Sei sulla pagina 1di 54

Quantum Mechanics

Sagar J C
St. Joseph’s College, Bengaluru

19PCM21015
3rd year Bachelor of Science

( Physics, Chemistry and Mathematics )

2nd October 2021


Contents

1 Transition from Classical Mechanics to Quantum Mechanics 1


1.1 The atomic structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Black-body radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Photoelectric effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.4 Bohr’s Planetary model of atom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.5 Matter waves and De Broglie’s Hypothesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2 Mathematical Basis for Quantum Mechanics 7


2.1 Vector space and Hilbert space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Matrices and Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 Eigen Value problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.4 Generalized Uncertainty Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.5 Function space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3 Wave Function and Quantum mechanics 15


3.1 Wave Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2 Postulates of Quantum Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3 Schrodinger’s wave equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.4 Operators in Quantum Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.5 Measurements in quantum Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.5.1 Wave Function Collapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.5.2 Quantum Indeterminacy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.5.3 Expectation Values of Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.6 Heisenberg’s Uncertainty Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.6.1 Position-Momentum uncertainty . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.6.2 Schrodinger’s equation in momentum space . . . . . . . . . . . . . . . . . . . . . . . 23
3.6.3 Energy-Time uncertainty . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.7 Time Independent Schrodinger’s Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.8 Stationary states in quantum mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

4 Application of Schrodinger’s Theory 29


4.1 Infinite square well . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2 Quantum Harmonic Oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.2.1 Quantum Harmonic Oscillator - Analytical Method . . . . . . . . . . . . . . . . . . . 34
4.3 The Hydrogen Atom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

5 Properties of particles in 3D 45
5.1 Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.2 Spin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.2.1 Spin 1/2 Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.2.2 Half Integral spin (n/2) and Integral spin (n) Particles . . . . . . . . . . . . . . . . . 51

Bibliography 52

i
Chapter 1

Transition from Classical Mechanics to


Quantum Mechanics
Quantum mechanics was not born all of a sudden. It was a gradual process where the efforts from physicists
over decades gave birth to Quantum Mechanics. The transition from classical mechanics to quantum was
by physicists who started to postulate the motion and behaviour of microscopic particles of very small mass
and size. Let us go through that journey in this particular chapter.

1.1 The atomic structure


The very birth of Quantum mechanics was because of the interest of man to understand the structure of
atom starting from Dalton’s atomic theory where he postulated the properties of atoms.
1. Dalton’s atomic theory
1. All matter is made of atoms. Atoms are indivisible and indestructible.
2. All atoms of a given element are identical in mass and properties
3. Compounds are formed by a combination of two or more different kinds of atoms.
4. A chemical reaction is a rearrangement of atoms.
2. J.J Thomson’s Plum Pudding model
But by the experiments done by J.J Thomson namely cathode ray experiments(proving the existence of
electrons), anode ray experiments (proving positively charged ions), etc. He broke one of the postulates
of Dalton’s theory where he showed the atom is composed of negatively charged hard particles called as
electrons present in the sea of positively charged bodies in his Plum-Pudding model of atom. He also found
out the value of charge to the mass ratio of electron to be 1.7588 × 1011 C Kg−1
3. Rutherford’s Nuclear model of atoms
In his famous Gold Foil experiment, Rutherford showed that 99.9% of atom is empty where the center
consists of a very tiny dense +ve nucleus which is surrounded by electrons revolving around the nucleus.

1
He showed that the electron was balanced by the inward coulomb force and outward centrifugal force.
Hence Fe = Fc .
1 Ze2 me v 2 Ze2 1 Ze2
⇒ = ⇒ v2 = ⇒ K.E. = me v 2 = (1.1)
4πε0 r2 r 4πε0 rme 2 8πε0 r
P.E Ze2
F =− ⇒ P.E. = − (1.2)
r 4πε0 r
Ze2 Ze2 Ze2
Total Energy = K.E + P.E. = − ⇒ E=− (1.3)
8πε0 r 4πε0 r 8πε0 r
Thus he proved that the total energy of the atom is negative implying that the system is bounded and to
remove electrons, energy has to be supplied where ‘v’ is the velocity of electron, ‘e’ is the charge of electron,
‘me ’ is the mass of electron, ‘Z’ is the total number of electrons, ‘r’ is the radius of the path electrons
revolve, ‘ε0 ’ is the permittivity of free space. This model was a major success until the absorption and
emission spectral lines of gaseous elements were observed. Only certain definite frequencies of light were
emitted or absorbed for a particular element. Rutherford couldn’t give an explanation for that and some
more phenomena which are discussed below, thus disproving this model.

1.2 Black-body radiation


We know that any object when heated emits electromagnetic waves. So when the graph was plotted with
the wavelength λ of the electromagnetic wave on x-axis and the energy radiated by the object per unit
surface area per unit second (Eλ ) on the y-axis the graph was of this sort.

Here we see that the energy radiated by the object per unit surface area per unit second for a particular
wavelength emitted by object increases if the wavelength increases only up to a particular point called as
wavelength corresponding to peak radiation energy (λmax ) and then for wavelengths higher than that, the
radiation energy decreased. This is known as the black body phenomenon. Many scientists tried explaining
this phenomenon by the classical picture of light which is made up of electromagnetic waves which were
continuous and taking the body to be made of classical particles. But that was not working and gave rise
to faulty equations. But in the 1900s, Max Plank put forth his quantum theory of light and told that light
is made up of minute discrete packets of energy called ‘Photons’. And absorption or emission of light will
take place in an integral multiple of the energy of photon whose energy is,

E = hν (1.4)

When Plank formulated the phenomenon of the Black body in the quantum picture of light he got a
distribution called as ‘Plank’s Distribution curve’. By plank’s distribution curve for black bodies, the total
radiation energy density (dEλ ) within the range (dλ) at a particular wavelength (λ), emitted by a black
body of a particular temperature (T ) per unit wavelength, per surface area, per unit time is given by,

2πhc2 1
Eλ = 5 hc (1.5)
λ e λkB T − 1

2
1.3 Photoelectric effect
The photoelectric effect is a phenomenon where if an electromagnetic wave of certain radiation is incident
on a metal surface the electrons are ejected from the surface. If the setup is used in a circuit then the
ejected electrons called photoelectrons can travel and complete the circuit. This is called as Photoelectric
effect.

Characteristics of photoelectric effect


1. The emission of photoelectrons was instantaneous (∼ 10−9 s) and occurred only if the incident radiation
had frequency greater than a particular called ’The threshold frequency’, which depended on the metal
used.
2. Unless the incident light is higher than the threshold frequency no matter how intense the incident
light is, no photoelectrons were emitted.
3. Increasing the intensity of incident light, increased the photo-current (number of photoelectrons pro-
duced). So photo-current is directly proportional to the intensity. But increasing the intensity doesn’t
affect the kinetic energy of the photoelectrons.
4. Even though the incident light has very low intensity, if the frequency was higher than the threshold
frequency, the photoelectrons were ejected instantaneously with high kinetic energy except that the
number of photoelectrons ejected is less. Thus we can tell that the kinetic energy of photoelectrons
is directly dependent on frequency after threshold frequency.

metal - 1 metal - 2

KEmax

photo
electrons

ν01 ν02
0 ν 0 Intensity of light

Einstein’s Postulates for photoelectric effect


Einstein used plank’s theory of light to explain the photoelectric effect. Since light is made of discrete
packets of energy whose value is E = hν, the absorption of light energy by atoms will be in discrete
quantities. Let the total energy supplied by a photon be E = hν. This total energy provided by the photon
is used up in removing the electron from the surface by providing a certain amount of energy to the atom
called the work function (φ) which is related to the threshold frequency φ = hν0 and then the excess energy
is used in providing kinetic energy to the electron ( 12 me v 2 ).
1
E = Work function + K.E. ⇒ hν = φ + me v 2 (1.6)
2
1
⇒ hν = hν0 + me v 2 (1.7)
2
If we apply a reverse potential in the plates so that the photoelectron is emitted from the anode and
reaches the cathode, the photo-current decreases and at a particular negative potential the photocurrent
becomes Zero, which is called as the ’Stopping potential’ (V0 ) which is equal to the kinetic energy of the
1 h φ
photoelectron eV0 = me v 2 . Then the above equation becomes hν − φ = eV0 or V0 = ν−
2 e e

3
1.4 Bohr’s Planetary model of atom
Bohr’s Postulates
1. Electrons in the atom revolve around certain definite orbits where they don’t emit any energy as long
as they are in these stable orbits and have definite energy called as stationary states denoted by the
quantum number n = 1, 2, 3, . . . .

2. Electrons revolve around the nucleus only in those orbits where the ‘Angular momentum’ of the
h nh
electron is an integral multiple of . Thus he told that L = where h is the plank’s constant.
2π 2π
3. The electrons can go to higher energy states only by absorbing the light whose quanta corresponds
to the difference in the transiting energy levels, i.e., if an electron goes from n = 1 to n = 2 energy
level whose energy corresponds to E1 and E2 respectively, then the electron absorbs the light whose
quanta has energy E = hν = E2 − E1 . And if the electron goes back to the initial state then it will
emit a photon whose energy is exactly equal to the difference in the energy levels.

n=2
excited
electron
n=1
ground
electron

nucleus hν

Bohr’s radius
nh h 2h 3h 4h
By Bohr’s 2nd Postulate, the angular momentum of electron can be L = = , , , . . . only.
2π 2π 2π 2π 2π
nh nh
Thus the angular momentum is quantized. And is, L = me vr = ⇒ v= (1.8)
2π 2πrme
The inward electrostatic force will be balanced by the outward centrifugal force (Fe = Fc ), so we get,

1 Ze2 me v 2 Ze2
= ⇒ v2 = (1.9)
4πε0 r2 r 4πε0 rme
Now squaring Eq.(1.8) and using that value in the above equation we get,

n2 h2 Ze2 n2 h2 ε0 n2 h2 ε0 n2 a0 h 2 ε0
= ⇒ rn = = × = a0 = (1.10)
4π 2 r2 m2e 4πε0 rme Ze2 πme Z πe2 me Z πe2 me

If the atom has Z = 1 and if n = 1 then, r1 = a0 = 5.29177 × 10−11 m is called as Bohr’s radius.

Energy levels of an atom


Using the expression for energy in Eq.(1.3) and substituting the value of rn in that equation we get,

Ze2 Ze2 πme Z 2 me e4


En = − × 2 2 ⇒ En = − (1.11)
8πε0 n h ε0 8ε20 h2 n2

Z2 −18 13.605 Z 2
En = × 2.1788 × 10 Joules = eV (1.12)
n2 n2

4
E5 n=5
P f und Series
E4 n=4
Brackett Series
E3 P aschen Series
n=3

E2 n=2
Balmer Series

En
E1 n=1
Lyman Series

By this formula for energy levels, Bohr was able to explain the absorption and the emission spectrum of
the Hydrogen atom (Z = 1) successfully. So if electron transit in the atom then the difference in the final
energy level Ef and initial energy level Ei is given by,

me e4 me e4
Ef = − and Ei = − (1.13)
8ε20 h2 n2f 8ε20 h2 n2i
!
me e4 1 1
⇒ ∆E = Ef − Ei = − 2 (1.14)
8ε20 h2 n2f ni
In the transition of electron it absorbs or emits a photon whose energy will be ∆E = hν which will be equal
to the difference in the two energy levels,
hc
∆E = Ef − Ei = hν = = hcν (1.15)
λ
!
me e4 1 1
⇒ hcν = 2 2 − 2 (1.16)
8ε0 h n2f ni
!
me e4 1 1
⇒ ν= 2 3 − 2 (1.17)
8ε0 h c n2f ni
me e4 me e4
Then the factor is called as the Rydberg’s constant, denoted by R∞ = = 101973731.6 m−2
8ε20 h3 c 8ε20 h3 c
In the Hydrogen atom, the transition of an electron from various energy level to the ground state nf = 1,
corresponds to Lyman series which is in UV region. And if nf = 2 then it is to Balmer series which is in
the Visible Region. And for nf = 3, 4, 5 it is Paschen, Brackett, and Pfund series which lies in IR region.

Similarly for other Hydrogen like species, He+ , Li2+ , Be3+ , etc, the results shown by Bohr’s model was
accurate where the Rydberg’s constant was modified to,
Z 2 me e4
R= = Z 2 × 101973731.6 m−2 (1.18)
8ε20 h3 c
Thus this model was explaining the structure of only Hydrogen and Hydrogen like atoms. But if Multi-
electron atoms were taken into account, then the model failed completely. Mainly because it couldn’t
account for the coulombic force between the electrons. Also because of many other new hypotheses and
principles, Bohr’s model was slowly discarded.

5
1.5 Matter waves and De Broglie’s Hypothesis
We know by Double slit experiments that light behaves as a pure wave. But in the phenomenon of Black-
body radiation and Photoelectric effect, light behaves as a particle. Thus we can tell that Light or generally
Electromagnetic radiation behaves as both waves and as particles. This is the Dual nature of light. Then
how can we interpret light, can we actually see light and its quanta? No, we can’t see them rather we can
feel them only through their effects. Any light we see is the excitation of the rods and cones cells in our
retina which converts the excitation to electric signals and our brain interprets them accordingly and thus
we can see the light that too only if the wavelength is in the visible range.

So if electromagnetic waves show dual nature, will the matter particles show dual nature as both particles
and waves? This is De’ Broglie’s Hypothesis. He hypothesized that even matter particles show wave nature.
And showed that the wavelength of the matter particles relates with its momentum given by,
h h h 2π
λ= ⇒ P = = · = }k (1.19)
P λ 2π λ
If the particle is electron and is accelerated by a potential difference of V then the kinetic energy of the
electron will be equal to the charge times potential difference,
1 (me v)2 P2
K.E. = eV = me v 2 = = (1.20)
2 2me 2me
P2 h2 h2 h
eV = = ⇒ λ2 = ⇒ λ= √ (1.21)
2me 2me λ2 2me eV 2me eV

Davisson and Germer Experiment


This experiment verified the wave nature of electrons. In the experiment, an electron gun fueled by a high
voltage battery emitted fast-moving electrons. These electrons were collimated by a narrow hole and were
made incident on a Nickel crystal. When the electrons were scattered back it was detected by a detector
that could move in a circular arc-shaped trajectory. In the experiment, the position where the maximum
number of electrons were detected corresponded to the angle of 50◦ , when a voltage of 54V was applied, then
when the detector was moved in a circular path, there were regions where the detector detected maximum
and minimum number of electrons, this was similar to the maxima and minima of the diffraction pattern.
But diffraction is purely a wave phenomenon, thus it was confirmed that the electron behaved as a wave in
this experiment. When the wavelength of the electron was measured by the above formula. It was,
h 6.626 × 10−34
λ= √ =√ = 1.66 × 10−10 m = 0.166 nm (1.22)
2me eV 2 × 9.1 × 10−31 × 1.6 × 10−19 × 54
When the wavelength was calculated by Bragg’s law for the same parameters, it was 0.165 nm which is
close to the experimental value and hence proving the wave nature of electron.

G.P. Thomson’s Experiment


G.P Thomson’s experiment was also an electron diffraction experiment. When the cathode rays (stream
of electrons) generated by an induction coil were directed to a thin metallic film of gold, the rays passed
through the foil and the transmitted rays were collected in a photographic plate. The photographic plate
showed thin circular alternate black and white rings which were due to the circular diffraction of electrons
through the foil. When the calculations were done based on the distance from foil and the photographic
plate, radii of diffracting rings, the wavelength of electron (calculated using the potential difference), the
slit width (equal to the length of the unit cell of the gold crystal) was found out experimentally, which
matched with the theoretical value and thus proved the wave nature of electron through diffraction.

Thus Bohr’s model of atom failed to incorporate the dual nature of matter particles (electron) and multi-
electron atoms. And hence physicists built a radically new theory to explain the behavior of quantum
particles through ‘Wave Quantum Mechanics’. But before diving into that let us revise some mathematics.

6
Chapter 2

Mathematical Basis for Quantum Mechanics

2.1 Vector space and Hilbert space


Vector Spaces
A vector space V is a space whose elements are denoted by |V i and are of the form |V1 i, |V2 i, |V3 i, |V4 i,. . . .
and also has associated scalars a1 , a2 , a3 , a4 , . . . (real or complex). A vector |V i is represented as,
 
v1
v2 
|V i = v  (2.1)
 
 3
..
.
The necessary and sufficient conditions for a set of vectors to be a vector space is,
1. Closure property :- ∀ |V1 i, |V2 i ∈ V, ∃ |V3 i ∈ V such that |V1 i + |V2 i = |V3 i
2. Associative property:- ∀ |V1 i, |V2 i, |V3 i ∈ V, (|V1 i + |V2 i) + |V3 i = |V1 i + (|V2 i + |V3 i)
3. Commutative property:- ∀ |V1 i, |V2 i ∈ V, |V1 i + |V2 i = |V2 i + |V1 i
4. Has a unique additive identity:- ∀ |V i ∈ V, ∃ |0i ∈ V such that |V i + |0i = |0i + |V i
5. Has additive inverse:- ∀ |V i ∈ V, ∃ |V i−1 ∈ V such that |V i + |V i−1 = |V i−1 + |V i = 0
6. Scalar Multiplication:- ∀ |V i ∈ V, a|V i = |V ia ∈ V where a is some scalar
7. Distributive properties:- a1 (|V1 i + |V2 i) = a1 |V1 i + a1 |V2 i and |V1 i(a1 + a2 ) = a1 |V1 i + a2 |V1 i
8. Associative property of Scalar Multiplication:- ∀ |V i ∈ V, a1 (a2 |V i) = (a1 a2 )|V i
9. Identity of Scalar multiplication:- ∀ |V i ∈ V, ∃ a scalar 1 such that 1(|V i) = |V i
10. Formation of a null vector:- ∀ |V i ∈ V, ∃ a scalar 0 such that 0(|V i) = |0i
Note that the vector spaces need not have only vectors (quantity with magnitude and direction), it can also
have ‘Matrices’ and even ‘Functions’ in it. Since they have all the above-mentioned properties.

Hilbert Spaces
A Hilbert space H is a subspace of a vector space V where all elements follow the rules of vector space and
also in addition to them, the elements have a new operation called the Inner product and new rules. Let
us denote the elements of H by Ψ1 , Ψ2 , Ψ3 , . . . We know how a vector |Ψ i can be represented as in Eq.(2.1).
Now let us define another type of the same vector called as the ‘Dual’ of |Ψ i denoted by hΨ | where,
hΨ | = Conjugate transpose of |Ψ i = complex conjugate of |Ψ i = v1∗ v2∗ v3∗ . . . (2.2)
 

Where v ∗ is the conjugate of v and is got by replacing the imaginary number i by −i wherever it appears.

7
Properties of elements in a Hilbert space
1. Inner product operation:- The inner product of hΦ| and |Ψ i is defined as,
 
v1
 v2 

 X ∗
hΦ|Ψ i = w1∗ w2∗ w3∗ . . . v  = wi vi = Scalar quantity ∈ C (2.3)

 3
.. i
.
2. Conjugate symmetry (skew-symmetry):- hΦ|Ψ i = hΨ |Φi∗

3. Scalar multiplication is linear with respect to the second vector:- hΦ|aΨ i = ahΦ|Ψ i

4. Scalar multiplication is conjugate linear with respect to the first vector:- haΦ|Ψ i = a∗ hΦ|Ψ i

5. Linear with respect to second vector:- hΨ1 |aΨ2 + bΨ3 i = hΨ1 |aΨ2 i + hΨ1 |bΨ3 i = ahΨ1 |Ψ2 i + bhΨ1 |Ψ3 i

6. Anti-linear with respect to 1st vector:- haΨ1 + bΨ2 |Ψ3 i = haΨ1 |Ψ3 i + hbΨ2 |Ψ3 i = a∗ hΨ1 |Ψ3 i + b∗ hΨ2 |Ψ3 i

7. Positive definiteness - Norm is greater than or equal to zero:- hΨ |Ψ i = kΨ k2 ≥ 0 → kΨ k ≥ 0

8. Distance between two vectors in Hilbert space:- kΨ2 − Ψ1 k =


p
hΨ2 − Ψ1 |Ψ2 − Ψ1 i

9. Hilbert spaces are separable, i.e., they contain countable, dense subsets.

10. Hilbert spaces are complete, i.e., there are no gaps in between the elements (continuous), so every
element has an adjacent element that is very close such that the distance between the two elements limits
to zero. So we get lim kΨm − Ψn k = 0
n→m

Dirac Notations
We have seen the vectors and their conjugate transpose represented as |Ψ i and hΨ |. This is Dirac’s notation.
 
v1
 v2 
 
|Ψ i =  v3  hΨ | = (|Ψ i∗ )T = v1∗ v2∗ v3∗ . . . vn∗ (2.4)
   
 .. 
.
vn
So we have represented so many properties of a Vector and Hilbert spaces in Dirac notations already.

1. Triangle Inequality:- (equal only when Φ = Ψ = 0)


p p p
hΦ + Ψ |Φ + Ψ i ≤ hΦ|Φi + hΨ |Ψ i

2. Schwartz inequality:- hΦ|Ψ i2 ≤ hΦ|ΦihΨ |Ψ i ⇒ khΦ|Ψ ik ≤ kΦk kΨ k

3. Orthogonal vectors:- hΦ|Ψ i = 0

4. Normalized vector:- Ψ is normalized if hΨ |Ψ i = 1

5. Ortho-Normal vectors:- Two vectors are ortho-normal if hΨ1 |Ψ1 i = 1 and hΨ2 |Ψ2 i = 1 and ( hΨ1 |Ψ2 i = 0
1 , i=j
which can be written shortly as hΨi |Ψj i = δij where δij is the Kronecker delta given by δij =
0 , i 6= j
6. Any vector can be expressed inXterms of a basis vector
X whose norm is 1 and also as a linear combination
of normalized vectors. So, |Ψ i = vi |ii and hΨ | = vi∗ hi|
i i

8
2.2 Matrices and Operators
An operator  is a mathematical tool that can transform a vector |Ψ i to |Ψ 0 i given by, Â|Ψ i = |Ψ 0 i where
|Ψ i, |Ψ 0 i ∈ H. The operator  is more like a square matrix. In n-dimensional vector space, we can write,
    
a11 a12 a13 . . . a1n v1 Σa1i vi
 a21 a22 a23 . . . a2n   v2   Σa2i vi 
    
Â|Ψ i =  a31 a32 a33 . . . a3n   v3  =  Σa3i vi  = |Ψ 0 i
    
 .. .. .. .. ..   ..   .. 
 . . . . .  .   . 
an1 an2 an3 . . . ann vn Σani vi

Similarly the operator  can also act on the dual space of hΨ | where we get a transformed dual vector hΨ 0 |
but we have to operate  on the right side so as to be consistent with matrix multiplication,
 
a11 a12 a13 . . . a1n
 a21 a22 a23 . . . a2n 
 
 a31 a32 a33 . . . a3n   ∗
 = Σvi ai1 Σvi∗ ai2 Σvi∗ ai3 . . . = hΨ 0 |
 
hΨ |Â = v∗1 v∗2 v∗3 . . . v∗n 
 .. .. .. .. .. 
 . . . . . 
an1 an2 an3 . . . ann
These operators have some significant implications in quantum mechanics because they are used to get the
observable values of some physical quantities like position of a particle, its momentum, total energy, etc

Properties of Operators
1. All operators don’t commute under multiplication:- ÂB̂ 6= B̂Â
2. Operators are associative:- Â(B̂Ĉ) = (ÂB̂)Ĉ
3. Power property:- (Â)n (Â)m = (Â)n+m
4. hΦ|Â|Ψ i ∈ C is a scalar. Because Â|Ψ i = |Ψ 0 i and hΦ|Â|Ψ i = hΦ|Ψ 0 i ∈ C (inner product property)
5. Identity operator Î is an operator which doesn’t transform the vector. Î|Ψ i = Ψ and hΨ |Î = hΨ |
6. Â−1 is that operator which reverses the transformation done by Â, Hence ÂÂ−1 = Â−1 Â = Î
7. (ÂB̂)−1 = B̂−1 Â−1 only then we get (ÂB̂)(ÂB̂)−1 = (ÂB̂)B̂−1 Â−1 = Â(

 −1 )Â−1 = ÂÂ−1 = Î

8. (Ân )−1 = (Â−1 )n

hΨ |Â|Ψ i
9. The expectation value of  is given by
hΨ |Ψ i
10. The outer product of two vectors given by |Ψ ihΦ| is also an operator. Because this quantity gives rise
to an n × n matrix which is an operator as we know in n-dimensional space.

Linear Operators
Linear Operators are a special kind of operators and are extensively used in Quantum mechanics. For
operator  to be a linear operator, they have to follow these rules in addition to the above properties of a
normal operator. If Ψ ∈ H and a ∈ C then,
1. Â(|Ψ1 i + |Ψ2 i) = Â|Ψ1 i + Â|Ψ2 i and Â|aΨ i = aÂ|Ψ i
2. (hΨ1 | + hΨ2 |) = hΨ1 | + hΨ2 | and hΨ |a = ahΨ |Â
3. Â|Ψ i = |ÂΨ i = |Ψ 0 i
4. (ÂB̂)|Ψ i = Â(B̂|Ψ i) = Â|B̂Ψ i

9
Hermitian Conjugate Operation {}†
The Hermitian conjugate of a complex number a is equal to its complex conjugate. a† = a∗ . And for
a vector |Ψ i its hermitian conjugate denoted by |Ψ i† is defined as its dual vector. So, |Ψ i† = hΨ | and
similarly hΨ |† = |Ψ i. But for an operator, we can understand the hermitian conjugate in the following way.
If Â|Ψ i = |Ψ 0 i, then the operator that has to operate on hΨ | such that it transforms it to hΨ 0 | is called as
the Hermitian conjugate of  and is denoted by, † . So we can say to transform a |Ψ i we use  where
Â|Ψ i = |ÂΨ i = |Ψ 0 i and to transform a hΨ | we use † where, hΨ |† = hÂΨ | = hΨ 0 | . Thus we get,
 ∗
a11 a∗21 a∗31 . . . a∗n1
  
a11 a12 a13 . . . a1n
 a21 a22 a23 . . . a2n   a∗ a∗ a∗ . . . a∗ 
   12 22 32 n2 
 ∗ ∗ ∗ ∗ 
If  =   then † = (Â| )∗ =  a13 a23 a33 . . . an3 
 a31 a32 a33 . . . a3n 
(2.5)
 .. .. .. . . ..   .. .. .. . . .. 
 . . . . .   . . . . . 
an1 an2 an3 . . . ann a∗1n a∗2n a∗3n . . . a∗nn

|aÂΨ i = a∗ hΨ |† (2.6)



hΦ|ÂΨ i = h Φ|Ψ i ∀ Ψ, Φ ∈ H (2.7)
† ∗
hΦ|Â|Ψ i = hΨ |Â |Φi (2.8)

Properties of Hermitian Conjugate Operation


1. († )† = Â

2. (aÂ)† = a† †

3. († )n = (Ân )†

4. ( + B̂ + Ĉ)† = † + B̂† + Ĉ†

5. (ÂB̂Ĉ)† = Ĉ† B̂† † ⇒ (ÂB̂Ĉ|Ψ i)† = hΨ |Ĉ† B̂† †

6. (|ΦihΨ |)† = |Ψ ihΦ|

Commutator and Anti-Commutator brackets


Let  and B̂ be two operators, then their commutator is [Â, B̂] = ÂB̂ − B̂ and their anti-commutator
is {Â, B̂} = ÂB̂ + B̂Â . Some properties of commutator brackets are:-

1. Commutator of the same operator is Zero:- [Â, Â] = 0

2. Operators commute with scalars:- [Â, a] = a − Âa = 0

3. Commutator is anti-symmetric:- [Â, B̂] = −[B̂, Â]

4. Commutator is linear:- [Â, B̂ + Ĉ + . . . ] = [Â, B̂] + [Â, Ĉ] + . . .

5. Hermitation conjugate of commutator reverses the order of operators:- [Â, B̂]† = [B̂† , † ]

6. Commutator is distributive with respect to second operator:- [Â, B̂Ĉ] = [Â, B̂]Ĉ + B̂[Â, Ĉ]

7. Commutator is distributive with respect to first operator:- [ÂB̂, Ĉ] = Â[B̂, Ĉ] + [Â, Ĉ]B̂

8. Jacobi Identity:- Cyclic operation of commutator is zero, i.e., [Â, [B̂, Ĉ]]+[B̂, [Ĉ, Â]]+[Ĉ, [Â, B̂]] = 0

9. Commutator of two Hermitian operators are anti-hermitian, i.e., If  and B̂ are hermitian operators
then [Â, B̂] = ÂB̂ − B̂Â is anti-hermitian.

10. Anti-commutator of two Hermitian operators are hermitian, i.e., If  and B̂ are hermitian operators
then {Â, B̂} = ÂB̂ + B̂Â is hermitian.

10
Types of Operators
 
1 0 0 ... 0
0 1 0 . . . 0
 
1. Identity Operator:- Î = 0 0 1 . . . 0  = diag(1, 1, 1 . . . 1)n

 .. .. .. . . .. 
. . . . .
0 0 0 ... 1

2. Hermitian and Anti-hermitian Operators:- An operator is said to be Hermitian if  = † and


an operator is said to be anti-hermitian if  = −† . Any linear operator can be expressed as the
sum of a hermitian and an anti-hermitian operator by,
 + †  − †
 = +
2 }
| {z 2 }
| {z
hermitian anti-hermitian

3. Unitary Operator:-  is unitary if † = Â−1 . So, if  and B̂ are unitary, then ÂB̂ is unitary.
4. Projection Operator:-  is a projection operator if it is Hermitian † =  and if  = Â2 . So two
projection operators P̂1 and P̂2 are orthogonal if P̂1 P̂2 = 0. And the sum of projection operators
P̂1 + P̂2 + P̂3 + · · · + P̂n is also projection operator if and only if P̂1 , P̂2 , P̂3 , . . . , P̂n are mutually
orthogonal to each other.

2.3 Eigen Value problem


Consider an operator  and a vector |Ψ i, then if they satisfy the below condition,
Â|Ψ i = λ|Ψ i (2.9)
Where λ is a scalar quantity. The vector |Ψ i is called as the Eigen vector of  and λ is the eigen value of
the eigen vector Ψ of the operator Â. And the whole equation is called as the Eigen value equation. To
get the eigen values λ1 , λ2 , λ3 , . . . of an operator and a vector we can do this below transformation,
Â|Ψ i = λ|Ψ i = λÎ|Ψ i (2.10)
⇒ Â|Ψ i − λÎ|Ψ i = |0i (2.11)
⇒ ( − λÎ)|Ψ i = |0i (2.12)

Now if we say that the operator  − λÎ is having an inverse, i.e., if its determinant is non zero, then we
get a trivial solution |Ψ i = |0i, but for non-trivial solutions, we have to restrict that  − λÎ doesn’t have
an inverse, i.e., whose determinant is zero. Thus we get,

a11 − λ a12 a13 ... a1n

a21
a22 − λ a23 ... a2n
a31
| − λÎ| = a32 a33 − λ . . . a3n = 0 (2.13)
.. .. .. .. ..
. . . . .



an1 an2 an3 . . . ann − λ
The expansion of the determinant yields an algebraic Polynomial expression for λ which is called as the
characteristic equation for the operator  given by,
P (λ) = C1 λ1 + C2 λ2 + C3 λ3 + · · · + Cn λn = 0 (2.14)
Which has not more than n number of solutions. If the distinct solutions are of the form λ1 , λ2 , λ3 , . . . , λm ,
where m ≤ n, then by using  and eigen values one at time we get ‘m’ number of eigen vectors
(|Ψ1 i, |Ψ2 i, |Ψ3 i, . . . , |Ψm i) which satisfy Â|Ψi i = λi |Ψi i where 1 ≤ i ≤ m. If  is an operator and Ψ is
its eigen function whose corresponding eigen value is λ, then we can tell,
Â|Ψ i = λ|Ψ i ⇒ f (Â)|Ψ i = f (λ)|Ψ i where f is some polynomial function (2.15)

11
Some useful theorems
Theorem 1. The Eigen values of a Hermitian operator are always real.
Proof. Consider a Hermitian operator Â. According to Eq.(2.9), it has to satisfy,
Â|Ψ i = |ÂΨ i = λ|Ψ i = |λΨ i (2.16)
Taking the inner product of the above equation after multiplying both sides by hΨ | we get,
hΨ |ÂΨ i = hΨ |λΨ i = λhΨ |Ψ i (2.17)
Now according to Eq.(2.7) we can write,
hΨ |ÂΨ i = h† Ψ |Ψ i (2.18)
† †
Since  =  ⇒ hÂ Ψ |Ψ i = hÂΨ |Ψ i (2.19)
Since ÂΨ = λΨ ⇒ hÂΨ |Ψ i = hλΨ |Ψ i = λ∗ hΨ |Ψ i (2.20)
As we know that if |Ψ i =
6 |0i then hΨ |Ψ i =
6 0 because of positive definiteness, we can equate Eq.(2.17) and
Eq.(2.20) and get the result λ = λ∗ which implies λ is a real quantity. Thus the eigen value of a hermitian
operator is always a real quantity.
Theorem 2. The expectation value of a Hermitian operator corresponding to an eigen vector is real.
hΨ |Â|Ψ i
Proof. We know that the expectation value of an operator is given by,
hΨ |Ψ i
hΨ |Â|Ψ i hΨ |ÂΨ i hΨ |λΨ i hΨ |Ψ i
⇒ = = =λ =λ which is real
hΨ |Ψ i hΨ |Ψ i hΨ |Ψ i hΨ |Ψ i
Theorem 3. Eigen functions of a Hermitian operator of distinct eigenvalues are mutually orthogonal.
Proof. Let  be a Hermitian operator. And Ψ1 , Ψ2 be two eigen functions whose eigen values are λ1 and
λ2 respectively where λ1 6= λ2 . So we get,
Â|Ψ1 i = λ1 |Ψ i and Â|Ψ2 i = λ2 |Ψ i (2.21)
Since  is Hermitian we can write, hΨ1 |ÂΨ2 i = hÂΨ1 |Ψ2 i ⇒ hΨ1 |λ2 Ψ2 i = hλ1 Ψ1 |Ψ2 i
⇒ λ2 hΨ1 |Ψ2 i = λ∗1 hΨ1 |Ψ2 i (2.22)
But, λ∗1 = λ1 as Hermitian operator has real eigen values. Thus we get,
λ2 hΨ1 |Ψ2 i = λ1 hΨ1 |Ψ2 i (2.23)
But we know that λ1 and λ2 are distinct and not necessarily be zero. So only way to hold the above
expression is only by equating the inner product of Ψ1 and Ψ2 to Zero. ⇒ hΨ1 |Ψ2 i = 0
Theorem 4. Eigen vectors of a hermitian operator corresponding to distinct eigen values will form a
complete set of mutually orthonormal basis vectors and the linear combination of these eigen vectors can
represent any vector in hilbert space.
Proof. Mathematically we can tell, if Ψ1 , Ψ2 , Ψ3 , . . . , Ψn are eigen vectors of a hermitian operator  whose
corresponding eigen values are λ1 , λ2 , λ3 , . . . , λn , then we can concise the above statement as ÂΨi = λi Ψi
where λi 6= λj ⇒ Ψi 6= Ψj ∀ 1 ≤ i, j ≤ n and i 6= j. If hΨi |Ψj i = δij and if Φ is some vector in
Hilbert space H, then we can write it as,
Xn
Φ = c1 Ψ1 + c2 Ψ2 + c3 Ψ3 + · · · + cn Ψn = ci Ψi (ci is a scalar) (2.24)
i=1

Theorem 5. The eigen values and expectation values of an Anti-Hermitian operator are purely imaginary.
Theorem 6. Eigen values of a unitary operator are complex numbers whose magnitude is always 1. And
Eigen vectors of unitary operators are mutually orthogonal to each other provided that all the eigen values
are distinct.

12
2.4 Generalized Uncertainty Principle
We wish to statistically find the relation between the standard deviations of various measurements by two
operators. Consider two operators  and B̂, the expectation value of an operator is given by the mean
value of all the observed measurements. So expectation value of  is hÂi and B̂ is hB̂i. The deviation
from the observed value by measurement and expectation value is denoted by ∆ and ∆B̂,

∆ =  − hÂi and ∆B̂ = B̂ − hB̂i (2.25)


⇒ 2 2 2
(∆Â) =  + hÂi − 2hÂi and 2 2 2
(∆B̂) = B̂ + hB̂i − 2hB̂iB̂ (2.26)

The expectation value of (∆Â)2 is called as the variance of a set of measurements which is the sum of the
expectation values of Â2 and −2hÂi summed with hÂi2 which is hÂ2 i − 2hÂihÂi − hÂi2 , thus,

(∆Â)2 = σA
2
= hÂ2 i − hÂi2 and (∆B̂)2 = σB
2
= hB̂2 i − hB̂i2 (2.27)

∆B̂∆ ∆B̂∆Â
Consider the term ∆Â∆B̂ = ∆Â∆B̂ + − (2.28)
2 2
∆Â∆B̂ ∆B̂∆ ∆Â∆B̂ ∆B̂∆Â
= − + + (2.29)
2 2 2 2
1 1
= [∆Â, ∆B̂] + {∆Â, ∆B̂} (2.30)
2 2

Let us find the expression for [∆Â, ∆B̂] now, expanding the terms we get,

[∆Â, ∆B̂] = ∆Â∆B̂ − ∆B̂∆ (2.31)


= (Â − hÂi)(B̂ − hB̂i) − (B̂ − hB̂i)(Â − hÂi) (2.32)
 
= ÂB̂ − 
hÂi
 B̂ −  hÂihB̂i − B̂Â − 
ÂhB̂i − 
hÂi hÂih (2.33)
    
ÂhB̂i +   B̂ +   B̂i
= ÂB̂ − B̂Â (2.34)
⇒ [∆Â, ∆B̂] = [Â, B̂] (2.35)

1 1
Thus we get, ∆Â∆B̂ = [Â, B̂] + {∆Â, ∆B̂} (2.36)
2 2
1 1
Taking their expectation values h∆Â∆B̂i = h[Â, B̂]i + h{∆Â, ∆B̂}i (2.37)
2 2
1
⇒ h∆Â∆B̂i ≥ h[Â, B̂]i (2.38)
2
1
⇒ h∆Â∆B̂i2 ≥ h[Â, B̂]i2 (2.39)
4
By Schwartz’s inequality we get, h∆Â2 ih∆B̂2 i ≥ h∆Â∆B̂i2 (2.40)
Thus by the above two equations we get,
1
h∆Â2 ih∆B̂2 i ≥ h[Â, B̂]i2 (2.41)
4
1
Using Eq.(2.27) we get, h∆Â2 i = σA
2
and h∆B̂2 i = σB
2
⇒ 2 2
σA σB ≥ h[Â, B̂]i2 (2.42)
4
1
Taking the square root we get the generalized uncertainty principle σA σB ≥ h[Â, B̂]i (2.43)

2
where σA and σB are the standard deviations of a set of measurements of the two physical quantities
described by the operators  and B̂ respectively.

13
2.5 Function space
We know how a vector and its dual looked in a finite dimension space in Eq.(2.4). So we can say that
the vector |Ψ i has n-number of components. We can also denote it by Ψ (i) = vi . But if the vector |Ψ i
was considered to be a function whose domain is space and time coordinates, then we can represent it as,
|Ψ i = Ψ (t, x, y, z). Since the coordinates can take values from −∞ to +∞, it would mean |Ψ i has an infinite
number of components and it follows all properties mentioned in the sections (2.1, 2.1, and 2.1).

Some analogous operations in Function space


1. Hermitian conjugate of functions in Hilbert space:- Ψ ∗ (x, y, z, t) is the hermitian conjugate of
Ψ (x, y, z, t) which is got by substituting i → (−i) where ever it appears in the function.
Example:- If Ψ = ei(kx−ωt) then Ψ ∗ = e−i(kx−ωt) and the term Ψ ∗ Ψ = |Ψ |2

2. Inner product of functions in Hilbert space:- The inner product of two functions in Hilbert space
is given by, ˆ ∞
hΦ|Ψ i = Φ∗ (x, y, z, t)Ψ (x, y, z, t) dx dy dz = finite scalar (2.44)
−∞
ˆ ∞
3. Norm of a function is given by:- hΨ |Ψ i = Ψ ∗ Ψ dx dy dz
−∞
We have to note that all the functions in Hilbert space have to be such that the norm is a finite scalar
quantity. Only then do the functions
ˆ fit in the Hilbert space. Thus all the functions (Ψ ) in Hilbert space

|Ψ |2 = finite scalar quantity.
have to be square-integrable, i.e.,
−∞ ˆ ∞

4. Two functions are orthonormal if :- Ψm Ψn dx dy dz = δmn
−∞

5. Operators:- In function space the operators are not matrices, rather they are mathematical tools like
differentiation, etc. And all operators in function space has the properties mentioned in section(2.2, 2.2).
d ∂ ∂2 ∂2 ∂2 ∂2 ∂ ∂
Examples:- , , ia 2 , 2
+ 2 + 2, + − 1 etc.
dx ∂t ∂y ∂x ∂y ∂z ∂t ∂x
ˆ ∞
Ψ ∗ (ÂΨ ) dx dy dz
hΨ |Â|Ψ i
6. Expectation value of an operator = = ˆ ∞
−∞
hΨ |Ψ i
Ψ ∗ Ψ dx dy dz
−∞

14
Chapter 3

Wave Function and Quantum mechanics

3.1 Wave Function


According to De Broglie’s hypothesis even matter had wave properties and behaves as a wave. Then there
has to be a mathematical description for these waves. One such tool is called a ‘Wave Function’ denoted
by Ψ or |Ψ i.

d2 x dV
In classical mechanics we have an equation of motion like F = m 2 = − when solved for x, it gives a
dt dx
function of its position and then we could find its velocity, momentum, kinetic energy, potential energy, etc,
and know all the properties of the particle. The same analog in quantum mechanics is the Wave function
which has the information regarding a system it describes.

Born’s Interpretation
ˆ x+dx ˆ y+dy ˆ z+dz
According to Born’s probabilistic interpretation of wave function, hΨ |Ψ i = Ψ ∗ Ψ dx dy dz
x y z
gives the probability of a particle to exist in a region of space between x, y, z and x + dx , y + dy , z + dz.
Then the quantity Ψ ∗ Ψ = |Ψ |2 is the probability density function of the particle. This implies that Ψ is the
probability amplitude function of the particle. Consider a wave function of a 1D particle when Ψ and |Ψ |2
plotted looks like the below figure.

So here we can see the wave function Ψ and also its probability density |Ψ |2 . So we see that the location
where |Ψ |2 = 0 namely x = 0, x2 , x6 and a, the implies that at those locations the probability of finding
the particle is zero. But at x = x7 the density function is the highest, so the particle is most likely to be
in that region. And at x = x4 we see that the probability is very less so particle is very less likely to be
present. And at x = x3 , x5 the probabilities are the same, and at x = 1 the probability is lesser than that
of x = x3 , x5 . Thus we can arrange the probabilities as, x7 > x5 = x3 > x1 > x4 > x2 = x6 = 0.

15
Properties of a Wave Function
1. |Ψi can interfere with itself thus we see that particles showing diffraction and interference patterns.
2. Ψ is a complex-valued function of the form ψ1 + iψ2 or reiθ . And Ψ ∗ is got by replacing i → (−i).
3. The quantity Ψ ∗ Ψ = |Ψ |2 is the probability density function of the particle.

Conditions for a wave function


1. Ψ should be a finite quantity for all values of x, y, z, t.
2. At x, y, z → ∞ or at the Boundary of the surface where the particle resides Ψ → 0
3. |Ψi and its derivatives have to be continuous and differentiable, except at places where P.E. = ∞
and Ψ should be a single-valued function. i.e., Ψ can’t have two or more values for same input, i.e.,
at same location and time (x, y, z, t), Ψ can’t have two or more values.
ˆ ∞
4. Ψ has to be square-integrable. i.e., the quantity hΨ |Ψ i = Ψ ∗ Ψ dx dy dz = Finite scalar value.
−∞

5. Ψ has to be normalized. i.e. the inner product should be hΨ |Ψ i = 1 because of Born’s interpretation.
Since the particle has to be in at least one place, the probability of it being throughout space from
−∞ to ∞ will be 1.
6. If c1 |Ψ1 i, c2 |Ψ2 i, . . . , are wave functions of a system
X then |Ψ i = c1 |Ψ1 i + c2 |Ψ2 i + . . . is also a wave
functions of the system subject to the condition c2i = 1 because of the normalization condition.
i

The normalization condition for Ψ is time-invariant


Consider that we normalize a wave function and some time t, but will the normalization condition change
for some other time? For simplicity consider a 1D wave function, So if Ψ changes as time evolve, then let
us try to total differentiate the normalization condition with time,
ˆ ˆ ∞
d ∞ ∗ ∂
Ψ ∗ Ψ dx (3.1)

Ψ Ψ dx =
dt −∞ −∞ ∂t
ˆ ∞
∂Ψ ∂Ψ ∗ 
= Ψ∗ +Ψ dx (3.2)
∂t ∂t
ˆ−∞
∞  ˆ ∞ 
∂Ψ  ∂Ψ ∗ 
= Ψ∗ dx + Ψ dx (3.3)
−∞ ∂t −∞ ∂t
| {z }
I

Let us apply integration by parts to the above integral I.


ˆ ∞ ˆ ∞
∂Ψ ∗   ∗  +∞ ∂Ψ 
I= Ψ dx = Ψ Ψ −∞ − Ψ∗ dx (3.4)
−∞ ∂t | {z } −∞ ∂t
0
+∞
Now Ψ ∗ Ψ = 0 because of the boundary conditions. Then putting the value of I back we get,

−∞
ˆ ˆ ∞ ˆ ∞
d ∞ ∗ ∗ ∂Ψ
 ∂Ψ 
Ψ Ψ dx = Ψ dx − Ψ∗ dx = 0 (3.5)
dt −∞ −∞ ∂t −∞ ∂t
ˆ
d ∞ ∗
⇒ Ψ Ψ dx = 0 (3.6)
dt −∞
ˆ ∞
⇒ Ψ ∗ Ψ dx = Constant with time (3.7)
−∞
Thus we got the normalization condition to be time-independent. Implies that if a wave function is nor-
malized at one time then it will continue to remain normalized at any future time.

16
3.2 Postulates of Quantum Mechanics
Postulate 1:-
The state of the particle at a region (x, y, z) and time t is represented by a vector (or a wave function) |Ψ i
that belongs to Hilbert space. And the wave function is square-integrable, continuous (including its first
and second derivatives) and tends to zero at boundary conditions or when x, y, z → (∞, −∞) ⇒ |Ψ i → 0

Postulate 2:-
The probability of finding a particle (described by |Ψ i) in a region of space between x, y, z and x + dx , y +
dy , z + dz i.e. in the region (dx dy dz) is given by the inner product of |Ψ i.
ˆ x+dx ˆ y+dy ˆ z+dz
Probability = hΨ |Ψ i = Ψ ∗ Ψ dx dy dz (3.8)
x y z

Thus the quantity Ψ ∗ Ψ which we integrate upon space to find the probability is called Probability density.
And since the particle if it exists it has to be in space, the Probability
ˆ of finding it between −∞ and ∞

should be 1. This is the Normalization condition given by, hΨ |Ψ i = Ψ ∗ Ψ dx dy dz = 1
−∞

Postulate 3:-
To every observable quantity in classical mechanics there exists a linear hermitian operator in quantum
mechanics which describes the motion of quantum particles. A hermitian operator satisfies the conditions,

† =  and hΨ |ÂΨ i = hÂΨ |Ψ i and hΨ |Â|Ψ i = hΨ |Â|Ψ i∗ (3.9)

Postulate 4:-
The only possible values that can be obtained from the measurement of an observable of a system whose
operator is  are the eigen values an of the equation whose eigen functions (vectors) are |Ψn i.

Â|Ψn i = an |Ψn i (Â is the operator and an is the observable quantity) (3.10)

Postulate 5:-
The expectation value of an observable quantity whose corresponding operator is  for a system that is
represented by a normalized |Ψ i is a real quantity, and is got by the inner product of Ψ with Â|Ψ i.
ˆ ∞
hÂi = hΨ |Â|Ψ i = Ψ ∗ (ÂΨ ) dx dy dz (3.11)
−∞

3.3 Schrodinger’s wave equation


There is no actual derivation for Schrodinger’s equation, it was just proposed by Schrodinger. But there
is a subtle way to derive it. Now we know that any matter particle behaves as a wave according to De
Broglie’s equation. That wave is represented by wave function Ψ . Any wave has the form of,

Ψ (x, y, z, t) = Aei(kx x+ky y+kz z−ωt) (3.12)

Where A is the amplitude, kx , ky , kz are the wave constants in the respective directions and ω is the angular
frequency of the wave. Now we shall use De Broglie’s hypothesis in Eq.(1.19) since P = }k we can write,
Px Py Pz
kx = ky = kz = (3.13)
} } }
h E
Also we can write E = hν as E = · 2πν = }ω ⇒ ω= Now the wave equation becomes,
2π }
i
Ψ (x, y, z, t) = Ae } (Px x+Py y+Pz z−Et) (3.14)

17
Let us find the derivatives of Ψ ,
∂Ψ i i i ∂Ψ
= − EAe } (Px x+Py y+Pz z−Et) = − EΨ ⇒ EΨ = i} (3.15)
∂t } } ∂t
∂Ψ i i i ∂Ψ
= Px Ae } (Px x+Py y+Pz z−Et) = Px Ψ ⇒ Px Ψ = −i} (3.16)
∂x } } ∂x
∂Ψ i i i ∂Ψ
= Py Ae } (Px x+Py y+Pz z−Et) = Py Ψ ⇒ Py Ψ = −i} (3.17)
∂y } } ∂y
∂Ψ i i i ∂Ψ
= Pz Ae } (Px x+Py y+Pz z−Et) = Pz Ψ ⇒ Pz Ψ = −i} (3.18)
∂z } } ∂z
∂2Ψ i2 2 i (Px x+Py y+Pz z−Et) Px2 ∂ 2Ψ
= P Ae } = − Ψ ⇒ P 2
Ψ = −}2
(3.19)
∂x2 }2 x }2 x
∂x2
2
∂2Ψ i2 i Py ∂2Ψ
2
= 2 Py2 Ae } (Px x+Py y+Pz z−Et) = − 2 Ψ ⇒ Py2 Ψ = −}2 2 (3.20)
∂y } } ∂y
2
∂ Ψ 2
i 2 i (Px x+Py y+Pz z−Et) Pz2 2
2∂ Ψ
= P Ae } = − Ψ ⇒ P 2
Ψ = −} (3.21)
∂z 2 }2 z }2 z
∂z 2
Now we have the relation for the total energy of the wave to be T + V = E where T is the kinetic energy
1 P2 Px2 + Py2 + Pz2
and V is the potential energy. We also know that T = mv 2 = = Thus we get,
2 2m 2m
Px2 Py2 P2
+ + z +V =E (3.22)
2m 2m 2m
Px2 Ψ Py2 Ψ P 2Ψ
(multiplying the above equation by Ψ ) + + z + V Ψ = EΨ (3.23)
2m 2m 2m
Substituting the expressions for EΨ , Px Ψ , Py Ψ , and Pz Ψ in the above equation we get,
2 2 2

}2 ∂ 2 Ψ }2 ∂ 2 Ψ }2 ∂ 2 Ψ ∂Ψ
− 2
− 2
− 2
+V (x, y, z, t) Ψ = i} V (x, y, z, t) → time dependent potential (3.24)
2m ∂x 2m ∂y 2m ∂z ∂t
" #
}2 ∂ 2 Ψ ∂2Ψ ∂2Ψ ∂Ψ
⇒ − 2
+ 2
+ 2
+ V (x, y, z, t)Ψ = i} (3.25)
2m ∂x ∂y ∂z ∂t

The above equation is called as Time-Dependent Schrodinger equation where Ψ depends on (x, y, z, t).
When solved for Ψ , it gives the wave function of the system. We can also further simplify the above
equation and get,

}2 2 ∂Ψ
− ∇ Ψ + V Ψ = i} where ∇2 is the Laplacian operator (3.26)
2m ∂t
" #
}2 2 ∂Ψ
⇒ − ∇ + V Ψ = i} (3.27)
2m ∂t

Let us define the Hamiltonian operator to be Ĥ = i} , then the above equation becomes,
∂t
" #
}2 2 ∂Ψ
ĤΨ = − ∇ + V Ψ = ÊΨ = i} (3.28)
2m ∂t
This is a way to arrive at Schrodinger’s equation which describes the state of a quantum system. This
equation can be more understood by its implications. It is a statement of conservation of energy for
quantum systems. Which is of the form Total energy = Kinetic energy + Potential energy. Though the
equation is self-explanatory, it explains the particles moving at non-relativistic speeds only and changes
under Lorentz transformations, so for systems moving at relativistic speeds, we need to modify the above
equation where the quantities will become Lorentz invariant.

18
3.4 Operators in Quantum Mechanics
The quantum realm is not deterministic because we can’t find the exact values of the physical quantities
(observable) of a particle, rather the corresponding operators when operated on the wave function Ψ yields
an eigen value that is an observable property. Some operators are,

Quantum Mechanical operators


Observable Operator Symbol Expression

Position x̂ Multiply by x

Momentum in x-axis P̂x −i}
∂x

Momentum P̂ ~
−i}∇

∂2
Momentum squared P̂2 −}2
∂x2

Time t̂ Multiply by t

Kinetic Energy in x-axis T̂x −}2 ∂ 2


2m ∂x2

−}2 2
Kinetic Energy T̂ ∇
2m

Hamiltonian / ∂
Ĥ, Ê i}
Total energy ∂t

Multiply by V (x, y, z, t)
Potential Energy V̂

These are some of the operators which we have dealt with while deriving Schrodinger’s equation. x̂ ,P̂ ,t̂ ,
and Ĥ are the fundamental operators and all the rest of the operators are derived from them. And since
these fundamental operators are Hermitian (proof is dealt next), all derived operators are also Hermition.

Momentum operator is Hermitian


An operator  is Hermitian if hΦ|ÂΨ i = hÂΦ|Ψ i. Using this identity we can show that P̂x is hermitian.
ˆ ∞ ˆ ∞ ! ˆ ∞
∗ ∗ ∂Ψ ∂Ψ
hΦ|P̂x Ψ i = Φ (P̂x Ψ )dx = Φ − i} dx = −i} Φ∗ dx (3.29)
−∞ −∞ ∂x −∞ ∂x
ˆ ∞ " ˆ ∞ #
∗ ∂Ψ
 ∗  +∞ ∂Φ∗
(Integrating by parts) − i} Φ dx = −i} Φ Ψ −∞ − Ψ dx (3.30)
−∞ ∂x | {z } −∞ ∂x
0
ˆ ∞ ˆ ∞
!∗
∂Φ∗ ∂Φ
⇒ hΦ|P̂x Ψ i = i} Ψ dx = − i} Ψ dx = hP̂x Φ|Ψ i (3.31)
−∞ ∂x −∞ ∂x

Thus we got hΦ|P̂x Ψ i = hP̂x Φ|Ψ i which implies that P̂x is a Hermitian operator. Similarly, we can also
show that P̂y and P̂z are also Hermitian. Thus P̂ = P̂x î + P̂y ĵ + P̂z k̂ = −i}∇
~ is also Hermitian. Also, we
can easily show x̂ is also hermitian since it is just the multiplication of x to the wave function. And since
P̂2
T̂ = which is derived from a hermitian operator P̂ we can tell that T̂ is also Hermitian.
2m

19
Hamiltonian operator is Hermitian
ˆ ∞
hΦ|ĤΨ i = Φ∗ (ĤΨ )dx (3.32)
ˆ−∞
∞ ˆ ∞
∂Ψ∗ ∂Ψ
= Φ i} dx = i} Φ∗ dx (3.33)
−∞ ∂t −∞ ∂t
ˆ ∞ " ˆ ∞ ∗
#
∂Ψ +∞ ∂Φ
(Integrating by parts) Φ∗ dx = i} Φ∗ Ψ −∞ − (3.34)
 
i} Ψ dx
−∞ ∂t | {z } −∞ ∂t
0
ˆ ∞ ∗
! ˆ ∞ !∗
∂Φ ∂Φ
⇒ hΦ|ĤΨ i = − i} Ψ dx = i} Ψ dx = hĤΦ|Ψ i (3.35)
−∞ ∂t −∞ ∂t

Thus we got hΦ|ĤΨ i = hĤΦ|Ψ i which implies that Ĥ is a Hermitian operator. Since all quantum mechanical
operators are Hermitian operators we can say that their Eigen values and Expectation values are real.

3.5 Measurements in quantum Mechanics


3.5.1 Wave Function Collapse
In quantum mechanics, we have a very unique property of quantum systems. The act of measurement of
a physical quantity or observing the quantum system changes the state of the system completely. This
is known as the Observers effect. Before measurement of a quantity, that quantity in the system will be
probabilistic, whose probability amplitude is given by the wave function Ψ . But once the measurement is
done, the probabilistic value will be known because of the measurement and then the wave function will
collapse into a peak. This phenomenon is called as the collapsing of the wave function. This is also similar
to the change in the wave function that is done by the operation of an operator on a wave function.

We can understand this concept experimentally, consider the interference of electrons through the double
slit. Even when single electrons are shot by an electron gun one after the other towards the double slit, it
forms a pattern where the probability of finding the electron goes from peak to zero in form of a wave that
corresponds to the bright and dark bands of the interference pattern.

But if we try to observe the electron through a detector or any such device to know through which slit the
electron passes, then it acts as a particle and the interference pattern disappears and the resulting pattern
is in the form of a narrow peak with a central maximum. Thus the act of measurement has changed the
result of the experiment completely

20
3.5.2 Quantum Indeterminacy
Consider that we measure some physical quantity of a particle, and then we find the value to be x1 , then
we pose a question of where the particle was just before we measure it? There could be three situations
that characterize Quantum Indeterminacy:-

The realist value


The physical quantity we measured was x1 itself. Though it is reasonable it makes quantum mechanics
an incomplete theory because though the physical quantity was x1 before itself, but it was indeterminate
before measurement just because the observer was not known of it.

The orthodox value


According to this stand, the particle was not having a particular value of the physical quantity before
measurement, but the act of measurement made the particle have a particular value for the measured
physical quantity. But why it had to take the value of x1 itself we don’t know. But it could be forced to
take the value and the system was disturbed. This is known as the ‘Copenhagen interpretation’ which is
widely accepted by physicists.
1. Quantum mechanics is probabilistic. And probabilities are primary.
2. The observational means have to be explained in terms of classical mechanics.
3. The act of observation is irreversible. And will leave a mark on the system.
4. The act of observation involves an action upon the object measured and reduces the wave packet
5. Wave function collapse when the measurement is made is a transition from potentiality to actuality.
6. Complementary properties cannot be observed simultaneously (Uncertainty principle).
7. Only the results of observation are true. All the rest are probabilistically existing.

The agnostic value


This is means that there is no answer for such questions. Because To get to know the value of some quantity
we need to measure it as we haven’t measured still, there is no need to ask such a question.

3.5.3 Expectation Values of Operators


Expectation value is not the average of consecutive measurements of a quantity in a particular system
continuously. Rather it is the average of the measured physical quantity on an ‘n’ number of identical
systems or ‘n’ number of measurements on the same system where consecutive measurements are done
only after the system returns to the original state after the current measurement. Because if we do the
measurement again and again on the same system without it returning to initial conditions the values will
repeat because the wave function has changed to a spike and since it is a spike, the values won’t change.
As we have seen earlier about the wave function collapse. Thus it is necessary to give breaks between
consecutive measurements or we have to measure the value of the quantity in identical systems. If  is an
operator corresponding to an observable physical quantity of a system described by a wave function |Ψi,
then the expectation value of  is given by,
ˆ ∞
Ψ ∗ (ÂΨ ) dx dy dz
hΨ |Â|Ψ i
hÂi = ˆ ∞
= −∞ (3.36)
hΨ |Ψ i
Ψ ∗ Ψ dx dy dz
−∞
ˆ ∞
But if |Ψ i is normalized then hΨ |Ψ i = Ψ ∗ Ψ dx dy dz = 1, thus the expectation value is hÂi = hΨ |Â|Ψ i.
−∞

21
3.6 Heisenberg’s Uncertainty Principle
We came across the Generalized Uncertainty principle in section(2.4) which was derived based on the
statistical aspects of standard deviation whose equation is given in Eq.(2.43),
1
σA σB ≥ h[Â, B̂]i (3.37)

2
where σA is the standard deviation in the measured values or the outcome values of the physical quantity
corresponding to the operator  and similarly σB is the standard deviation in the measured values of the
physical quantity corresponding to the operator B̂. The product of these two quantities will be proportional
to the absolute value of the expectation of the commutator of the two operators.

3.6.1 Position-Momentum uncertainty


We know x̂ = x and P̂x = −i} ∂x∂
. Substituting  = x̂ and B̂ = P̂x in generalized uncertainty equation,
1
σx σPx ≥ h[x̂, P̂x ]i (3.38)

2
 h ∂ i
Now let us find the RHS of the above equation, x̂, P̂x = x, −i} (3.39)

∂x
! !
h ∂ i ∂Ψ ∂
(Multiplying throughout by Ψ ) x̂, P̂x Ψ = x, −i} (3.40)
 
Ψ = x − i} − − i} (xΨ )
∂x ∂x ∂x
!
∂Ψ ∂x ∂Ψ ∂Ψ
 ∂Ψ

= −i}x + i} Ψ +x = −i}x + i}Ψ + i}x (3.41)
 
∂x ∂x ∂x  ∂x  ∂x
 

(3.42)
   
⇒ x̂, P̂x Ψ = i}Ψ ⇒ x̂, P̂x = i}
Since the commutator value is a constant, the expectation value is also constant. So, h x̂, P̂x i = i} . Now
 

the modulus value of this quantity is given by,


  √ }
h x̂, P̂x i = }2 = } ⇒ σx σPx ≥ (3.43)

2
We know that more the standard deviation of a set of values more spread is the values and the measurement
error will be more and if the standard deviation for a set of values is less that means the values are very
close and the error in measurement is very less. And since the product of standard deviations of position
and momentum is greater or equal to a finite quantity then that means that the lesser the measurement
error of one quantity, the more is the measurement error in the other quantity. Thus the error cannot be
zero for both the quantities since their product is a non-zero number.

The above figure gives a pictorial representation of the uncertainty principle in position and momentum.

22
Case - 1
We see that the wave function of a particle is in form of a sinusoidal curve where the wavelength is highly
defined and the error is very very less and λ is known accurately, and since we know that momentum is
Px = λh , the error in measuring Px is very less. But the position of the particle is not known at all as the
wave function is spread throughout the x-axis, thus the error in position is very large.

Case - 2
In case 2 the wave function of the particle is in to form of a steep peak which is called a Delta function.
The location of the particle is highly defined which is given to be x1 and the error in position is very less.
But the wavelength of the particle is not defined correctly, thus the momentum of the particle as it depends
on the wavelength cant be defined correctly, thus the error in momentum is very large.

Case - 3
But here we can see that the wave function is localized in x-axis within a distance of ∆x, in such condition
the location of the particle lies within that distance whose error doesn’t exceed ∆x, also the wavelength of
the particle can be found out pretty well whose error won’t exceed ∆λ and then the error in momentum
will be ∆Px . The value of ∆x and ∆Px is such that their product is greater than or equal to }/2.

Note that this uncertainty in the particle’s position and momentum is a fundamental property of quantum
systems and is not because of the limitations of the observer or instrument of measurement. We can show
that this is a fundamental property by writing Schrodinger’s equation in momentum space.

3.6.2 Schrodinger’s equation in momentum space


We are familiar with the wave function that depends on the position and time denoted by Ψ (x, y, z, t). This
represents the probability amplitude of the particle which describes the probability of the particle to be in
a particular position (x, y, z) at some time (t). Such a wave has the form,
~
Ψ (x, y, z, t) = Aei(k·~r−ωt) (3.44)
But the same analogy in momentum space, the momentum wave function denoted by ΨP (Px , Py , Pz , t) that
depends on the momentum of the system and time. This represents the probability amplitude of the particle
that describes the probability of the particle to have momentum P~ = (Px , Py , Pz ) at some time (t). To get
ΨP (Px , Py , Pz , t) we need to apply Fourier transformation which results in,
ˆ
1 i ~
ΨP (Px , Py , Pz , t) = p A(P ) e } (P ·~r−Et) dp (3.45)
(2π})3

Then when we do the same process that we did in arriving at Schrodinger’s equation in position space, we
will get the Schrodinger’s equation in momentum space that is given by,
! ˆ
P2 1 i ~
−E p A(P ) e } (P ·~r−Et) dp + V Ψ = 0 (3.46)
2m (2π})3

When this equation is solved we get the wave function of the particle in momentum space. A striking fact
that we get is, if we have the wave function in position space to be a Delta function in form of a peak then
the wave function in momentum space which is got be the Fourier transform of delta function becomes
a sinusoidal curve where the momentum of the particle is spread out throughout the momentum space
implying that if the position is known accurately the error in momentum is very large. And if the wave
function of a particle in position space is sinusoidal, i.e., if the momentum is known accurately but the
error in position is very large, then in momentum space the wave function is in the form of a delta function
implying that momentum is known accurately in momentum space. Thus uncertainty is a fundamental
property and a consequence of the nature of Fourier transform of sinusoidal and delta functions.
Fourier Transformation
Sinusoidal curve Delta function
Inverse Fourier Transformation

23
3.6.3 Energy-Time uncertainty
We know Ê = i} ∂t

and t̂ = t . Substituting  = Ê and B̂ = t̂ in generalized uncertainty equation,

1
σE σt ≥ h[Ê, t̂]i (3.47)

2
 h ∂ i
Now let us find the RHS of the above equation, Ê, t̂ = i} , t (3.48)

∂t
! !
h ∂ i ∂ ∂Ψ
(Multiplying throughout by Ψ ) Ê, t̂ Ψ = i} , t Ψ = i} (tΨ ) − t i} (3.49)
 
∂t ∂t ∂t
!
∂t ∂Ψ ∂Ψ ∂Ψ ∂Ψ
(3.50)

 

= i} Ψ + t − i}t = i}Ψ + i}t − i}t
∂t ∂t ∂t  ∂t  ∂t

(3.51)
   
⇒ Ê, t̂ Ψ = i}Ψ ⇒ Ê, t̂ = i}
Since the commutator value is a constant, the expectation value is also constant. So, h Ê, t̂ i = i} .
 

  √ }
⇒ h Ê, t̂ i = }2 = } ⇒ σE σt ≥ (3.52)

2

Thus we get the product of standard deviations of measurements of Energy at some time to be greater or
equal to }/2. Since the product of standard deviations of Energy and time is greater or equal to a finite
quantity implies that the lesser the error in measurement of one quantity more is the error in measurement
in the other quantity. Thus we can’t find the energy of a system accurately at an exact time t. If we know
the energy of a system with high accuracy (very less error) then we won’t be knowing at what time the
energy of particle was having that value implying high error in measurement of time.

This uncertainty relation has a very important application. In the phenomenon of quantum tunneling
where the particle can cross barriers whose potential energy is greater than the particle’s total energy. Here
the particle can increase its total energy to ∆E greater than the potential energy within a time interval
of ∆t and can cross the barrier such that the inequality ∆E∆t ≥ }2 . According to this principle, the
Energy of a particle becomes uncertain if its lifetime is very short. i.e, if we exactly measure the time
interval of the quantum state or the lifetime of a particle (say unstable particle), then the energy of the
particle cannot be measured. So, then if we measure the particle’s energy, then its lifetime will be uncertain.

Note that for any two operators  and B̂, if they don’t commute, i.e., if the commutator bracket of
the operators is non-zero, then there will be uncertainty in the measurement of the physical observables
represented by the corresponding operators because of the generalized uncertainty principle.

3.7 Time Independent Schrodinger’s Equation


We know that Ψ (x, y, z, t) is the probability amplitude of the particle in the space coordinate (x, y, z) at
some time t. This is a bit tough to deal with because of the time component. What if we need to know
the wave function at a constant time, in the sense let it be any time, what would be the wave function?
For that let us analyze the Time-dependent Schrodinger’s equation in Eq.(3.25). If the potential term was
time-independent, i.e., if V → V (x, y, z), then the equation becomes,
" #
}2 ∂ 2 Ψ ∂2Ψ ∂2Ψ ∂Ψ
− 2
+ 2
+ 2
+ V (x, y, z)Ψ = i} (3.53)
2m ∂x ∂y ∂z ∂t

The above partial differential equation is homogeneous (no term independent of Ψ ) and linear (no term has
powers of Ψ and it’s derivatives). Thus we can use a technique called as separation of variables and write
the wave function of the form Ψ (x, y, z, t) = φ(t)ψ(x, y, z) . Substituting this in Schrodinger’s equation,

24
" #
}2 ∂ 2 ∂2 ∂2 ∂
(3.54)
 
− + + φ(t)ψ(x, y, z) + V (x, y, z)φ(t)ψ(x, y, z) = i} φ(t)ψ(x, y, z)
2m ∂x2 ∂y 2 ∂z 2 ∂t
" #
}2 ∂2ψ ∂2ψ ∂2ψ ∂φ(t)
⇒ − φ(t) 2
+ 2
+ 2
+ V (x, y, z)φ(t)ψ = i}ψ (3.55)
2m ∂x ∂y ∂z ∂t
" # !
1 }2 ∂ 2 ψ ∂ 2 ψ ∂ 2 ψ i} ∂φ(t)
⇒ − 2
+ 2
+ 2
+ V (x, y, z)ψ = (3.56)
ψ 2m ∂x ∂y ∂z φ(t) ∂t
Now we see that the LHS of the above differential equation is dependent only on space coordinates (x, y, z),
and the RHS depends only on time. So even if space coordinate changes the RHS remains constant with
respect to them since RHS doesn’t depend on space at all. Similarly, even if time changes the LHS remains
constant as it doesn’t depend on time, thus we can equate the above partial differentiation equation to some
constant E (the reason for it being E is discussed later). Now we can write the above equation as,
" # !
1 }2 ∂ 2 ψ ∂ 2 ψ ∂ 2 ψ i} ∂φ(t)
− + + + V (x, y, z)ψ = =E (3.57)
ψ 2m ∂x2 ∂y 2 ∂z 2 φ(t) ∂t

Thus we get two separate equations for ψ(x, y, z) and φ(t) which is,
" #
}2 ∂ 2 ψ ∂ 2 ψ ∂ 2 ψ
− + + + V (x, y, z)ψ = E ψ (3.58)
2m ∂x2 ∂y 2 ∂z 2

∂φ(t)
i} = E φ(t) (3.59)
∂t
Let us solve the time equation since it is simple, it becomes,
1 ∂φ(t) iE
=− (3.60)
φ(t) ∂t }
∂φ(t) iE
⇒ = − ∂t (3.61)
φ(t) }
iE iE
On integrating we get, log(φ(t)) = − t + log(A) ⇒ φ(t) = Ae− } t (3.62)
}
Then the equation corresponding to space given by Eq.(3.58) is called as the Time-independent Schrodinger’s
equation, when solved gives the wave function of the system at time a constant time t = t1 . Thus the full
solution of the wave function will be of the form,
iE
Ψ (x, y, z, t) = ψ(x, y, z) e− }
t
(3.63)

Note that A since is just a numerical constant, it can be absorbed into the time-independent wave function.

3.8 Stationary states in quantum mechanics


The stationary state in quantum mechanics means that all the physical observables of the system do not
iE iE
depend on time. And the wave function is of the form Ψ = ψ e− } t and its conjugate is Ψ ∗ = ψ ∗ e } t .
The exponential factors with the imaginary number i in it are called as the wiggle factors because they
can be shown to be of the form of a sinusoidal function of time, which means that as time proceeds they
make the space-dependent wave function oscillate hence they are called as wiggle factors. The separation of
variables was mainly because the potential energy of the system is time-independent, so in all future cases
if we want to use time-independent Schrodinger’s equation, then the potential energy of the system has to
be time-independent. Now let us understand some properties of a system whose wave function is separable
of space and time dependency.

25
Properties of stationary states
1. The Probability density function remains constant with time
iE iE
Though Ψ and Ψ ∗ depend on time, the probability density Ψ Ψ ∗ = ψ  e }t = ψψ ∗ = |ψ|2 which is
e−} t · ψ ∗ 
 
independent of time. Thus the probability density function remains constant and the probability of finding
the particle at a particular location where Ψ is defined remains constant with time.

2. The expectation values are constant with time


The expectation values of any observable which can be derived from r̂ (position operator) and P̂ (momentum
operator) remains constant with time.  Let  be anoperator of some physical quantity which can be derived
from r̂ and P̂, so, Â = f̂ (r̂, P̂) = f (x, y, z), −i}∇
~ . Since hAi = hΨ |ÂΨ i we get,

ˆ ∞   ˆ ∞  
iE
hAi = ∗ ~ 3
Ψ f̂ (x, y, z), −i}∇ Ψ d x = ψ∗ e }
t ~ ψ e− iE
f̂ (x, y, z), −i}∇ }
t 3
d x (3.64)
−∞ −∞

iE
The factor e− }doesn’t depend on space. So, it comes out of the operator f̂ and will be canceled.
t

ˆ ∞
iE  
iE

ψ∗ 
 −
hAi = hΨ |ÂΨ i = e }t  ~ ψ d3 x = hψ|Âψi
e } t f̂ (x, y, z), −i}∇ ⇒ hAi is time-independent
−∞

3. They have a definite total energy


By time-independent Schrodinger’s equation we get,
" # " #
}2 ∂ 2 ψ ∂ 2 ψ ∂ 2 ψ }2 2 }2 2
− + + + V ψ = E ψ ⇒ − ∇ ψ+Vψ =Eψ ⇒ − ∇ +V ψ =Eψ
2m ∂x2 ∂y 2 ∂z 2 2m 2m

}2 2
We know that the Hamiltonian operator is defined by, Ĥ = − ∇ +V (3.65)
2m
Thus the time-independent Schrodinger’s equation can be written concisely as, Ĥψ = E ψ (3.66)

Note that the above equation is an ‘Eigen value’ equation where the operator Ĥ operates on an Eigen
function ψ and yields an Eigen value E on operation. Thus the Eigen value of Hamiltonian is the total
energy of the system which is a constant. Then the expectation value of the Hamiltonian operator is,
ˆ ∞ ˆ ∞ ˆ ∞
hHi = hψ|Ĥψi = ψ ∗ (Ĥψ) d3 x = ψ ∗ (E ψ) d3 x = E ψ ∗ ψ d3 x = E (3.67)
−∞ −∞ −∞

Thus the expectation value of Ĥ is E. Now let us find the error in Ĥ which is σH = hH 2 i − hHi2 ,
p

ˆ ∞ ˆ ∞ ˆ ∞
∗ ∗
2 2
hH i = hψ|Ĥ ψi = 2 3
ψ (Ĥ ψ) d x = 3
ψ Ĥ(E ψ) d x = E 2
ψ ∗ ψ d3 x = E 2 (3.68)
−∞ −∞ −∞

Then the error in the Hamiltonian is σH = hH 2 i − hHi2 = E 2 − E 2 = 0 , Thus a separable solution
p

has the property that every measurement of the total energy is certain to return the value E which is the
total energy of the system. That is why the separation constant was denoted by E.

4. The general solution for a system of constant potential energy is a linear combination of
various separable solutions.
The time-independent Schrodinger’s equation yields an infinite collection of solutions ψ1 , ψ2 , ψ3 , . . .
which is associated with different total energy E1 , E2 , E3 , . . . ,. Thus all these wave functions satisfy
the Hamiltonian eigen value equation, that can be denoted by Ĥψn = En ψn where n runs from 1 to ∞.

26
Here the wave function ψn is called as the Eigen state of the system. We know that the time-dependent
Schrodinger’s equation has a property that any linear combination of the solutions is also a solution.

X
Ψ = c1 Ψ1 + c2 Ψ2 + c3 Ψ3 + · · · + cn Ψn + · · · = cn Ψn (3.69)
n=1

iEn
But we know that they all represent stationer states, hence we can write Ψn = ψn e− } t where En corre-
sponds to the total energy of the system whose wave function is given by ψn . So we get,

iE1 iE2 iE3 X iEn
− − −
Ψ = c1 ψ1 e }
t
+ c2 ψ2 e }
t
+ c3 ψ3 e }
t
+ ··· = cn ψn e− }
t
(3.70)
n=1

The constants c1 , c2 , c3 , . . . , cn , . . . are got by the boundary conditions of a system. The constant cn
represents the probability of finding the system whose total energy is En . To understand consider, some
nth state whose wave function is P sin , the probability density function is given by,
iEn iEn
Probability density = Ψn∗ Ψn = |Ψ |2 = c∗n ψn∗  t
e−} t
= |cn |2 ψn∗ ψn = |cn |2 |ψn |2 (3.71)
 

e} · cn ψn 

Thus the probability density is proportional to |cn |2 . Thus the probability of a system to have the wave
function Ψn is |cn |2 . Since it represents the probabilities, we should put a constrain that,

X
|cn |2 = 1 (3.72)
n=1

Since the total energy of the system is got by the Hamiltonian Eigen function,
∞ ∞
− iE}n t
X X
ĤΨ = EΨ where, Ψ = cn ψn e ⇒ hHi = |cn |2 En (3.73)
n=1 n=1

Some Theorems on Stationary states


Theorem 7. The total energy of a system whose solutions are separable (time-independent potential energy)
is a real quantity.

Proof. Let us assume that the total energy is a complex quantity. Thus, E = Er + iEi where Er is the
real component of energy and Ei is the imaginary component. Thus the total wave function becomes,
i(Er +iEi ) (iEr −Ei ) iEr Ei
Ψ = ψ e− }
t
= ψ e− }
t
= ψ e− }
t
e }
t
(3.74)
ˆ ∞
By the normalization condition, the wave function Ψ has to satisfy the condition Ψ ∗ Ψ d3 x = 1
−∞
ˆ ∞ iEr Ei iEr
 Ei
ψ∗  t t
e−} t e t
d3 x = 1 (3.75)
   
⇒ e} e } ψ }
−∞
ˆ ∞ 2Ei
⇒ ψ∗ψ e }
t
d3 x = 1 (3.76)
−∞
ˆ ∞
2Ei
⇒ e }
t
ψ ∗ ψ d3 x = 1 (3.77)
−∞
So we see that the normalization condition has a factor of time multiplied to it, which implies that the
normalization condition is not constant in time. But we have already seen in Eq.(3.7) that the normalization
condition doesn’t depend on time, thus the only way to satisfy that is by equating Ei to zero. This implies
that the imaginary component of total energy doesn’t exist and it is a purely real physical quantity.

27
}2 2
Theorem 8. The solution ψ to the time-independent Schrodinger’s equation − ∇ ψ + V ψ = E ψ can
2m
safely be assumed to be real or real solutions can be constructed easily from the complex solutions ψ and ψ ∗
Proof. Take the conjugate of the above Schrodinger’s equation which gives,
}2 2 ∗
− ∇ ψ + V ∗ψ∗ = E ∗ ψ∗ (3.78)
2m
Now V is the potential energy of the system which cant have a complex component hence V = v ∗ , then we
have seen in the previous theorem that E is a real quantity. Thus, the equation becomes,
}2 2 ∗
− ∇ ψ + V ψ∗ = E ψ∗ (3.79)
2m
Now if we compare the time-independent Schrodinger’s equation of ψ and ψ ∗ , they both are exactly the
same except that ψ → ψ ∗ , thus we can tell that ψ and ψ ∗ are both equally suitable solutions for the time-
independent Schrodinger’s equation. Thus even the linear combination is a suitable solution. Consider
ψ = ψr + iψi where ψr and ψi are real functions. Then ψ ∗ = ψr − iψi . So linear combination of ψ and ψ ∗
are equally valid solutions. So consider the linear combinations given by,

ψ + ψ ∗ = 2ψr and i(ψ − ψ ∗ ) = 2ψi (3.80)

So such linear combinations can construct real solutions for time-independent Schrodinger’s equation.

Theorem 9. If the potential energy of the system is an even function, i.e., V (−x, −y, −z) = V (x, y, z),
then the solutions for the Schrodinger equation can be constructed to be either odd or even.
Proof. Let us transform the time-independent Schrodinger’s equation by replacing all space coordinates
with their negatives. Thus it becomes,
" #
}2 ∂2 ∂2 ∂2
− + + ψ(−x, −y, −z) + V (−x, −y, −z)ψ(−x, −y, −z) = E ψ(−x, −y, −z)
2m ∂(−x)2 ∂(−y)2 ∂(−z)2

But the potential energy is an even function. And taking the −ve signs out from the differentials we get
(−1)2 which is nothing but 1 so the equation becomes,
" #
}2 ∂ 2 ∂2 ∂2
− + + ψ(−x, −y, −z) + V (x, y, z)ψ(−x, −y, −z) = E ψ(−x, −y, −z) (3.81)
2m ∂x2 ∂y 2 ∂z 2

Thus the above equation is same to the time-independent Schrodinger’s equation except ψ(x, y, z) is replaced
by ψ(−x, −y, −z). So, both the functions are valid solutions including their linear combination given by
ψ(x, y, z) + ψ(−x, −y, −z) which is even and ψ(x, y, z) − ψ(−x, −y, −z) which is odd.

Theorem 10. Total energy of the system has to be greater than the minimum value of its potential energy.
Proof. Consider the time-independent Schrodinger’s equation given by,
}2 2
− ∇ ψ + Vmin ψ = E ψ (3.82)
2m
Isolating the laplacian of ψ we rearrange to get,
2m
∇2 ψ = (Vmin − E)ψ (3.83)
}2
Now if the total energy is less than minimum potential energy, then both ψ and ∇2 ψ will have the same
sign. It implies that the value of ψ and ∇2 ψ increases or decreases indefinitely, which means that the
wave function is not normalizable. Thus for the wave function to be normalizable, the function ψ and its
laplacian ∇2 ψ should have opposite signs and that is possible only if E > Vmin

28
Chapter 4

Application of Schrodinger’s Theory


4.1 Infinite square well

Let us model a particle that is trapped in a 1D-region whose length is a. This is also called an Infinite
potential well where the potential of the system is of the form,
(
0 , 0<x<a
V = (4.1)
∞ , Otherwise
Since V = ∞ on the boundary and outside the well, we can say that the particle cannot cross the boundaries
of the well and the wave function of the particle is 0 outside and on the boundary. Since the potential
doesn’t depend on time we shall use the time-independent Schrodinger’s equation given by,
" #
}2 ∂ 2 ψ ∂ 2 ψ ∂ 2 ψ
− + + + V (x, y, z)ψ = E ψ (4.2)
2m ∂x2 ∂y 2 ∂z 2
Since V = 0 and as the system is one dimensional we get ψ → ψ(x) and the P.D.E becomes,
}2 ∂ 2 ψ ∂2ψ 2Em
− =Eψ ⇒ = ψ (4.3)
2m ∂x2 ∂x2 }2
2Em ∂ψ
Substituting, k2 = and rearranging we get, − k2 ψ = 0 (4.4)
}2 ∂x2
This is a simple P.D.E whose solution is of the form, ψ = A sin(kx) + B cos(kx) (4.5)
We can find the value of A and B by using the boundary and the normalization conditions on wave function.
So we have, ψ(0) = ψ(a) = 0. Thus we get ψ(0) = 0 = A sin(0) + B cos(0) ⇒ B = 0 and then,

ψ(a) = A sin(ka) = 0 ⇒ ka = nπ ⇒ k= (4.6)
a
 nπx 
Thus we got ψ(x) = A sin . To get the value of A, we can use the normalization condition.
a
ˆ ∞ ˆ ∞  nπx 

ψ ψ dx = 1 ⇒ A2 sin2 =1 (4.7)
−∞ −∞ a
The integral
r can be solved by using the identity 2 sin (x) = 1 − cos(2x) . Thus we get the value of A to
2

2
be A = and the fully solved wave function for a particle in an infinite square well is,
a
r
2  nπx 
ψn (x) = sin (4.8)
a a
As said earlier, we have infinite solutions for the system where n has the range 1, 2, 3, . . . , ∞. So the different
solutions corresponding to the value of n, denoted by ψn .

29
Properties of the particle in Infinite square well
1. Wave function
The wave function is even and odd alternatively with respect to the center of the square well. i.e., the
particle in the odd stationary state described by the wave function ψ2n−1 is an even function with respect
to the center of the well and the particle in the even stationary state described by the wave function ψ2n is
an odd function with respect to the center of the well.

2. Probability density
The probability density of the particle in a particular stationary state n is given by ψn∗ ψn which becomes,
a  nπx 
Probability density = ψn∗ ψn = |ψn |2 = sin2 (4.9)
2 a

3. Energy Eigen values


The total energy of a system is got by the Hamiltonian operator Ĥ. In time-independent system it is,
}2 ∂
Ĥ = − (4.10)
2m ∂x2
Operating Ĥ on the stationary states ψn we get,
r ! r
}2 ∂ψn }2 ∂ 2  nπx  }2 n2 π 2 2  nπx  n2 π 2 }2
Ĥψn = − = − sin = − × − sin = ψn (4.11)
2m ∂x2 2m ∂x2 a a 2m a2 a a 2ma2

n2 π 2 }2
⇒ Ĥψn = ψn (4.12)
2ma2
n2 π 2 }2
Thus the Eigen values of Ĥ are the energy Eigen values En which is given by, En = (4.13)
2ma2
Here is a pictorial representation of the wave function ψ, its probability density and energy Eigen values.

Nodes and their relation with the position


Nodes are the regions in space where the probability density of the particle is zero where the probability
of finding the particle regions is zero. A particle in 1st state has has no nodes, a particle in 2nd state has
a a 2a
1 node at x = , and a particle in 3rd state has 2 nodes at x = and x = , similarly, a particle at nth
2 3 3
a 2a 3a (n − 1)a
energy state has n − 1 number of nodes at x = , , , . . . , .
n n n n

30
Momentum of the particle
The momentum of the particle is given by the momentum operator P̂x when operated r on an Eigen wave
a  nπx 
function gives the Eigen values of momentum. But the wave function given by ψn = sin is not
2 a
an Eigen function as the form of the function changes. But we can modify the wave function ψn by using
the identity sin(x) = −i sinh(ix) then the wave function becomes,
inπx inπx r inπ r −inπ
a e a − e− a
r
i a x i a x
ψn = −i =− e a + e a = ψn+ + ψn− (4.14)
2 2 2 2 2 2
Thus the wave function is divided into two the positive part and the negative part which are the Eigen
functions for the momentum operator. Then operating ψn+ and ψn− on P̂x we get,
r inπ  r inπ 
∂ψn+ ∂  i a x }nπ  i a x }nπ +
+
P̂x ψn = −i} = −i} − e a = − e a = ψ (4.15)
∂x ∂x 2 2 a 2 2 a n
r −inπ  r −inπ 
− ∂ψn− ∂ i a x }nπ  i a x }nπ −
P̂x ψn = −i} = −i} e a = e a = ψ (4.16)
∂x ∂x 2 2 a 2 2 a n
n}π nh h
Thus the Eigen values of both ψn+ and ψn− are the same and are p−
n = = = }k = (4.17)
a 2a λ

Expectation value of momentum


The expectation value of momentum is given by hPx i = hψn |P̂x ψn i where the integral goes from one
boundary to the other. So we get,
ˆ a ˆ a  ˆ a r
∗ ∗ ∂ψn  ∗ ∂ a  nπx 
hPx i = hψn |P̂x ψi = ψn (P̂x ψn )dx = ψ − i} dx = −i} ψn sin dx (4.18)
0 0 ∂x 0 ∂x 2 a
ˆ ˆ
i}nπ a  nπx   nπx  i}nπ a  2nπx 
⇒ hPx i = − sin cos dx = − sin dx (4.19)
2 0 a a 4 0 a
i}nπ a h  2nπx ia i}a  i}a
⇒ hPx i = × cos = cos(2nπ) − cos(0) = (1 − 1) = 0 (4.20)
4 2nπ a 0 8 8

Mutual Orthonormality of the wave functions


Two wave functions are said to be orthogonal if their inner product is zero, i.e., hψm |ψn i = 0. So, let’s
check whether the stationary states of a particle in an infinite square well are mutually orthogonal or not.
ˆ ar ˆ
a  mπx r a  nπx  1 a  mπx   nπx 
hψm |ψn i = sin sin dx = 2 sin sin dx (4.21)
0 2 a 2 a a 0 a a
Using the trigonometric identity of 2 sin(A) sin(B) = cos(A − B) − cos(A + B) we get,
ˆ a
1  πx   πx 
hψm |ψn i = cos (m − n) − cos (m + n) dx (4.22)
a 0 a a
" #a
1 a 1  πx  1  πx 
hψm |ψn i = × sin (m − n) − sin (m + n) (4.23)
a π (m − n) a (m + n) a
0
"   #
1 sin (m − n)π sin (m + n)π
hψm |ψn i = − (4.24)
π (m − n) (m + n)
We know that m and n are integers, thus m − n and m + n are also integers, thus the values of sines will
become zero and the inner product will become zero. And this will work only if m 6= n. If n = m then it is
just the normalization condition where the inner product becomes 1. Thus for the stationary solutions of
the particle in infinite square well, we can write,
ˆ (
∗ 1, m=n
hψm |ψn i = ψm ψn dx = δmn = ⇒ ψm and ψn are mutually orthonormal. (4.25)
0 , m 6= n

31
4.2 Quantum Harmonic Oscillator
This is a system where the quantum particles are in oscillatory motion. Classically, this can be explained by
Hooke’s law where the force on the particle is directly proportional to the displacement and acts opposite
to the direction of displacement ‘F = −kx’. Similarly, in the quantum harmonic oscillator the potential is,
1
V (x) = mω 2 x2 (4.26)
2
where m is the mass and ω is the angular frequency of the oscillatory motion exhibited by the particle.

In such potential, Schrodinger equation becomes,


" #
}2 ∂ 2 ψ 1 ∂2ψ 2Em m2 ω 2 x2
− + mω 2 x2 ψ = Eψ rearranging we get, + − ψ=0 (4.27)
2m ∂x2 2 ∂x2 }2 }2
r r r
mω ∂ξ mω ∂ψ ∂ψ ∂ξ mω ∂ψ
Substituting ξ = x , we get, = , and = = (4.28)
} ∂x } ∂x ∂ξ ∂x } ∂ξ
! r ! r ! r !
∂2ψ mω ∂ 2 ψ
r
∂ ∂ψ mω ∂ ∂ψ mω ∂ ∂ψ mω ∂ mω ∂ψ
= = = = = (4.29)
∂x2 ∂x ∂x } ∂x ∂ξ } ∂ξ ∂x } ∂ξ } ∂ξ } ∂ξ 2
Using all these substitutions and rearranging them, we get,
" #
mω ∂ 2 ψ 2Em m2 ω 2  ξ 2 }  ∂ 2 ψ h 2E i
+ − ψ=0 ⇒ + − ξ 2
ψ=0 (4.30)
} ∂ξ 2 }2 }2 mω ∂ξ 2 }ω

2E 1
If we substitute ε = where we measure energy in units of }ω then it can be expressed in terms of ε.
}ω 2
∂2ψ 2
(4.31)

+ ε − ξ ψ=0
∂ξ 2
1 2
The un-normalized asymptotic solution is of the form ψ = H(ξ)e− 2 ξ where the solution is expressed as a
power series H(ξ). Then using this in the modified Schrodinger’s equations we get the P.D.E to be,

∂ 2 H(ξ) ∂H( ξ)
2
− 2ξ + (λ − 1)H(ξ) = 0 (4.32)
∂ξ ∂ξ
2 ∂ n −ξ2 
The solutions are Hermite polynomials of degree n defined by, Hn (ξ) = (−1)n eξ e (4.33)
∂ξ n
Then by using the normalization condition hψn |ψn i = 1 we get the normalized solution to be,
 mω  1 1  r
mω  − mω x2
4
ψn = √ Hn x e 2} (4.34)
π} 2n n! }

32
First few solutions for the wave functions of a quantum harmonic oscillator are given by,
 mω  1 mω 2
e− 2} x
4
ψ0 = (4.35)
π}
 mω  1 mω  − mω x2 √  mω  34 − mω x2
r
4 1 
ψ1 = √ 2x e 2} = 2π xe 2} (4.36)
π} 2 } π}
! !
 mω  1 1  r mω 2 mω 2
 mω  5 √ } mω 2
− 2 e− 2} x = e− 2} x
4 4
ψ2 = √ 4 x 2π 2 x2 − (4.37)
π} 8 } π} 2mω
! r !
 mω  1 1  r mω 3  r mω   mω  13 4π 3 3}
− mω 2 mω 2
x e− 2} x (4.38)
4 x 4 3
ψ3 = √ 8 x − 12 x e 2} = x −
π} 48 } } π} 3 2mω

Properties of Harmonic Oscillator


1. Energy Eigen values
The Eigen values of the total energy for a quantum harmonic oscillator at nth state is given to be,
 1
En = n + }ω n = 0, 1, 2, 3 . . . (4.39)
2
Thus we can note that even at 0th energy state the total energy of the quantum harmonic oscillator is not
zero. And is a non-zero quantity given by E0 = 12 }ω. So, this minimum energy is a significant physical result
since it predicts that in any system that behaves as a quantum oscillator, the total energy cannot be zero.
So, atoms and molecules vibrate even at the lowest state or at 0 K temperatures because of this minimum
energy. Such vibrations are referred to as ‘Zero point vibrations’ whose energy is ‘Zero point energy’. Also,
we can note that the difference in energy between any two states is given to be ∆E = En+1 − En = }ω
which is a constant for a system. Thus, the energy levels are placed at uniform distances.

2. Wave function and Probability Density


• The wave function of a harmonic oscillator is real and is of the form ψ0 , ψ1 , ψ2 , . . . where they are
alternatively even and odd functions with respect to the origin.
• The probability density is got by squaring the wave function and there are ‘n − 1’ number of nodes
(zero probability region) for a harmonic oscillator in nth energy state.
• All even energy states have a maximum probability near the origin and for all odd energy states, the
probability of the particle being in origin is Zero. Also, there are n probability lobes for a particle
that is in nth energy state.
• We also can see that the wave function can come out of the potential energy barrier with a very low
probability. This phenomenon is quantum tunneling which is unique to quantum mechanics.

33
4.2.1 Quantum Harmonic Oscillator - Analytical Method
The Classical analogue of harmonic oscillator whose mass is m and angular frequency is ω is governed by,

p2 1 1 2
H= + mω 2 x2 = (p + m2 ω 2 x2 ) (4.40)
2m 2 2m
In classical mechanics, both position(x) and momentum(p) can simultaneously be zero and in turn the
energy of the classical oscillator will be zero at x = p = 0. But that is not the case in quantum realm.
Here we should write the equation in the form of operators as all observables in quantum mechanics are
described by operators.. So, we use H → Ĥ, p → P̂, and x → x̂. Then we get,

P̂2 1 1
+ mω 2 x̂2 = P̂2 + (mωx̂)2 (4.41)

Ĥ =
2m 2 2m
Here both Position and Momentum cannot be zero as it violates Uncertainty principle. So there can’t be
a quantum state with zero energy. Now let us define two operators as,
1 1
and (4.42)
 
a= √ P̂ + imωx̂ a= √ P̂ − imωx̂
2mω 2mω
Let us also find the commutator of these operators,
1 1  
(4.43)
  
[a, a] = [P̂ + imωx̂, P̂ − imωx̂] = P̂ + imωx̂ P̂ − imωx̂ − P̂ − imωx̂ P̂ + imωx̂
2mω 2mω
1  ˆ2 ˆ
 1
= 2 2 2 2 2 2 2
P + m ω x̂ + imω[x̂, P̂] − P − m ω x̂ + imω[x̂, P̂] = · 2imω[x̂, P̂] (4.44)
2mω 2mω
⇒ [a, a] = i [x̂, P̂] = i · i = −1 ⇒ aa − aa = −1 ⇒ aa + 1 = aa (4.45)
| {z }
=i
So in this notation we can write the Hamiltonian in a simplified manner. Consider,
1 1 1
P̂2 + m2 ω 2 x̂2 + imω[x̂, P̂] = P̂2 + m2 ω 2 x̂2 − mω
   
aa = P̂ + imωx̂ P̂ − imωx̂ =
2mω 2mω 2mω
1 ω  1   1
P̂2 + (mωx̂)2 = Ĥ = ωaa + = ω aa + (4.46)

⇒ ⇒ Ĥ = ω aa +
2m 2 2 2
1
The potential energy of the system is V = mω 2 x2 . So in such time independent potentials we get the
2
Schrodinger’s equation to be in form of an Eigen value equation given by Ĥ|Ψ i = E|Ψ i.
 1  a  1
Consider, Ĥ(a|Ψ i) = ω aa + (a|Ψ i) = ω aaa + |Ψ i = ωa aa + |Ψ i (4.47)
2 2 2
" #
 1  1 h i h i
= ωa aa + 1 + |Ψ i = a ω aa + + ω |Ψ i = a Ĥ + ω |Ψ i = a Ĥ|Ψ i + ω|Ψ i (4.48)
| {z } 2 2
by Eq.(4.45)
h i
= a E|Ψ i + ω|Ψ i = (E + ω)a|Ψ i ⇒ Ĥ(a|Ψ i) = (E + ω)a|Ψ i (4.49)
Thus we see that the operator a has just increased the energy of the oscillator by a factor of ω. Then
consider the operation given by Ĥ(a|Ψ i) in the similar manner, we get,
 1  a  ω
Ĥ(a|Ψ i) = ω aa + a|Ψ i = ω (aa − 1)a + |Ψ i = a ω(aa − 1) + |Ψ i = a(E − ω)|Ψ i
2 2 2
Thus we see that the operator a has just decreased the energy of the oscillator by a factor of ω. Hence the
operators a and a are called as the ‘raising’ and the ‘lowering’ operators respectively. Because their work
is to just increase or lower energy state of the wave function. But a cannot indefinitely keep on lowering
the energy state. It can lower the energy state only till ground state with some non-negative energy. So if
a operate on the ground state (lowermost) wave function (|Ψ0 i) then it has to yield zero.

34
1 i  ∂|Ψ0 i 
(4.50)

∴ a|Ψ0 i = 0 ⇒ a|Ψ0 i = √
P̂ − imωx̂ |Ψ0 i = √ − − mωx|Ψ0 i = 0
2mω 2mω ∂x
ˆ ˆ
d|Ψ0 i d|Ψ0 i 1 2
⇒ = −mωx|Ψ0 i ⇒ = − mωx dx ⇒ |Ψ0 i = N e− 2 mωx (4.51)
dx |Ψ0 i
To find the normalization constant N we need to use the condition hΨ0 |Ψ0 i = 1,
ˆ ∞ ˆ ∞  mω  1
2
N 2 e−mωx dx = 1
4
2
|Ψ0 i dx = 1 ⇒ when solved we get, N= (4.52)
∞ ∞ π
 mω  1 1 2
 mω  1 mω 2
e− 2 mωx e− 2} x
4 4
⇒ |Ψ0 i = In SI units we get, |Ψ0 i = (4.53)
π π}
To get its Energy state we shall use the equation Ĥ|Ψ0 i = E0 |Ψ0 i. So we get,
 1  1  ω ω
Ĥ|Ψ0 i = ω aa + |Ψ0 i = ω a a|Ψ0 i + |Ψ0 i = |Ψ0 i ⇒ E0 = (4.54)
2 | {z } 2 2 2
=0

Thus the ground state energy of the system is given by E0 = ω/2 and in SI units we get, E0 =
2
We know that the raising operator increases the energy state by one i.e., a|Ψ0 i → |Ψ1 i. So, to get the First
excited state Ψ1 we just have to use the raising operator once on |Ψ0 i. And hence we get,
√  mω  43 − 1 mωx2

P̂ + imωx̂ |Ψ0 i 1  ∂ 
|Ψ1 i = a|Ψ0 i = √ =√ −i + imωx̂ |Ψ0 i ⇒ Ψ1 = 2π xe 2
2mω 2mω ∂x π}
1
Thus the general formula to get the wave function at nth state is Ψn = √ (a)n |Ψ0 i (4.55)
n!
Now we have, a|Ψ0 i = 0|Ψ0 i . Consider the operation of aa on the vector a|Ψ0 i ,
 

(4.56)
   
aa a|Ψ0 i = a(aa)|Ψ0 i = a (aa + 1) |Ψ i = a(a a|Ψ0 i +|Ψ0 i) = 1 a|Ψ0 i
| {z } | {z }
by Eq.(4.45) =0

⇒ aa|Ψn i = n|Ψn i (4.57)

Similarly, aa|Ψn i = (aa + 1)|Ψn i = (n + 1)|Ψn i ⇒ aa|Ψn i = (n + 1)|Ψn i (4.58)


Hence, |Ψn i is an eigen vector of the operators aa and aa with eigen value n and n + 1 respectively.
 1  |Ψn i 
The energy of the oscillator at nth state is, Ĥ|Ψn i = ω aa + |Ψn i = ω aa|Ψn i + (4.59)
2 2
 |Ψn i   1  1
⇒ Ĥ|Ψn i = ω n|Ψn i + =ω n+ |Ψn i ⇒ En = n + ω (4.60)
2 2 2
We know that when a acts on |Ψn i, the state is transformed to |Ψn+1 i. And when a acts on |Ψn i, the state
is transformed to |Ψn−1 i So, we get,
a|Ψn i = λn |Ψn+1 i and a|Ψn i = κn |Ψn−1 i (4.61)
Let us write the corresponding hermitian conjugate of the above equation,
hΨn |a = λ∗n hΨn+1 | and hΨn |a = κ∗n hΨn−1 | (4.62)
Taking their inner product, we get,
hΨn |aa|Ψn i = λ∗n λn hΨn+1 |Ψn+1 i and hΨn |aa|Ψn i = κ∗n κn hΨn−1 |Ψn−1 i (4.63)

Solving LHS, hΨn |aa|Ψn i = hΨn |(aa + 1)|Ψn i = (n + 1)hΨn |Ψn i and hΨn |aa|Ψn i = nhΨn |Ψn i

Since the wave functions are normalized we get hΨn |Ψn i = hΨn+1 |Ψn+1 i = hΨn−1 |Ψn−1 i = 1. So,

√ √
|λn |2 = n + 1 ⇒ a|Ψn i = n + 1|Ψn+1 i and |κn |2 = n ⇒ a|Ψn i = n|Ψn+1 i (4.64)

35
4.3 The Hydrogen Atom
The potential energy of the interaction between the electron and the proton is given by,
−e2
V = (4.65)
4πε0 r
Where e is the charge of the electron, ε0 is the permittivity of free space, and r is the distance between
the proton and the electron. We see that the potential energy depends only on the radial distance, so the
system has spherical symmetry where a sphere of radius r is an equipotential surface. So if we use the
spherical polar coordinates then it will be easy to tackle the system.
Z x = r sin(θ) cos(ϕ)
z y = r sin(θ) sin(ϕ)
z = r cos(θ)

P (x, y, z)
(r, θ, ϕ)
r
θ
y
rs
ϕ in(
θ)
Y
x
X
}2 2
The time-independent Schrodinger’s equation is given by, −
∇ ψ + V ψ = Eψ (4.66)

Where µ is the reduced mass of the system. Then the Laplacian operator in spherical polar coordinates is,
1 ∂  2∂  1 ∂  ∂  1 ∂2
∇2 = 2 r + 2 sin(θ) + 2 2 (4.67)
r ∂r ∂r r sin(θ) ∂θ ∂θ r sin (θ) ∂ϕ2
Thus the Schrodinger equation in spherical polar coordinate becomes,
" #
}2 1 ∂  2 ∂ψ  1 ∂  ∂ψ  1 ∂2ψ
− r + 2 sin(θ) + 2 2 + V ψ = Eψ (4.68)
2µ r2 ∂r ∂r r sin(θ) ∂θ ∂θ r sin (θ) ∂ϕ2

Solving this equation is pretty time-consuming, and the technique of separations of variables has to be used.
Let the wave function ψ(r, θ, ϕ) be separable and has the form,
ψ(r, θ, ϕ) = R(r) Θ(θ) Φ(ϕ) = R Θ Φ (4.69)
Now we shall substitute this expression of the wave function in Schrodinger’s equation then it becomes,
" #
}2 1 ∂  2 ∂  1 ∂  ∂  1 ∂2  
− 2
r RΘΦ + 2 sin(θ) RΘΦ + 2 2 2
R Θ Φ +(V −E) R Θ Φ = 0
2µ r ∂r ∂r r sin(θ) ∂θ ∂θ r sin (θ) ∂ϕ
" #
}2 Θ Φ ∂  2 ∂R  RΦ ∂  ∂Θ  R Θ ∂2Φ
(4.70)

− 2
r + 2 sin(θ) + 2 2 2
+ (V − E) R Θ Φ = 0
2µ r ∂r ∂r r sin(θ) ∂θ ∂θ r sin (θ) ∂ϕ
Then dividing the equation throughout by R Θ Φ we get,
" #
}2 1 ∂  2 ∂R  1 ∂  ∂Θ  1 ∂2Φ
− r + sin(θ) + + (V − E) = 0 (4.71)
2µ Rr2 ∂r ∂r Θr2 sin(θ) ∂θ ∂θ Φr2 sin2 (θ) ∂ϕ2

2µr2
Multiplying throughout by − we get,
}2
" #
1 ∂  2 ∂R  1 ∂  ∂Θ  1 ∂2Φ 2µr2
r + sin(θ) + + (E − V ) =0 (4.72)
R ∂r ∂r Θ sin(θ) ∂θ ∂θ Φ sin2 (θ) ∂ϕ2 }2

36
Rearranging the equation by putting the radial terms on the Left-hand side of the equation and the angular
terms on another side of the equation we get,
1 ∂  2 ∂R  2µr2 1 ∂  ∂Θ  1 ∂2Φ
r + (E − V ) 2 = − sin(θ) − (4.73)
R ∂r ∂r } Θ sin(θ) ∂θ ∂θ Φ sin2 (θ) ∂ϕ2
So now we see that when we vary the angular variables θ and ϕ, as LHS doesn’t depend on them at all, we
can tell that the RHS behaves as a constant for the LHS and similarly when the radial terms are varied,
the RHS of the equation doesn’t change and hence LHS behaves as a constant with respect to the RHS
thus we can separate the equations by using a separation constant l(l + 1). Thus, we get two equations for
the angular and radial parts of the wave function to be,
1 ∂  2 ∂R  2µr2
r + (E − V ) 2 = l(l + 1) (4.74)
R ∂r ∂r }
1 ∂  ∂Θ  1 ∂2Φ
− sin(θ) − = l(l + 1) (4.75)
Θ sin(θ) ∂θ ∂θ Φ sin2 (θ) ∂ϕ2
The Angular equation Eq.(4.75) further can be separated in terms of Θ and Φ. So, multiply the equation
by sin2 (θ) and rearranging the equation by putting θ terms on one side and ϕ terms on another side of the
equation, it becomes,
sin(θ) ∂  ∂Θ  1 ∂2Φ
sin(θ) + l(l + 1) sin2 (θ) = − (4.76)
Θ ∂θ ∂θ Φ ∂ϕ2
So, the LHS and RHS are now separable, thus the separating the equations by using a separation constant
m2 we get,
sin(θ) ∂  ∂Θ  1 ∂2Φ
sin(θ) + l(l + 1) sin2 (θ) = m2 and − = m2 (4.77)
Θ ∂θ ∂θ Φ ∂ϕ2
Thus the three separated P.D.E. for the wave function of the hydrogen atom is given by,

∂  2 ∂R  2µr2 R
r + (E − V ) = l(l + 1)R (4.78)
∂r ∂r }2
∂  ∂Θ 
sin(θ) sin(θ) + l(l + 1) sin2 (θ)Θ = m2 Θ (4.79)
∂θ ∂θ

∂2Φ
= m2 Φ (4.80)
∂ϕ2

The Φ Equation
The partial differentiation equation of Φ is a simple one given by,
∂2Φ
− m2 Φ = 0 (4.81)
∂ϕ2
The solution is of the form Φ = A eimϕ where A is a constant. Now we also know that the angle ϕ has the
limits of 0 ≤ ϕ ≤ 2π and Φ(ϕ + 2π) = Φ(ϕ). Thus we get Ae2imπ = 1 = Ae2iπ . Thus for this to be valid,
m has to be an integer. So, m = 0, ±1, ±2, ±3, . . . . To get the value of A we need to normalize the Φ
solution within the limits of 0 to 2π. Thus we get, hΦ|Φi = 1,
ˆ 2π ˆ 2π ˆ 2π
∗ −imϕ
Φ Φ dϕ = 1 ⇒ Ae Ae imϕ
dϕ = 1 ⇒ A2 dϕ = 1 (4.82)
0 0 0
ˆ 2π  2π 1
⇒ A2 dϕ = A2 ϕ 0 = A · 2π = 1 ⇒ A= √ (4.83)
0 2π
Thus the normalized Φ equation is given to be,
1
Φ = √ eimϕ (m = 0, ±1, ±2, ±3, . . . ) (4.84)

37
The Θ Equation
The P.D.E for the Θ function is given in Eq.(4.79) which is,
∂  ∂Θ 
sin(θ) sin(θ) + l(l + 1) sin2 (θ)Θ = m2 Θ (4.85)
∂θ ∂θ
Substituting z = cos(θ) in the above P.D.E., we get,
∂z p ∂Θ ∂Θ ∂z p ∂Θ
z = cos(θ) ⇒ = − sin(θ) = − 1 − z 2 ⇒ = = − 1 − z2 (4.86)
∂θ ∂θ ∂z ∂θ ∂z
∂  ∂Θ  ∂  p p  ∂Θ  ∂   ∂Θ  ∂z ∂   ∂Θ 
1 − z2 1 − z2

sin(θ) = 1 − z2 − 1 − z2 =− = sin(θ)
∂θ ∂θ ∂θ ∂z ∂z ∂z ∂θ ∂z ∂z
∂  ∂Θ  h ∂ Θ2 ∂Θ i
= 1 − z2 1 − z2 (4.87)
 
⇒ sin(θ) sin(θ) − 2z
∂θ ∂θ ∂z 2 ∂z
Thus the P.D.E. becomes,
h  ∂2Θ ∂Θ i h i
1 − z2 1 − z2 2 2
(4.88)

− 2z + l(l + 1) 1 − z − m Θ=0
∂z 2 ∂z
Dividing throughout by 1 − z 2 we get,


 ∂2Θ ∂Θ h m2 i
1 − z2 − 2z + l(l + 1) − Θ=0 (4.89)
∂z 2 ∂z 1 − z2
This is a Legendre equation whose solutions are associated Legendre polynomials P lm (z). Thus
Θ = AP lm (z) (4.90)
where A is the normalization constant. And the Associated Legendre polynomials are given by,
|m|
(1 − z 2 ) 2 d l+|m|  2 l
P lm (z) = z − 1 (4.91)
2l l! dz l+|m|
s (
(2l + 1)(l − |m|)! 1, m≤0
The Normalization constant is, A= where  = m (4.92)
2(l + |m|)! (−1) , m > 0
Thus the normalized Θ equation is given by,
s
(2l + 1)(l − |m|)! sin|m| (θ) d l+|m|  l
Θm
l = cos 2
(θ) − 1 (4.93)
2(l + |m|)! 2l l! d cos(θ) l+|m|
For the solution to be valid |m| ≤ l. Thus m = 0, ±1, ±2, ±3, . . . , ±l. Few solutions of Θm
l are given by,

1
Θ00 = √
2
r r r
3 3 3
Θ01 = cos(θ) Θ11 =− sin(θ) Θ−1
1 = sin(θ)
2 4 4
r r r
5 15 15
Θ02 = 3 cos2 (θ) − 1 Θ12 = − Θ−1

sin(θ) cos(θ) 2 = sin(θ) cos(θ)
4 4 4
r r
15 15
Θ22 = − sin2 (θ) Θ−2
2 = sin2 (θ)
16 16
r
0 7
5 cos3 (θ) − 3 cos(θ)

Θ3 =
8
r r
21 21
Θ13 = − sin(θ) 5 cos2 (θ) − 1 Θ−1 sin(θ) 5 cos2 (θ) − 1
 
3 =
16 16
r r
105 105
Θ23 = − sin2 (θ) cos(θ) Θ−2
3 = sin2 (θ) cos(θ)
16 16
r r
35 35
3
Θ3 = − sin3 (θ) Θ−3
3 = sin3 (θ) Note that Θm −m
l = −Θl
32 32

38
Spherical Harmonics
The combined equation of Θ and Φ is known as the Spherical Harmonics Y (θ, ϕ) = Θ(θ) Φ(ϕ) given by,
s
(2l + 1)(l − |m|)! eimϕ sin|m| (θ) d l+|m|  l
Ylm (θ, ϕ) =  cos 2
(θ) − 1 (4.94)
4π(l + |m|)! 2l l! d cos(θ) l+|m|

Solutions for Spherical harmonics are generally complex (when m 6= 0) as they have the term eimϕ multiplied
with them. So to get the real solutions we employ the advantage of the linear combination of the wave
functions. Now we have,
1 1
Ylm = √ Θm imϕ
= √ Θm (4.95)

l e l cos(mϕ) + i sin(mϕ)
2π 2π
1 −1
Yl−m = √ Θ−m e−imϕ = √ Θm (4.96)

l l cos(mϕ) − i sin(mϕ)
2π 2π
which are complex. But we can choose the linear combinations given by,

Ylm − Yl−m 1
= √ Θml cos(mϕ) (4.97)
2 2π

Ylm + Yl−m 1
= √ Θm l sin(mϕ) (4.98)
2i 2π
These linear combinations yield real wave functions. A few spherical harmonics functions are given below,

l = 0, 1 and m = −1, 0, 1
r
1 3
Y00 = Y10 = cos(θ)
2π 4π
r r
3 3
Y11 = − sin(θ) cos(ϕ) Y1−1 = sin(θ) sin(ϕ)
8π 8π

l = 2 and m = −2, −1, 0, 1, 2

r
5
Y20 3 cos2 (θ) − 1

=

r r
15 15
Y21 = − sin(θ) cos(θ) cos ϕ Y2−1 = sin(θ) cos(θ) sin(ϕ)
8π 8π
r r
15 15
Y22 = − sin2 (θ) cos(2ϕ) Y2−2 = sin2 (θ) sin(2ϕ)
32π 32π

l = 3 and m = −3, −2, −1, 0, 1, 2, 3

r
7
Y30 5 cos3 (θ) − 3 cos(θ)

=
16π
r r
21 21
Y31 = − sin(θ) 5 cos2 (θ) − 1 cos(ϕ) Y3−1 sin(θ) 5 cos2 (θ) − 1 sin(ϕ)
 
=
32π 32π
r r
105 105
Y32 = − sin2 (θ) cos(θ) cos(2ϕ) Y3−2 = sin2 (θ) cos(θ) sin(2ϕ)
32π 32π
r r
35 35
3
Y3 = − sin3 (θ) cos(3ϕ) Y3−3 = sin3 (θ) sin(3ϕ)
64π 64π

39
The Radial (R) equation
∂  2 ∂R  2µr2 R
The Radial P.D.E. is given by the Eq.(4.78) as, r + (E − V ) = l(l + 1)R (4.99)
∂r ∂r }2
χ
Let us simplify this equation by using suitable substitutions. Using R = we get,
r
∂χ
∂R R −χ ∂R ∂χ ∂  2 ∂R  ∂ 2 χ ∂χ ∂χ ∂2χ
= ∂r 2 ⇒ r2 =R −χ ⇒ r =r 2 +  −  =r 2 (4.100)
∂r r ∂r ∂r ∂r ∂r ∂r ∂r ∂r ∂r
Thus the radial P.D.E becomes,
∂2χ h 2µr2 iχ
r 2 + (E − V ) 2 − l(l + 1) =0 (4.101)
∂r } r
}2
Multiplying throughout by − and separating E from the equation,
2µr
}2 ∂ 2 χ h }2 l(l + 1) i
− + V + χ = Eχ (4.102)
2µ ∂r2 2µ r2
So here we see that the above equation is similar to Schrodinger’s equation except that the potential term
is having an extra factor added to it thus the effective potential energy becomes,
}2 l(l + 1)
Veff = V + (4.103)
2µ r2
| {z }
centrifugal term

The centrifugal term tends to throw the particle away from the origin which depends inversely on the
reduced mass of the system. Notice that for a system whose potential energy depends only on the radial
distance i.e. V → V (r) the spherical harmonic wave functions and the radial P.D.E remain the same till
the previous equation. The potential energy of Hydrogen atom is given by,
e2
V (r) = − (4.104)
4πε0 r
Now using the potential energy expression of the hydrogen atom in the P.D.E., we get,

}2 ∂ 2 χ h e2 }2 l(l + 1) i
− + − + χ = Eχ (4.105)
2µ ∂r2 4πε0 r 2µ r2

40
Dividing throughout by E and taking χ common we get,
}2 ∂ 2 χ h e2 }2 l(l + 1) i
− = 1 + − χ (4.106)
2µE ∂r2 4πε0 rE 2µE r2
r
−2µE
Let us use a substitution κ = we get,
}2
1 ∂2χ h e2 µ l(l + 1) i
= 1 − + χ (4.107)
κ2 ∂r2 2π}2 ε0 κ2 r (κr)2
µe2
Now if we substitute ρ = κr and ρ0 = then the equation reduces to,
2πε0 }2 κ
∂2χ h ρ0 l(l + 1) i
= 1 − + χ (4.108)
∂ρ2 ρ ρ2
Now, let’s review the substitutions and the P.D.E. The radial equation is reduced to the below equation,
∂2χ h ρ0 l(l + 1) i
= 1 − + χ (4.109)
∂ρ2 ρ ρ2
q
−2µE
where, χ = rR(r) is the modified radial wave function. Then, ρ = κr where κ = }2
and ρ is the
independent variable instead of r. For bound states the total energy is a negative quantity, hence κ , is real.
Let us now find the solution for the above equation. Taking the asymptotic solution as (ρ → ∞) we get,
∂2χ
=χ ⇒ χ = Ae−ρ + B |{z}

∂ρ2
not valid

Thus asymptotic solution is χ = Ae−ρ . But as ρ → 0 the centrifugal term dominates so,
∂2χ l(l + 1)
= χ ⇒ χ = Cρ−l + D ρl+1 (4.110)
∂ρ2 ρ2 |{z}
not valid

Thus the total solution will contain a power series and be in the form,

X
χ(ρ) = N (2ρ)−l e−ρ L(2ρ) where L(2ρ) = cj (2ρ)j (4.111)
j=0

When the power series is solved through the recursion formula we get certain conditions like, as j can’t go
to infinity, there has to be a maximum +ve integer n such that,

n = jmax + l + 1 and ρ0 = 2n (4.112)

Thus the power series is called as the associated Laguerre polynomials Ln−(l+1)
2l+1
(x) whose degree is n−(l+1).
" #
2l+1 n−(l+1) 
2l+1 d x d

−x n−(l+1)
2l+1
Ln−(l+1) (x) = (−1) e e x (4.113)
dx2l+1 dxn−(l+1)

Few associated Laguerre polynomials are given below,

n = 1, l = 0 ⇒ L10 (x) = 1
n = 2, l = 0 ⇒ L11 (x) = −2x + 4
n = 2, l = 1 ⇒ L30 (x) = 6
n = 3, l = 0 ⇒ L12 (x) = 3x2 − 18x + 18
n = 3, l = 1 ⇒ L31 (x) = −24x + 96
n = 3, l = 2 ⇒ L50 (x) = 120

41
The normalization constant N is got by the condition hχ|χi = 1 which is,
v !3
u
u µe2 (n − (l + 1))!
N =t 2 3 (4.114)
2nπε0 } 2n (n + l)!
For the wave function to be normalizable n ≥ (l + 1), Thus we get l = 1, 2, 3, . . . (n − 1). Since we got the
solution we shall undo the substitutions one by one. Notice that the condition for the radial wave function
solution was ρ0 = 2n i.e.,
s
µe2 µe2 }2 µ2 e4 }2
2n = = − squaring both sides 4n 2
= − · (4.115)
2πε0 }2 κ 2πε0 }2 2µE 4π 2 ε20 }4 2µE

µe4 h µe4 13.605


⇒ E=− using } = we get E=− ⇒ En = eV (4.116)
32π 2 ε20 }2 n2 2π 8ε20 h2 n2 n2
This is nothing but the Bohr’s formula in Eq.(1.11) for energy which was purely derived from classical
notion of particle and assuming that angular momentum is integral multiple of }. Thus Bohr’s energy
values are the Eigen values of Total Energy of the hydrogen atom. Also from the relation ρ = rκ we get,
r s
−2µE 2µ µe4
ρ = rκ = r = r − × − (4.117)
}2 }2 32π 2 ε20 }2 n2
s
µ2 e 4 rµe2 r 1 r
⇒ ρ=r 2 2 4 2
= 2
= × 2 = (4.118)
16π ε0 } n 4πε0 } n n  4πε0 }  a0 n
µe2
| {z }
Bohr’s radius Eq.(1.10)
The normalization constant can also be simplified further by using the expression for Bohr’s radius a0 as
s
 2 3 (n − (l + 1))!
Nnl = 3 (4.119)
na0 2n (n + l)!
χ(ρ)
We get the radial wave function from the solutions χ(ρ) where, Rnl (r) = 2l+1
= Nnl (2ρ)−l e−ρ Ln−(l+1) (2ρ),
r
then by substituting all the values back we get,
 r 
 2 3 (n − (l + 1))! 2r −l −
s !
 2r 
Rnl (r) = 3 e a0 n L2l+1n−(l+1) a n (4.120)
na0 2n (n + l)! a0 n 0

A few radial wave functions of Hydrogen atom are given below,


!
r
− 23 −
n=1 l=0 R10 (r) = 2a e a0

!
− 32 r  − 2ar

n=2 l=0 R20 (r) = (2a0 ) 2− e 0
a0
!
1 3 r − r
n=2 l=1 R21 (r) = √ (2a0 )− 2 e 2a0
3 a0
!
2 3
 18r 2r2  − 3ar
n=3 l=0 R30 (r) = (3a0 )− 2 27 − + 2 e 0
27 a0 a0
√ !
4 2 − 23 6r
 r2  − 3ar
n=3 l=1 R31 (r) = (3a0 ) − e 0
54 a0 a20
!
4 3 r
2 r

n=3 l=2 R32 (r) = √ (3a0 )− 2 2 e 3a0
27 10 a0

42
Complete Wave function
While finding the solutions for the wave function of hydrogen atom we came across few integers which
determine the form of the solution they are n, l, m which are called the principal quantum number, azimuthal
quantum number, and magnetic quantum number respectively. Thus the wave function is of the form,
ψnlm = Rnl (r) Θlm (θ) Φm (ϕ) = Rnl (r) Ylm (θ, ϕ) (4.121)
s ! !−l  r 
 2 3  2l + 1  (n − (l + 1))!(l − |m|)! 2r − + imϕ 2l+1  2r  m
ψnlm =
na0 8nπ
3
a0 n
e a0 n Ln−(l+1) a0 n
P l (cos(θ))
(n + l)! (l + |m|)!
The wave functions here are imaginary for m 6= 0, but we get real form by linear combination.
1 r
1s orbital → ψ100 = √ e− a
πa3
1  r  − 2ar
2s orbital → ψ200 = p 2 − e 0
4 2πa30 a0
1 r
− 2ar
2pz orbital → ψ210 = p e 0 cos(θ)
4 2πa30 a0
1  r  − 2ar
2px orbital → ψ211 = p 3 e 0 sin(θ) cos(ϕ)
8 πa0 a0
1  r  − 2ar
2py orbital → ψ21−1 = p 3 e 0 sin(θ) sin(ϕ)
8 πa0 a0
1  18r 2r2  − 3ar
3s orbital → ψ300 = p 27 − + 2 e 0
81 3πa30 a0 a0

2  6r r2  − 3ar
3pz orbital → ψ310 = p 3 − e 0 cos(θ)
81 πa0 a0 a20
1  6r r2 
− r
3px orbital → ψ311 = p 3 − 2 e 3a0 sin(θ) cos(ϕ)
81 πa0 a0 a0
1  6r r2 
− r
3py orbital → ψ31−1 = p 3 − 2 e 3a0 sin(θ) sin(ϕ)
81 πa0 a0 a0

43
Probability density function
The probability density of a particle is given by the square of its wave function as |ψ|2 = ψ ∗ ψ. For Hydrogen
atom ψnlm = Rnl (r) Θlm (θ) Φm (ϕ) and ψnlm ∗ ∗ (r) Θ∗ (θ) Φ∗ (ϕ). But R (r) Θ (θ) are real functions
= Rnl lm m nl lm
− 21 imϕ 1
so their complex conjugates are same. But Φm (ϕ) = (2π) e and Φm (ϕ) = (2π)− 2 e−imϕ . Thus the

probability density function is,

|ψnlm |2 = |Rnl (r)|2 |Θlm (θ)|2 (2π)eimϕ e−imϕ = (2π)|Rnl (r)|2 |Θlm (θ)|2 (4.122)

Thus the probability density function is real and doesn’t depend on the ϕ coordinate. Thus the probability
of finding the electron is independent of angle ϕ. We know ϕ is measured in xy−plane from the z−axis.
That implies that the electron density will be symmetrical about the z−axis. When the probability density
plots are graphed they will be symmetrical about z− axis and looks like the one below which have the 3D
graphs of first few energy states of Hydrogen.

Figure 4.1: Probability density functions for first few wave functions. (Griffith’s, Introduction to QM)

44
Chapter 5

Properties of particles in 3D
5.1 Angular Momentum
While solving the radial wave function we came across the effective potential energy given by,

}2 l(l + 1)
Veff = Velectrostatic + Vcentrifugal where Vcentrifugal = (5.1)
2µ r2
1
The angular momentum as we know is given by, L = µωr2 . The centrifugal energy is Vcentrifugal = µω 2 r2 .
2
In radial P.D.E. the centrifugal force was given as shown in the above equation. Thus equating both,

1 2 2 }2 l(l + 1)
µω r = ⇒ µ2 ω 2 r4 = }2 l(l + 1) ⇒ L2 = }2 l(l + 1) (5.2)
2 2µ r2
Angular momentum is a vector given by, L
~ = Lx î+Ly ĵ +Lz k̂ whose magnitude square is L2 = L2 +L2 +L2 .
x y z
The relation between angular and linear momentum is L ~ = ~r × p~. Thus the components are,

Lx = ypz − zpy Ly = zpx − xpz Lz = xpy − ypx (5.3)

} ∂ } ∂ } ∂ } ∂ } ∂ } ∂
⇒ L̂x = y −z L̂y = z −x L̂z = x −y (5.4)
i ∂z i ∂y i ∂x i ∂z i ∂y i ∂x
From the above relation can get the commutator brackets of the components of Angular momentum to be,

[L̂x , L̂y ] = i}L̂z [L̂y , L̂z ] = i}L̂x [L̂z , L̂x ] = i}L̂y (5.5)
So we see that the components of Angular momentum don’t commute between themselves. So we can say
that no two components of Angular momentum can not be found simultaneously.
h i
Consider the commutator [L̂2 , L̂x ] = L̂2x + L̂2y + L̂2z , L̂x = [L̂2x , L̂x ] + [L̂2y , L̂x ] + [L̂2z , L̂x ] (5.6)


Clearly, [L̂2x , L̂x ] = 0. And using the identity [ÂB̂, Ĉ] = Â[B̂, Ĉ] − [Â, Ĉ]B̂ we get,

[L̂2y , L̂x ] = L̂y [L̂y , L̂x ] − [L̂y , L̂x ]L̂y = L̂y (i})L̂z − (i})L̂z L̂y = i} L̂y , L̂z = −}2 L̂x (5.7)
 

[L̂2z , L̂x ] = L̂z [L̂z , L̂x ] − [L̂z , L̂x ]L̂z = L̂z (i})L̂y − (i})L̂y L̂z = i} L̂z , L̂y = +}2 L̂x (5.8)
 

⇒ [L̂2 , L̂x ] = 0 − }2 L̂x + }2 L̂x = 0 (5.9)


In the same manner we can show [L̂2 , L̂y ] = 0 and [L̂2 , L̂z ] = 0 . So, we can measure the total magnitude
of the angular momentum with only one of the components where the other two will be unknown. Thus
let us arbitrarily choose to measure L̂2 and L̂z quantities simultaneously. Now since these two operators
commute, they can have the same Eigen functions.

L̂2 ψ = λ1 ψ and L̂z ψ = λ2 ψ (5.10)

45
Now we shall define two operators called the Ladder operators which can raise or lower the Eigen values.

L+ = L̂x + iL̂y and L− = L̂x − iL̂y (5.11)


Since L̂x and L̂y commute with L̂2 we can tell that even the ladder operators L̂± will commute with L̂2 .
Thus even the ladder operators have the same Eigen function ψ. The commutator of L̂z and L̂± is,

[L̂z , L̂± ] = [L̂z , L̂x ± iL̂y ] = [L̂z , L̂x ] ± i[L̂z , L̂y ] = i}L̂y ± i(−i}L̂x ) = ±}(L̂x ± iL̂y ) = ±}L̂± (5.12)

Now L̂z (L̂± ψ) = L̂z L̂± ψ − L̂± L̂z ψ + L̂± L̂z ψ = (L̂z L̂± − L̂± L̂z )ψ + L̂± L̂z ψ (5.13)
= [L̂z , L̂± ]ψ + L̂± L̂z ψ = ±}L̂± ψ + L̂± (λ2 ψ) = (λ2 ± })(L̂± ψ) (5.14)

Thus we see that L̂± operator will add or subtract a factor of } to the Eigen value when operated on
an operator. Thus when the Eigen values are extremized, i.e., the ladder operator has to stop adding or
removing the } factor after an l number of energy states we get a condition that λ1 = }2 l(l + 1). Based on
the l values we can get the λ2 to be m}.

⇒ L̂2 ψlm = }2 l(l + 1)ψlm and L̂z ψlm = m}ψlm (5.15)

Thus the Eigen values of L2 is nothing√ but the√ value√we got earlier as } l(l+1), i.e., the magnitude of angular
2

momentum can take values of only } 2, } 6, } 12, . . . . And moreover we see that the z-component of
angular momentum can be only integer values of } i.e., Lz = }, 2}, 3}, . . . . And for a given value of l
there are 2l + 1 values of m which are 0, ±1, ±2 . . . , ±l. We have to note that these results are for a general
case. We haven’t specifically mentioned it is for the Hydrogen atom alone. Except for the conditions for
the values of l which need it to be an integer. In fact any vector J~ that satisfies the Eq.(5.5) has the result
of J 2 = }j(j + 1) and Jz = m}. There is also no condition that j has to be an integer, it can also be
fractions as the result was got only by algebra and Ladder operators. For a given j, m can take values of
−j, (−j + 1), (−j + 2), . . . , 0, . . . , (j − 2), (j − 1), j a total of 2j+1 values.

Angular momentum in spherical coordinates


The Angular momentum vector is given by L~ = ~r × P~ . and P~ = −i}∇~ . Thus in polar coordinates ~r = rr̂

and ~ = r̂ ∂ + θ̂ 1 ∂ + ϕ̂ 1


⇒ L̂ =
}h ∂ ∂
r(r̂ × r̂) + (r̂ × θ̂) + (r̂ × ϕ̂)
1 ∂ i
(5.16)
∂r r ∂θ r sin(θ) ∂ϕ i ∂r ∂θ sin(θ) ∂ϕ

}h ∂ θ̂ ∂ i
But r̂ × r̂ = 0 r̂ × θ̂ = ϕ̂ r̂ × ϕ̂ = −θ̂ . Thus we get, L̂ = ϕ̂ − (5.17)
i ∂θ sin(θ) ∂ϕ
Now the unit vectors φ̂ and θ̂ in Cartesian coordinates are given by,

θ̂ = cos(θ) cos(ϕ)î + cos(θ) sin(ϕ)ĵ − sin(θ)k̂ ϕ̂ = − sin(θ)î + cos(ϕ)ĵ (5.18)

Thus we get the x, y, z components of angular momentum in terms of polar coordinates to be,
" # " # " #
} ∂ ∂ } ∂ ∂ } ∂
L̂x = − sin(ϕ) − cos(ϕ) cot(θ) L̂y = − cos(ϕ) − sin(ϕ) cot(θ) L̂z =
i ∂θ ∂ϕ i ∂θ ∂ϕ i ∂ϕ

We previously saw that the only measurable property of angular momentum is its Magnitude and its z-
component which is got by differentiating the Eigen function with respect to ϕ alone. So for Hydrogen
1
atom ψnlm = Rlm Θml Φm . where Φm = (2π) e
2 imϕ Thus we get,

} ∂ψlmn } ∂  Rlm Θml } ∂Φm Rlm Θml } 1


L̂z ψlmn = = Rlm Θm
l Φ m = = (im)(2π)− 2 eimϕ = m}ψlmn (5.19)
i ∂ϕ i ∂ϕ i ∂ϕ i

⇒ L̂z ψlmn = m}ψlmn (5.20)

46
Also when a magnetic field is applied to the atom, the electrons behave as small magnetic dipole of magnetic
moment M ~ = − eL~ where µ is the reduced mass and L ~ is the angular momentum. So when the measurement

of M
~ is done it is actually the Z component that can be measured that is given by Mz = − em} . So even

Mz is quantized and can take only integral values of − 2µe}
which is called Bohr Magneton. The states
corresponding to different values of the quantum number m can be observed by an external magnetic field
where the direction of the field is taken to be the Z−axis as only that can be measured. So that is the
reason why m is called the magnetic quantum number. So the direction of Angular momentum is not
known except the z component and its magnitude. Now the z− component and the magnitude are related
by Lz = L cos(θ) where the angular momentum vector makes a constant angle of θ with the z−axis but
the angle that it makes with the x or y−axis is not known. Thus it makes a cone whose vertex is at the
origin and angle of θ.

5.2 Spin
We learned the Angular momentum of the particle where the components depended on the spatial position
of the particle. This is extrinsic angular momentum which is analogous to the orbital angular momentum
of earth. But there is also another type of angular momentum associated with the particle which doesn’t
depend on its external movement. For example, if there is a system where an electron is having linear motion
only (angular momentum is zero), still there was a non zero angular momentum when measured, which
implies that the particle should have an intrinsic spinning nature which is analogous to the spin angular
momentum of a classical particle. The particle’s spin has nothing to do with the classical rotational motion of
particles. But it is present intrinsically and behaves as spin. We have already formulated the mathematical
theory for Angular momentum in the previous section. The exact same math is enough to work on spin
angular momentum. Thus the spin angular momentum as a vector denoted by S ~ = Sx î + Sy ŝ + Sz k̂ whose
magnitude square is S = Sx + Sy + Sz which follow the same commutation rules given by,
2 2 2 2

[Ŝx , Ŝy ] = i}Ŝz [Ŝy , Ŝz ] = i}Ŝx [Ŝz , Ŝx ] = i}Ŝy (5.21)

[Ŝ2 , Ŝx ] = 0 [Ŝ2 , Ŝy ] = 0 [Ŝ2 , Ŝz ] = 0 (5.22)


Only the magnitude of the spin and one of its components (say Sz ) can be measured so,

Ŝ2 ψ = }2 s(s + 1)ψ and Ŝz ψ = ms }ψ (5.23)

The Ladder operators Ŝ± = Ŝx ± iŜy are defined by, = } s(s + 1) − m(m ± 1) ψsms ±1 (5.24)
p
Ŝ± ψsms
Here s is the angular spin quantum number and ms is the magnetic spin quantum number. Since this result
has not come from the spherical harmonics we have the freedom to choose s to be even half-integer values
and ms can take values depending on s as, −s, (−s + 1), (−s + 2), . . . , 0, . . . , (s − 2), (s − 1), s . These results
were actually invoked by Dirac while working on his relativistic quantum mechanics. But we shall use his
results to continue our understanding regarding particles.

47
5.2.1 Spin 1/2 Particles
Spin Matrices
They are the particles whose spin quantum number is given to be s = 1/2. These particles are of special
interest as most of the particles we encounter are particles whose spin is half-integer. There are two Eigen
states for spin state for such particles which are Spin-Up denoted by | ↑ i = |1/2i and for Spin-Down it is
denoted by | ↓ i = | − 1/2i. So using these as basis vectors we can get the general state for the half-spin
particle which can be expressed by a two-column matrix as,
     
a 1 0
χ= = aχ+ + bχ− where χ+ = and χ− = (5.25)
b 0 1
| {z } | {z }
spin-up spin down

So χ+ and χ− are state vectors where the spin operators can operate on. Thus we get,

Ŝ2 χ+ = }2 s(s + 1)χ+ and Ŝ2 χ− = }2 s(s + 1)χ− (5.26)


3 3
Now as s = 1/2 we get, Ŝ2 χx = }2 χ+ and Ŝ2 χ− = }2 χ− (5.27)
4 4
       
1 3 1 0 3 0
⇒ Ŝ2 = }2 and Ŝ2 = }2 (5.28)
0 4 0 1 4 1
So we can think Ŝ2 to be a matrix which has 2 rows and 2 columns. Let the matrix have some elements
a11 , a12 , a21 , a22 then we get,
   3}2 
3}2 1
    
a11 a12 1 a11
= ⇒ = 4 (5.29)
a21 a22 0 4 0 a21 0

So we get a11 = 3}2 /4 and a21 = 0. Then by the second equation we get,

3}2 0
        
a11 a12 0 a12 0
= ⇒ = 3}2 (5.30)
a21 a22 1 4 1 a22 4

3}2
So we get a12 = 0 and a22 = 4 . Thus Ŝ2 operator can be written in matrix format as,

3}2
 
1 0
2
Ŝ = (5.31)
4 0 1

Similarly we can also find the Ŝz operator. Since s = 1


2 we get ms = + 12 and − 1
2 which corresponds to
spin-up and spin-down. Thus,
} }
Ŝz χ+ = χ+ and Ŝz χ− = − χ− (5.32)
2 2
Let Ŝz be a matrix whose elements are z11 , z12 , z21 , z22 . Using the matrix form of χ+ and χ− we get,
       }
z11 z12 1 } 1 z11
= ⇒ = 2 (5.33)
z21 z22 0 2 0 z21 0
So we get z11 = }
2 and z21 = 0. Then by the second equation we get,
        
z11 z12 0 } 0 z12 0
=− ⇒ = (5.34)
z21 z22 1 2 1 z22 − }2
So we get z12 = 0 and z22 = − }2 . Thus Ŝz operator can be written in matrix format as,
 
} 1 0
Ŝz = (5.35)
2 0 −1

48
To get the Ŝx and Ŝy matrices we shall use the definition of the ladder operators defined by,
Ŝ± ψsms = } s(s + 1) − m(m ± 1) ψsms ±1 (5.36)
p

By this definition we get, Ŝ+ χ− = }χ+ Ŝ+ χ+ = }χ− Ŝ+ χ+ = Ŝ− χ− = 0 (5.37)
To get the Ladder spin matrices we shall expand the equation in matrix notations. And when solved for
Ŝ+ and Ŝ− in the same method we did earlier we get,
   
0 1 0 0
Ŝ+ = } and Ŝ− = } (5.38)
0 0 1 0

Now to get Ŝx and Ŝy we shall just use the basic definition of ladder operators where Ŝ+ = Ŝx + iŜy and
Ŝ− = Ŝx − iŜy we get,
Ŝ+ + Ŝ− Ŝ+ − Ŝ−
Ŝx = and Ŝy = (5.39)
2 2i
Thus we get the matrix form for Ŝx and Ŝy as,
   
} 0 1 } 0 −i
Ŝx = and Ŝy = (5.40)
2 1 0 2 i 0

As all these operators Ŝx , Ŝy , Ŝz have a factor of }/2 we can combine them and write,
}
Ŝ = σ̂ (5.41)
2
where ~σ = σx î + σy ŝ + σz k̂ which are vector operators called as Pauli Spin matrices given by,
     
0 1 0 −i 1 0
σ̂x = σ̂y = σ̂z = (5.42)
1 0 i 0 0 −1

We can see that the Spin matrices Ŝ2 , Ŝx , Ŝy , Ŝz are all Hermitian matrices but Ŝ+ , Ŝ− are not
Hermitian matrices. Thus only the observable of Ŝ2 , Ŝx , Ŝy , Ŝz can be measured as the eigen values
and expectation values are real. But for Ŝ+ , Ŝ− as they are not hermitian they don’t have real observable.
Now if we try to measure the spin of the particle in some general state χ where χ = aχ+ + bχ− then the
probability of measuring the spin to be + }2 i.e., with the state χ+ is |a|2 and the probability of measuring
the spin to be − }2 i.e., with the state χ− is |b|2 . Since these are the only two possible states that can be
observed, the total probability |a|2 + |b|2 = 1. So the spinor χ has to be normalized.

Consider that you want to measure the x− component of spin using the operator Ŝx , then what are the
Eigen values of Ŝx , for that,
}2

0 − λ 1 −λ 1
}
Ŝx − Iλ = = = 0 ⇒ λ2
− =0 ⇒ λ=± (5.43)

1 0−λ 1 −λ 4 2
Thus the Eigen values of spin in x−axis is also ± }2 corresponding to spin-up or down. Here if the spinor
 
(x) (x) α
is χ = αχ+ + βχ− then we get χ =
(x) (x) . Now consider the eigen value equation Ŝx χ(x) = ± }2 χ(x) ,
β
expanding it,
        
} 0 1 α } α } β } α
=± ⇒ = ⇒ β = ±α (5.44)
2 1 0 β 2 β 2 α 2 β
Thus the normalization condition given by 2 2
 |α| + |β| = 1 becomes
 2|α| = 2|β| = 1 which implies that
2 2

(x) 1 1 (x) 1 1
the Eigen spinors of Ŝx are χ+ = √ and χ− = √ a general x spinor is given by,
2 1 2 −1
! ! ! !
a + b (x) a − b (x) a + b a − b
χ(x) = √ χ+ + √ χ− α= √ β= √ (5.45)
2 2 2 2

49
We have to note that the spin is a new degree of freedom that a particle can occupy. Since s = 1/2 we
get ms = ±1/2 which implies that particle along with having the quantum numbers n, l, m can have either
spin-up state and spin down state. Thus there are two types of wave functions for a particle given by ψ+ for
a particle with spin-up state and ψ− for a particle with spin down state. Thus the wave function becomes,
   
1 ψnlml (r, θ, ϕ)
⇒ ψnlml 1/2 (r, θ, ϕ) = ψnlml (r, θ, ϕ)χ+ = ψnlml (r, θ, ϕ) = = ψ+ (5.46)
0 0
   
0 0
similarly, ψnlml −1/2 (r, θ, ϕ) = ψnlml (r, θ, ϕ)χ− = ψnlml (r, θ, ϕ) = = ψ− (5.47)
1 ψnlml (r, θ, ϕ)
Here ml is the orbital magnetic quantum number associated with the extrinsic angular momentum and ms
is the spin magnetic quantum number associated with the intrinsic spin of the particle.

Stern-Gerlach Experiment

This experiment measures the spin angular momentum of the particle quantitatively. Consider a magnetic
field where B~ is in the direction of the z−axis. Then if a stream of particles whose spin quantum number
s = 1/2 is made to flow in the x−axis then there will be half the number of particles that gets deflected
upwards as ms = 1/2 and the other half the number of particles get deflected downward as ms = −1/2. This
deflection is because of the z− component of the spin angular momentum of particles when interacted with
the magnetic field. The Eigen values of Sz are given by
1 1
Sz = ms } where ms = for spin-up, and ms = − for spin down. (5.48)
2 2
But the spin magnetic moment vector is given by M ~ s = −gs e S ~ where gs is the spin factor. So the
2me
z−component of observed magnetic moment is
e }e
Msz = −gs · ms } = −gs · ms = −gs µB ms (5.49)
2me 2me
}e 1
where µB = = 9.274 × 10−24 A m2 is the Bohr Magneton. Now since ms = ± the observed z−
2me 2
µB
component of the magnetic moment has to be Msz = ±gs . Now the spin Gyro-magnetic ratio (γspin ) is
2
given by the ratio of z−component of magnetic moment to the z−component of the spin angular moment.
1

Ms z
}e
gs 2m ± 2 e
γspin = = e
1
 = gs (5.50)
Sz } ±2 2me
Similar even for orbital angular momentum we can define a gyro-magnetic ratio given by,
}e }e
Mz e
γorbital = = 2me = 2me = (5.51)
Lz ml } } 2me
Now if we take the ratio of γspin and γorbital we get,
γspin
= gs (5.52)
γorbital
When experimentally calculated the ratio between the spin gyro-magnetic ratio and orbital gyro-magnetic
ratio was found out to be 2[1.0011596521884(±43)]. Also theoretically when the value of spin factor derived
by Dirac in his relativistic quantum mechanical theory it came out to be 2[1.001159652140(±28)].

50
5.2.2 Half Integral spin (n/2) and Integral spin (n) Particles
Broadly all the particles can be classified into two types.
1. Particles with Integer-spins are called Bosons whose behavior is governed by a particular class of
quantum statistics called as Bose-Einstein statistics.

2. Particles with Half-integer spins are called as Fermions whose behavior is governed by a class of
quantum statistics called as Fermi-Dirac Statistics.
Consider a two-particle system whose individual wave function is Ψ (~ra , t) and Ψ (~rb , t). The total wave
function of the system is given by Ψ (~ra , ~rb , t). So one single wave function describes the two particles. Then
can we actually distinguish between the states of two particles? i.e., could we tell that specific particle is
in state Ψa and other is in Ψb ?? No, because once the particles are together they lose their individuality.
We cannot name an electron and keep track of it because they are not observable unless we change the
particle’s state. Now the conditions this total wave function has to follow are the same as that we learned
in the third chapter. Now the Hamiltonian of the system is given by,

}2 2 }2 2
Ĥ = − ∇a − ∇ +V (5.53)
2ma 2mb b
The subscripts indicate that the differentiation is done with respect to their respective coordinates. Now
the wave function can be separable if the potential energy is time-independent. Then we get,
E
Ψ (~ra , ~rb , t) = ψ(~ra , ~rb )ei } t (5.54)

The time factor is the same as E is the total energy of the system and not of the individual particles. Now
suppose one particle is in state ψ(~ra ) and one particle is in state ψ(~rb ) but we don’t know which is which.
Then the total wave function is given by the linear combination,

ψ(~ra , ~rb ) = c1 ψ(~ra )ψ(~rb ) ± c2 ψ(~ra )ψ(~rb ) (5.55)

When the normalization condition is applied we get c1 = c2 = √1 ,


2
thus,

1 h i
ψ(~ra , ~rb ) = √ ψ(~ra )ψ(~rb ) ± ψ(~ra )ψ(~rb ) (5.56)
2
Now by quantum statistics, it is seen that for particles with integer spins, their wave functions are anti-
symmetric and for particles with integer spins, their wave function is symmetric, i.e.,

ψ(~ra , ~rb ) = +ψ(~rb , ~ra ) ⇒ Boson (5.57)

ψ(~ra , ~rb ) = −ψ(~rb , ~ra ) ⇒ Fermion (5.58)


1 h i
Boson ⇒ ψ(~ra , ~rb ) = √ ψ(~ra )ψ(~rb ) + ψ(~ra )ψ(~rb ) (5.59)
2
1 h i
Fermion ⇒ ψ(~ra , ~rb ) = √ ψ(~ra )ψ(~rb ) − ψ(~ra )ψ(~rb ) (5.60)
2
If the quantum numbers are the same for both the particles, i.e., if ψ(~ra ) = ψ(~rb ) = ψ(~r), then we get,

Boson ψ(~ra , ~rb ) = 2ψ 2 (~r) |ψ(~ra , ~rb )| = 2 ψ 2 (~r) (5.61)

⇒ ⇒

Fermion ⇒ ψ(~ra , ~rb ) = 0 ⇒ |ψ(~ra , ~rb )| = 0 (5.62)


Thus for Bosons, if two particles with the same quantum state are in the same system they become one
single entity whose probability density function doubles, i.e., bosons are more likely to occupy the same
quantum state. And for Fermion the probability density becomes zero, i.e., the probability of fermions
being in the same energy state is Zero. So no two fermions in the same system can occupy the same energy
state. This is known as the ‘Pauli’s exclusion principle’.

51
Bibliography

[1] David J Griffiths and Darrell F Schroeter. Introduction to quantum mechanics. Cambridge University
Press, 2018.
[2] Ramamurti Shankar. Principles of quantum mechanics. Springer Science & Business Media, 2012.
[3] Nouredine Zettili. Quantum mechanics: concepts and applications, second edition. 2003.
[4] R Gilmore. “The Interpretation of Quantum Mechanics by Roland Omnes”. In: MATHEMATICAL
INTELLIGENCER 18.1 (1996), pp. 70–74.
[5] CL Arora and PS Hemne. Physics for Degree Students B. Sc Third Year. S. Chand Publishing.
[6] Puri Sharma, RL Sharma, and Pathania. Principles of Physical Chemistry, Chapter 1 and 2. 2004.

52

Potrebbero piacerti anche