=1
a n=1 a

The coefficients cn can be evaluated by exploiting the orthonormality of ψn . Multiply



both sides of the above equation by ψm (x) and integrate

Z ∞ ∞
X Z ∞
∗ ∗
ψm (x)f (x) dx = cn ψm (x)ψn (x) dx
−∞ n=1 −∞

X∞
= cn δmn
n=1

= cm
Z ∞
∴ cn = ψn∗ (x)f (x) dx (10)
−∞

These four properties are extremely powerful and are not special to the infinite square
well. The first is true whenever the potential itself is a symmetric function. The second
is universal regardless of the shape of the potential. Orthogonality and completeness are

also quite general.

The stationary states of the infinite square well are

n2 π 2 ~
 
r
2  nπx  −i t
Ψn (x, t) = sin e 2ma2
a a

The most general solution to the time dependent Schrödinger equation is a linear combination

of stationary states
n2 π 2 ~
 
∞ r
2  nπx  −i t
2ma2
X
Ψ(x, t) = cn sin e
n=1
a a

Any prescribed initial wave function, Ψ(x, 0) can be fitted to the general solution by appropriate
choice of the coefficients cn .

X
Ψ(x, 0) = cn ψn (x) (11)
n=1

The completeness of ψs guarantees that one can always express Ψ(x, 0) in this way, and their
17

orthonormality licenses the use of Fourier’s trick, eq. 10, to determine the actual coefficients –

r Z a
2  nπx 
cn = sin Ψ(x, 0) dx .
a 0 a

|cn |2 is the probability that a measurement of the energy would yield the value En . The sum

of these probabilities should be 1.



X
|cn |2 = 1
n=1

This follows from the normalization of Ψ(x, t), use eq. 11.

Proof.

Z ∞
|Ψ(x, 0)|2 dx = 1
−∞

!∗ ∞
!
Z ∞ X X
or cm ψm (x) cn ψn (x) dx = 1
−∞ m=1 n=1
X∞ X
∞ Z ∞
or c∗m cn ψm∗
(x)ψn (x) dx =1
m=1 n=1 −∞
∞ X
X ∞
or c∗m cn δmn = 1
m=1 n=1

X
or |cn |2 = 1
n=1

The expectation value of the energy must be


X
hHi = |cn |2 En
n=1

Proof.

Ĥψn = En ψn ← time independent Schrödinger equation


Z ∞
hHi = Ψ∗ (x, 0)ĤΨ(x, 0) dx
−∞
Z ∞ X ∞
!∗ ∞
!
X
= cm ψm (x) Ĥ cn ψn (x) dx
−∞ m=1 n=1

XX ∞ Z ∞
= c∗m cn En ∗
ψm (x)ψn (x) dx
m=1 n=1 −∞
18

∞ X
X ∞
= c∗m cn En δmn ← orthonormality condition
m=1 n=1
X∞
∴ hHi = |cn |2 En
n=1

The probability of getting a particular energy En which is |cn |2 , is independent of time and so
is the expectation value of H. This is a manifestation of conservation of energy in quantum

mechanics.

The harmonic oscillator

The paradigm for a classical harmonic oscillator is a mass m attached to a spring of force

constant k. The motion is governed by Hooke’s law,

d2 x
F = −kx = m .
dt2

The solution is

x(t) = A sin(ωt) + B cos(ωt)


r
k
where ω = is the angular frequency of oscillation.
m

The potential energy is

1 ∂V
V (x) = kx2 ∵F =−
2 ∂x

The quantum problem is to solve the Schrödinger equation for the potential

1
V (x) = mω 2 x2 .
2

Solve:
~2 d2 ψ 1
− + mω 2 x2 ψ = Eψ (12)
2m dx2 2
19

V(x)

Figure 4: A particle in a classical harmonic oscillator potential can have any energy E > 0.

The Schrödinger equation can be rewritten as

1  2 ~ d
p + (mωx)2 ψ = Eψ where p ≡

2m i dx
1  2
p + (mωx)2

∴H=
2m

We can solve the differential equation using the power series method or the ladder operator
method. We choose the ladder operator method. Let us examine the quantities

1
a± ≡ √ (∓ip + mωx)
2m~ω
1
Then, a− a+ = (ip + mωx)(−ip + mωx)
2~mω
1  2
p + ipmωx − imωxp + (mωx)2

=
2~mω
1  2
p + (mωx)2 − imω(xp − px)

=
2~mω

The quantity (xp − px) is called the commutator of x and p (note that x and p are not numbers
but operators, operators need not necessarily commute). In general,

[A, B] = AB − BA

In this notation,
1  2 i
p + (mωx)2 − [x, p]

a− a+ =
2~mω 2~
20

Evaluation of [x, p]:

 
~ df ~ d
[x, p]f (x) = x − (xf )
i dx i dx
 
~ df df
= x −x −f
i dx dx
= i~f (x)

∴ [x, p] = i~ ← canonical commutation relation

H 1 1  2
p + (mωx)2 ∧ [x, p] = i~

Therefore, a− a+ = +
∵H=
~ω  2  2m
1
or H = ~ω a− a+ −
2
1
Now, a+ a− = (−ip + mωx)(ip + mωx)
2~mω
1  2
p − imωpx + imωxp + (mωx)2

=
2~mω
1  2
p + (mωx)2 + imω(xp − px)

=
2~mω
H 1 1  2
p + (mωx)2 ∧ [x, p] = i~

= − ∵H=
~ω 2 2m
∴ [a− , a+ ] = 1

H can also be written as


 
1
H = ~ω a+ a− +
2

In terms of a± , the Schrödinger equation takes the form

 
1
~ω a± a∓ ± ψ = Eψ
2

We state that the wave function a+ ψ satisfies the Schrödinger equation with energy E + ~ω,

that is
H(a+ ψ) = (E + ~ω) (a+ ψ)

Proof.

 
1
H(a+ ψ) = ~ω a+ a− + (a+ ψ)
2
21
 
1
= ~ω a+ a− a+ + a+ ψ
2
 
1
= ~ωa+ a− a+ + ψ
2
  
1
= a+ ~ω a+ a− + 1 + ψ ∵ [a− , a+ ] = 1
2
 
1
= a+ (H + ~ω) ψ ∵ H = ~ω a+ a− +
2
= a+ (E + ~ω)ψ

= (E + ~ω)(a+ ψ)

Similarly, the wave function a− ψ satisfies the Schrödinger equation with energy E − ~ω, that is

H(a− ψ) = (E − ~ω) (a− ψ)

Proof.

 
1
H(a− ψ) = ~ω a− a+ − (a− ψ)
2
 
1
= ~ω a− a+ a− − a− ψ
2
 
1
= ~ωa− a+ a− − ψ
2
  
1
= a− ~ω a− a+ − 1 − ψ ∵ [a− , a+ ] = 1
2
 
1
= a− (H − ~ω) ψ ∵ H = ~ω a− a+ −
2
= a− (E − ~ω)ψ

= (E − ~ω)(a− ψ)

a± are called ladder operators because they allow us to climb up and down in energy.

a+ −→ raising operator
a− −→ lowering operator

a− ψ is a new solution to the Schrödinger equation, but there is no guarantee that it will be
normalizable – it might be zero or its square integral might be infinite. There occurs a lowest
22

state ψ0 such that


a− ψ0 = 0

We will use the above equation to determine ψ0 (x)

a− ψ0 = 0
1
or √ (ip + mωx)ψ0 =0
2m~ω
 
1 d ~ d
or √ ~ + mωx ψ0 =0 ∵p=
2m~ω dx i dx
dψ0 mωx
or =− ψ0
Z dx ~ Z
dψ0 mω
⇒ =− x dx
ψ0 ~
mω 2
⇒ ln ψ0 = − x +k
2~
2
⇒ ψ0 (x) = Ae− 2~ x

Normalize ψ0 (x) :
Z ∞
|ψ0 (x)|2 = 1
Z ∞ −∞
2
⇒ |A|2 e−mωx /~ dx = 1
Z−∞
∞ Z ∞ Z ∞
2
⇒ 2|A|2 e−mωx /~ dx = 1 ∵ f (x) dx = 2 f (x) dx , if f (−x) = f (x)
0 −∞ 0
mωx2 dz 2mωx
Let = z, =
~ r dx ~
Z ∞
1 ~
⇒ 2|A|2 e−z dz =1
0 2 mωz
r Z ∞
~
⇒ |A| 2
e−z z −1/2 dz =1
mω 0
r
∞ √
1
Z
π~
⇒ |A| 2
=1 ∵ e−z z −1/2 dz = Γ( ) = π
mω 0 2
r

∴ |A|2 =
π~
 mω 1/4 mω 2
∴ ψ0 (x) = e− 2~ x
  π~
1
Now, ~ω a+ a− + ψ0 (x) = E0 ψ0 (x)
2
Use, a− ψ0 (x) = 0
1
E0 = ~ω (ground state energy)
2
23
 
n 1
ψn (x) = An (a+ ) ψ0 (x) with En = n + ~ω
2
where ψn (x) is the nth excited state and An is the normalization constant.

A particle in a potential can either be in a bound state or a scattering state depending on the

energy of the particle and the potential function. Some potentials admit only bound states,
some allow only scattering states, while some permit both kinds depending on the energy of
the particle. In quantum mechanics, the solutions to the Schrödinger equation correspond to
either bound states or scattering states. If


E < V (−∞) and V (+∞) ⇒ bound state,


E > V (−∞) or V (+∞) ⇒ scattering state,

In real life most potentials go to zero at infinity, in which case the criterion simplifies even

further 

E < 0 ⇒ bound state,


E > 0 ⇒ scattering state,

We will now study a particle in an attractive delta function potential. The Dirac delta function,
δ(x), is defined as
 
 
0,
 if x 6= 0 
 Z +∞
δ(x) = , with δ(x) dx = 1.
  −∞
∞, if x = 0 
 

It is an infinitely high, infinitesimally narrow spike at the origin, whose area is 1. Note that
δ(x − a) is a spike of area 1 at the point a. Note some properties of the delta function

f (x)δ(x − a) = f (a)δ(x − a)

because the product is zero for all x except at x = a. Also,

Z +∞ Z +∞
f (x)δ(x − a) = f (a) δ(x − a) dx = f (a)
−∞ −∞
24

Consider a potential of the form

V (x) = −αδ(x), where α is some constant

Considering this potential, the Schrödinger equation reads

~2 d2 ψ
− − αδ(x)ψ = Eψ.
2m dx2

The solutions are both bound states (E < 0) and scattering states (E > 0). First, let us look
at bound states. In the region x < 0, V (x) = 0, so

d2 ψ 2mE
2
= − 2 ψ = κ2 ψ,
dx ~ √
−2mE
where κ ≡
~

E is negative by assumption, so κ is real and positive. The general solution to the above
equation is

ψ(x) = Ae−κx + Beκx ,

but the first term blows up as x → −∞, so A = 0, therefore

ψ(x) = Beκx , (x < 0).

In the region x > 0, V (x) is again zero and the general solution is of the form

ψ(x) = F e−κx + Geκx ,

now the second term blows up as x → +∞, so G = 0, and

ψ(x) = F e−κx , (x > 0).

Recall the standard boundary conditions for ψ,


25

1. ψ is always continuous,


2. is continuous except at points where the potential is infinite.
dx

In this case, the first boundary condition implies F = B, so



Beκx ,

 (x ≤ 0),
ψ(x) =
Be−κx , (x ≥ 0).

V is infinite at x = 0 and one can clearly see that the function ψ(x) has a kink at x = 0. Let us
calculate the discontinuity in the derivative of ψ at x = 0. Integrate the Schrödinger equation
from −ǫ to +ǫ and then take the limit ǫ → 0.

+ǫ +ǫ +ǫ
~2 d2 ψ
Z Z Z
− dx + V (x)ψ(x) dx = E ψ(x) dx
2m −ǫ dx2 −ǫ −ǫ


The first integral is evaluated between the two end points, the last integral is zero in the
dx
limit ǫ → 0, as it is the area of a strip with vanishing width and finite height. Therefore,

  +ǫ
dψ 2m
Z
∆ = 2 lim V (x)ψ(x) dx .
dx ~ ǫ→0 −ǫ

When V (x) is not infinite, the integral on the right hand side is zero. In this case, V (x) is

infinite at x = 0. Using the property of the delta function under an integral,

 
dψ 2mα
∆ =− ψ(0). (13)
dx ~2

In this case, 

= −Bκe−κx , dψ


 for (x > 0), so dx
= −Bκ
dx

+

= Bκeκx , dψ


 for (x < 0), so dx
= Bκ
dx


 

Therefore, ∆ = −2Bκ and ψ(0) = B. So, from Eq. 13 we get,
dx


κ= ,
~2
26

and the allowed energy is

~2 κ2 mα2
E=− =− 2 .
2m 2~

Now let us normalize ψ

+∞ +∞
|B|2
Z Z
2
|ψ(x)| dx = 2|B| 2
e−2κx dx = = 1,
−∞ 0 κ

√ mα
so B = κ= .
~

The delta function well has exactly one bound state. To summarize,


mα −mα|x|/~2 mα2
ψ(x) = e , E=− .
~ 2~2

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy