Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

AMC ENGINEERING COLLEGE

DEPARTMENT OF PHYSICS

Module 1 Notes

Quantum Mechanics

I/II SEMESTER Physics for EEE Stream


Subject code: BPHYE102/202
This page was intentionally left blank.
Contents

1 Quantum Mechanics 5
1.1 Wave-Particle Duality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.1 de Broglie’s Hypothesis . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.2 Worked Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 Wave Packets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.1 Phase Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.2 Group Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3 Heisenberg’s Uncertainty Principle . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3.1 Principle of Complementarity . . . . . . . . . . . . . . . . . . . . . . . 13
1.3.2 Non-existence of electron in the nucleus of an atom (non-relativistic case) 13
1.3.3 Worked Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.4 Wavefunction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.4.1 Properties of a wavefunction . . . . . . . . . . . . . . . . . . . . . . . 16
1.5 Schrodinger’s Time-Independent Wave Equation . . . . . . . . . . . . . . . . 16
1.5.1 Particle in an infintely deep potential well (Particle in a 1D box) . . . 20
1.5.2 Worked Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.6 Model and Previous Year Questions . . . . . . . . . . . . . . . . . . . . . . . 26
1.7 Numericals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
1.7.1 de Broglie Wavelength . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
1.7.2 Heisenberg’s Uncertainty Principle . . . . . . . . . . . . . . . . . . . . 26
1.7.3 Particle in a 1D box . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

3
This page was intentionally left blank.
Chapter 1

Quantum Mechanics

Syllabus
Quantum Mechanics
de Broglie Hypothesis and Matter Waves, de Broglie wavelength and derivation of expres-
sion by analogy, Phase Velocity and Group Velocity, Heisenberg’s Uncertainty Principle
and its application (Non existence of electron inside the nucleus-Non Relativistic), Princi-
ple of Complementarity, Wave Function, Time independent Schrödinger wave equation,
Physical Significance of a wave function and Born Interpretation, Expectation value,
Eigen functions and Eigen Values, Particle inside one dimensional infinite potential well,
Waveforms and Probabilities. Numerical Problems.

1.1 Wave-Particle Duality


Light (radiation) appears to behave as a wave in some scenarios and as a particle in other
scenarios.
A particle has mass; it occupies space; it can move from one place to another; it gives
energy when slowed down or stopped. Thus the particle can be specified by mass, velocity,
momentum, energy. And any two particles cannot occupy the same space at the same time.
But a wave spreads, as it travels, occupying a larger region of space. A wave is specified by
wavelength, frequency, amplitude, intensity, phase, and wave velocity.
In the case of interference, diffraction, and polarization, light behaves as a wave since the
presence of two waves are required at the same time and at the same space to explain these
phenomena. But we know that two particles cannot occupy the same position at the same
time. Thus light behaves as a wave in these situations.
In the case of photoelectric effect, Compton Effect, emission and absorption of radiation
(blackbody radiation, for example) etc., the radiation behaves as a particle (photon).
Thus we can conclude that radiation has dual nature (wave-like and particle-like). But
radiation cannot exhibit both the properties of a particle and a wave simultaneously.

1.1.1 de Broglie’s Hypothesis


We concluded that light (radiation) exhibits dual nature in the previous section. Due to this
result, in 1924, Louis de Broglie put forth a suggestion that matter also exhibits dual nature,
just like light, i.e., matter which is made up of discrete particles (for example, atoms, protons,
electrons etc.) might exhibit wave like properties under appropriate conditions. The waves
associated with moving material particles were called matter waves.

5
1.1. WAVE-PARTICLE DUALITY

Properties of matter waves


1. A matter wave represents the probability of finding a particle in space.
2. They can propagate in vacuum and hence they are not mechanical waves.
3. They are not electromagnetic in nature.
4. They are independent of the charge on the material particle.
5. The phase velocity of matter waves can be greater than the speed of light. Hence their
speeds do not have any physical significance.
Here we derive an expression for wavelength of such a wave by analogy. Consider a
particle of mass m moving with velocity v. From Einstein’s theory of relativity, we have a
relation between the particle’s mass and energy E given by
E = mc2 (1.1)
Now, we need an expression for energy in terms of a wave property like wavelength. We
have from Planck’s law, E = hν. Writing this in terms of wavelength (using c = νλ), we get
hc
E= (1.2)
λ
Let’s equate (1.1) and (1.2), we get
hc
mc2 =
λ
h
λ=
mc
Since the particle is moving with velocity v, let’s replace c (the velocity of light) in the
above expression with v. So we have
h
λ=
mv
But we know that p = mv is the momentum of the particle. So finally, we have
h
λ= (1.3)
p
The wavelength λ of a matter wave associated with a particle moving with momentum p
is called the de Broglie wavelength.
Note: Look at the Planck’s formula for energy and de Broglie’s formula for wavelength
h
E = hν and λ =
p
Both these formulae are indicating a relationship between particle characteristics (E and p)
and wave characteristics (ν and λ), and they are linked through the same constant h.
The above formula for de Broglie wavelength was in terms of momentum or velocity of the
particle. We can also write the formula in terms of the particle’s kinetic energy Ek = 12 mv 2 .
We do it as follows:
1 1 2 2
Ek = mv 2 = m v
2 2m
p2
Ek =
2m
p
=⇒ p = 2mEk

Department of Physics 6 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.1. WAVE-PARTICLE DUALITY

Substitute this p in de Broglie’s formula


h
λ= √ (1.4)
2mEk
This is the formula for de Broglie wavelength in terms of the particle’s kinetic energy.
Now, consider a charged particle of charge q, mass m accelerated through a potential
difference (voltage) V . The work done on this particle due to the potential difference is
given by qV . This work done is converted into the particle’s kinetic energy, i.e., Ek = qV .
Substitute this Ek in (1.4).
h
λ= √
2mqV
(Note: q is only magnitude of charge here.)
As a special case, consider the charged particle to be an electron. Then q = e, where
e = 1.6 × 10−19 C is the charge of the electron. Then the above formula becomes
 
h h 1
λ= √ = √ √
2meV 2me V
Here, h, m, e are all constants. After substituting in these constants and simplifying, we get
12.27
λ= √ Å
V

where V is in volts and 1 Å = 10−10 m.

1.1.2 Worked Examples

Q 1. Calculate the momentum and de Broglie wavelength associated with an electron with
a kinetic energy of 1.5 keV.
Given:
Ek = 1.5 keV = 1.5 × 1 × 103 × 1.6 × 10−19 = 2.4 × 10−16 J
me = 9.11 × 10−31 kg (because electron)
To find: p =?, λ =?
We know that
p
p = 2me Ek = 2 × 9.11 × 10−31 × 2.4 × 10−16 = 2.0911 × 10−23 Ns
p

p = 2.0911 × 10−23 Ns

Next, we know that

h 6.625 × 10−34
λ= = = 3.1682 × 10−11 m
p 2.0911 × 10−23

λ = 3.1682 × 10−11 m

Q 2. Find the energy of neutron in eV whose de Brogle wavelength is 1 Å.


Given:
λ = 1 Å = 1 × 10−10 m

Department of Physics 7 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.2. WAVE PACKETS

m = 1.674 × 10−27 kg (because neutron)


To find: Ek =?
h
We know that λ = √
2mEk
After rearranging terms, we have
2
h2 6.625 × 10−34
Ek (in J) = =
2mn λ2 2 × 1.674 × 10−27 × (1 × 10−10 )2
Ek (in J) = 1.311 × 10−20 J

Converting this to eV,

Ek (in J) 1.311 × 10−20


Ek (in eV) = = = 0.0819 eV
e 1.6 × 10−19

Ek (in eV) = 0.0819 eV

Q 3. Find the de Broglie wavelength of a particle of mass 0.58 MeV/c2 that has a kinetic
energy of 90 eV, where c is the speed of light.
Given:
0.58 MeV 0.58 × 1 × 106 × 1.6 × 10−19
m= = 2 = 1.0311 × 10−30 kg
c2 8
(3 × 10 )

Ek (in J) = 90 eV = 90 × 1.6 × 10−19 = 1.44 × 10−17 J


To find: λ =?
We know that
h 6.625 × 10−34
λ= √ =√ = 1.2157 × 10−10 m
2mEk 2 × 1.0311 × 10−30 × 1.44 × 10−17

λ = 1.2157 × 10−10 m

1.2 Wave Packets


Let us now look at how to represent the de Broglie wave associated with any particle. We will
start by introducing a mathematical structure to the wave and later give a physical meaning
to it.

1.2.1 Phase Velocity


For a travelling undamped sinusoidal wave, the vertical displacement y of any point on the
wave can be written in terms of the point’s position x and time t as

y = A sin(ωt − kx) (1.5)

where A is the amplitude, ω is the angular frequency, and k is the wave number. ω and k
contain information about the wave’s frequency ν and wavelength λ through the expressions
ω = 2πν and k = 2πλ .

Department of Physics 8 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.2. WAVE PACKETS

Figure 1.1: Travelling Wave

Look at figure 1.1 for example. It shows the graph of a travelling wave with y on the
y-axis and x on the x-axis at an instant of time t with some given value of k and ω. This
graph can be thought of as the path traced by a vertically oscillating object attached to a
spring which also moves with a certain velocity v to the right. The dot represents the vertical
distance y of the oscillating object at x = 2 at this instant of time.
In (1.5), the term (ωt − kx) gives the phase of the oscillating object. This term is a
constant with respect to time and thus we can write
d
(ωt − kx) = 0
dt
dx
ω−k =0
dt
dx ω
= (1.6)
dt k
If a point is imagined to be marked on a travelling wave, then it becomes a representative
point for a particular phase of the wave (for example, if the imagined point is the top most
location on the crest, then it represents the π/2 phase.), and the velocity with which it is
transported owing to the motion of the wave is called the phase velocity. Since in (1.6) x is
the distance travelled, dx
dt is taken as the phase velocity vphase . Therefore we have

ω
vphase = (1.7)
k
Now, you might be tempted to conclude that this phase velocity represents the velocity
of the particle vparticle whose wave nature we are trying to understand. This is not it as you
will see soon.

1.2.2 Group Velocity


When a group of two or more travelling waves, differing slightly in wavelengths, are super-
imposed on each other, a resultant pattern emerges in the shape of variation in amplitude
(see blue curve in figure 1.2). Such variation represents the wave group as a whole, and
is called the wave packet (represented by the red curve in the figure). If the velocities of
the component waves are equal, then the envelope enclosing the amplitude variation pattern
also, or the wave packet also, travels with the same velocity as the common velocity of the
waves. Since the group consists of waves of different wavelengths, they are bound to move

Department of Physics 9 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.2. WAVE PACKETS

Figure 1.2: Wave packet formed by superposition of two waves of differing wavelengths

with different velocities in a dispersive medium, at which time, the envelope or the wave
packet is carried with a velocity which will be different from the others (given the frequency
of the constituent waves are also different). The velocity with which the wave packet moves
is called the group velocity.
Group velocity is the velocity with which the envelope enclosing a wave group (called
wave packet, formed due to superposition of two or more travelling waves of slightly different
wavelengths) is transported. It is the velocity with which the energy transmission occurs in
a wave.
It may be noted that the phase velocity cannot be defined for a wave packet.
If we have two travelling waves with angular frequency ω and ω + ∆ω respectively and
wave number k and k + ∆k respectively, where the ∆ quantities are small, then it can be
shown that the velocity of the wave packet, called the group velocity, is
∆ω
vgroup =
∆k
In the limit, we have
∆ω dω
−→
∆k dk
Therefore we get the expression for group velocity as

vgroup = (1.8)
dk
This expression is applicable in general to a superposition of more than two travelling
waves also.
It turns out that the group velocity (and not the phase velocity) is the velocity of the
particle whose wave nature we are trying to understand. To see why, let’s find the relation
between vgroup and vparticle .

Relation Between Group Velocity and Particle Velocity


Consider the expression for group velocity (1.8),


vgroup =
dk
Here, ω(= 2πν) and k(= 2π
λ ) are properties of the wave. We need to introduce particle
properties, energy E and momentum p, here. But we know the relations connecting the two
pictures:
E
E = hν −→ ν =
h

Department of Physics 10 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.2. WAVE PACKETS

and
h
λ=
p
Therefore, we can write
E
ω = 2πν = 2π
h
and
1 p
k = 2π = 2π
λ h
Let us consider some small increments dω and dk in ω and k respectively. Then we can
write
dE
dω = 2π
h
and
dp
dk = 2π
h
Substitute these in (1.8). We have

dE

vgroup = h
dp

h
dE
vgroup = (1.9)
dp

So we have a relation between the group velocity and the properties of the particle. Now,
consider a particle of mass m moving with velocity vparticle in free space, i.e., no external
forces, which means the potential energy of the particle is zero. Therefore the total energy
E of the particle is completely due to its kinetic energy. So we can write

1 2 p2
E = mvparticle =
2 2m

where p = mvparticle is the momentum of the particle. Substitute this E in (1.9)

p2
 
d
vgroup =
dp 2m
p
vgroup =
m

But p = mvparticle . Thus we have finally

vgroup = vparticle (1.10)

Therefore, we can conclude that the de Broglie wave group (made up of many travelling
waves) associated with a free particle travels with a velocity equal to the velocity of the
particle itself.

Department of Physics 11 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.3. HEISENBERG’S UNCERTAINTY PRINCIPLE

Correlation b/w de Broglie Wavelength, Uncertainty principle, wavepacket


The main idea conveyed in the preceding section is that to accurately represent a particle as a
wave, a single traveling wave is insufficient. Instead, the wave nature of the particle can only
be accurately described by combining multiple traveling waves with different wavelengths, as
dictated by de Broglie’s formula (1.3). However, this implies considering multiple versions
of the same particle with different momenta simultaneously, leading to uncertainty regarding
the particle’s momentum.
Additionally, the wave packet representing the particle has a (group) velocity which is
equal to the particle’s velocity, but as material particles are point-like whereas the wave packet
occupies finite space, this representation introduces uncertainty in the particle’s position.
Attempting to reduce this uncertainty by making the wave packet smaller necessitates adding
more traveling waves to the superposition, increasing the uncertainty in momentum.
Conversely, aiming to decrease the uncertainty in the particle’s momentum leads to a
spread-out wave packet and increased uncertainty in position. Notably, the concept of ’zero
uncertainty’ is physically meaningless, as it would imply an infinite uncertainty in the other
quantity. For example, zero uncertainty in momentum would imply the particle could be
anywhere in space, resulting in infinite uncertainty in position, and vice versa.
Therefore, we say that there exists an inverse relationship between the uncertainties of a
particle’s position and momentum, as described by Heisenberg’s uncertainty principle. This
principle quantifies the fundamental limit on the precision with which certain pairs of physical
properties of a particle can be simultaneously known.

1.3 Heisenberg’s Uncertainty Principle


In classical mechanics, a particle occupies a definite place in space and possesses a definite
momentum. By knowing the position and momentum of a particle at any given instant of
time and knowing its interactions with its surroundings, it is possible to evaluate its position
and momentum at any later stage, and the trajectory of the particle could be continuously
traced. But in quantum mechanics, there is an inherent inability to accurately determine the
position and momentum of a particle simultaneously.
According to the Heisenberg’s Uncertainty Principle, it is impossible to determine simul-
taneously both the position and momentum of a particle with complete accuracy. This is
represented quantiatively by the following inequality:

∆x∆p ≥ (1.11)
2
where ∆x is the uncertainty in determining the position of the particle, ∆p is the uncertainty
in determining the momentum of the particle, and h̄ is called the reduced Planck’s constant
given by
h
h̄ =

(Note that here we are considering motion in only one dimension. If the motion was in
three dimensions, we would have three inequalities relating the three coordinates (x, y, z)
with the three components of momentum (px , py , pz ) respectively.)
Two consequences of the Heisenberg’s uncertainty principle are as follows:

ˆ An increase in the accuracy of determination of a particle’s position means a decrease


in the value of its uncertainty ∆x. As per (1.11), a decrement in the value of ∆x must
simultaneously result in the increment in the value of ∆p since the right hand side of

Department of Physics 12 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.3. HEISENBERG’S UNCERTAINTY PRINCIPLE

the equation is a constant. So in such a case, there is a decrease in the accuarcy of


simultaneous determination of momentum.

ˆ ∆x and ∆p can never be zero in the microscopic realm. If one of these is zero, then
the left hand side of (1.11) becomes zero and the inequality won’t hold. This tells
us that our inability to accurately determine these quantities is not because of poor
instruments or less sophisticated technology, but is due to the inherent nature of the
system under investigation. No matter how advanced an instrument one invents, it can
never determine these quantities with 100% accuracy. (Remember that this relation
applies to only microscopic (quantum) particles. Classically, for macroscopic particles,
h̄ is such a small quantity compared to the other quantities involved that it may be
assumed to be zero. So only in this case, the uncertainties can become zero.)

1.3.1 Principle of Complementarity


The complementarity principle holds that objects have certain pairs of complementary prop-
erties which cannot all be observed or measured accurately simultaneously. One of these
pairs is position and momentum. Other pairs include

ˆ Energy (E) and time (t) with ∆E∆t ≥
2

ˆ Angular displacement (θ) and angular momentum (L) with ∆θ∆L ≥
2
The reason these specific pairs satisfy this principle and not others is seen in the formalism
of quantum mechanics which is beyond the scope of this course. But the key takeaway
of this principle is the fact that setting up an experiment to measure one quantity of a
pair (say, position of an electron) excludes the possibility of accurately measuring the other
(momentum, in this case), yet understanding both experiments is necessary to characterize
the object under study. In other words, both the light and particle nature of any object
together can give us the complete picture of the object.

1.3.2 Application of Heisenberg’s Uncertainty Principle


Non-existence of electron in the nucleus of an atom (non-relativistic case)
In the β-decay process of radioactive radiation, it was observed that the emitted radiation
was composed of electrons (β rays), and the energy of these electrons was found to be in the
range of 3 to 4 M eV . But we know that electrons exist outside the nucleus of an atom and
not inside it. So does this experimental observation suggest that electrons can also be found
inside the nucleus? We’ll use Heisenberg’s Uncertainty Principle to check whether this is true
or not.
The typical radius of an atomic nucleus is nearly 5 × 10−15 m. If an electron is to exist
inside the nucleus, then the uncertainty in its position ∆x should not exceed this value, i.e.,
we can take ∆x = 5 × 10−15 m. Then from (1.11) we get the uncertainty in momentum ∆p
as
h
∆p ≥
4π∆x
6.625 × 10−34
∆p ≥
4π × 5 × 10−15
∆p ≥ 1.05 × 10−20 Ns

Department of Physics 13 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.3. HEISENBERG’S UNCERTAINTY PRINCIPLE

Now the momentum of the electron should have a minimum value equal to this uncer-
tainty, and we will assume that this is the case, i.e., p = 1.05 × 10−20 Ns.
Now, we know the relation between a particle’s kinetic energy and momentum is given
by
p2
Ek =
2m
Now, substituting the value of p from above and m = 9.1 × 10−31 kg (for an electron) in
this expression, we get

(1.05 × 10−20 )2
Ek =
2 × 9.1 × 10−31
Ek = 60.57 × 10−12 J
Ek ≈ 380MeV

We see that, theoretically, the electron must possess 380MeV of energy when it is emitted
from inside the nucleus. But experimentally we had only measured energy in the range of 3
to 4 MeV. This shows that our assumption of considering electron to exist inside the nucleus
is wrong. So the electron does not exist in the nucleus.
(Note: It turns out that the electrons emitted in the β-decay process are due to the decay
of neutrons into protons, antineutrinos, and electrons. An electron as such doesn’t exist in
the nucleus on its own.)

1.3.3 Worked Examples

Q 1. The position and momentum of an electron are simultaneously determined. If its


position is located within 1 Å, find the uncertainty in the determination of its momentum.
Given:
∆x = 1 Å = 1 × 10−10 m
To find: ∆p =?
So we have
h 6.625 × 10−34
∆p = = = 5.272 × 10−25 Ns
4π∆x 4 × 3.14159 × 1 × 10−10

∆p = 5.2720 × 10−25 Ns

Q 2. The position and momentum of an electron with energy 0.5 keV are determined. What is
the minimum percentage uncertainty in its momentum if the uncertainty in the measurement
of position is 0.5 Å.
Given:
Ek = 0.5 keV = 0.5 × 1 × 103 × 1.6 × 10−19 = 8 × 10−17 J
∆x = 5 × 10−11 m
me = 9.11 × 10−31 kg (because electron)
∆p
To find: × 100 =?
p
So first we need to find, p and ∆p. For p,
p
p = 2me Ek = 2 × 9.11 × 10−31 × 8 × 10−17 = 1.207 × 10−23 Ns
p

Department of Physics 14 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.4. WAVEFUNCTION

For ∆p,
h 6.625 × 10−34
∆p = = = 1.054 × 10−24 Ns
4π∆x 4 × 3.14159 × 5 × 10−11
So finally, the minimum percentage uncertainty is,

∆p 1.054 × 10−24
× 100 = × 100 = 8.733%
p 1.207 × 10−23

1.4 Wavefunction
From the de Broglie’s hypothesis, we had associated a wave with a moving particle and called
it a matter wave, and the wavelength of this wave was given by the de Broglie’s wavelength
formula (1.3). This wave can be a function of position and time, just like any other wave
(waves on a string, electromagnetic waves, etc.). The mathematical form of this function
of the wave is called the wavefunction of the particle and is denoted by Ψ(x, t) for the one-
dimensional case.
So far we haven’t given any physical meaning to the wavefunction (as in, what does the
wavefunction actually represent?). We will give one now called the Born’s interpretation of
a wavefunction.
According to Born, the wavefunction (representing a matter wave) is a probability wave
(or probability amplitude). The modulus-square of this probability amplitude |Ψ(x, t)|2 is
called the probability density P (x, t) and it tells us about the probability of finding the
particle at a position x at time t per unit length. (For three dimensions, we have |Ψ(⃗r, t)|2
which is the probability of finding the particle at a position ⃗r at time t per unit volume, hence
the name density.) We take the modulus-square instead of a regular square here because the
wavefunction could be a complex quantity. So the probability density can also be written as

P (x, t) = |Ψ(x, t)|2 = Ψ∗ (x, t)Ψ(x, t)

where Ψ∗ (x, t) is the complex conjugate of Ψ(x, t).


The probability density here is the probability per unit length. To determine actual
probabilities, i.e., a number between 0 and 1, we will consider the term

P (x, t)dx = |Ψ(x, t)|2 dx

which gives the probability P (x, t)dx of finding the particle between positions x and (x + dx)
at time t. (For three dimensions, we have P (⃗r, t)dV = |Ψ(⃗r, t)|2 dV which is the propbability
P (⃗r, t)dV of finding the particle between volumes V and (V + dV ) at time t.)
We know from probability theory that the sum of all the different probabilities correspond-
ing to all possibile events must be equal to 1. Similarly, if there are different probabilties of
finding the particle at different regions in space, then the sum of all these probabilities must
be equal to 1. This is nothing but asserting that the particle has to exist somewhere in space.
This condition is called the normalisation condition and is written as
Z ∞
P (x, t)dx = 1
−∞
Z ∞
=⇒ |Ψ(x, t)|2 dx = 1
−∞

A wavefunction which satisfies this condition is called a normalised wavefunction.

Department of Physics 15 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.5. SCHRODINGER’S TIME-INDEPENDENT WAVE EQUATION

1.4.1 Properties of a wavefunction


An acceptable wavefunction has the following properties:
ˆ The wavefunction must be single valued; because if it has more than one value at a
particular point then it means there is more than one value of probability of finding
the particle at that particular point, which is not possible.

ˆ The wavefunction must be square-integrable, which means the integral of the modulus-
square of the wavefunction must be finite (not zero or infinte). A consequence of this
is the wavefunction must tend to zero as the distance tends to infinity.

ˆ The wavefunction must be continuous everywhere, which means there should be no


sudden jumps of the wavefunction throughout space.

ˆ The wavefunction’s first-order derivatives must also be continous everywhere, which


means there should be no sudden jump of the slope of the wavefunction throughout
space. (Sometimes this condition is neglected in some hypothetical situations.)
Along with these properties, a wavefunction is postulated to contain all the information
that can be known about the system.

1.5 Schrodinger’s Time-Independent Wave Equation


Schrodinger’s equation is the fundamental equation of quantum mechanics. It is the wave
equation describing the behaviour of the wavefunction of a system. Sound waves and waves in
strings are described by the equations of Newtonian mechanics. Light waves are described by
Maxwell’s equations. Similarly, matter waves (wavefunctions) are described by Schrodinger’s
equation.
The wave equation that Schrodinger originally proposed is called the Schrodinger’s time-
dependent wave equation. After making a few assumptions and substitutions to this original
equation, we get the Schrodinger’s time-independent wave equation. We will only learn about
the latter.
Any kind of wave can be described by something called a general wave equation such as

∂ 2 u(x, t) 1 ∂ 2 u(x, t)
= (1.12)
∂x2 v 2 ∂t2
where u(x, t) is the mathematical function representing the wave (like waves on a string,
electromagnetic waves, etc.) and v is the velocity of the wave. (Note that here we are only
considering the one-dimensional case.)
Since (1.12) is an equation for a general wave, we can consider the wavefunction Ψ(x, t)
representing the matter wave to also be described by a similar equation, i.e.,

∂ 2 Ψ(x, t) 1 ∂ 2 Ψ(x, t)
= (1.13)
∂x2 v2 ∂t2
where, again, v is the velocity of the wave (here, matter wave).
Now we will consider a solution to this equation which is of the form

Ψ(x, t) = ψ(x)e−iωt (1.14)

where ω = 2πν is the angular frequency associated


√ with the wave, ν is the frequency of the
wave, and i is the imaginary number, i.e., i = −1. (Note the presence of both forms of the

Department of Physics 16 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.5. SCHRODINGER’S TIME-INDEPENDENT WAVE EQUATION

psi symbol Ψ and ψ. The former is a function of x and t while the latter is a function of only
x.)
Substituting (1.14) into (1.13), we get

∂ 2 (ψ(x)e−iωt ) 1 ∂ 2 (ψ(x)e−iωt )
=
∂x2 v2 ∂t2
2
d ψ(x) ω 2
e−iωt = − 2 e−iωt ψ(x)
dx2 v
2
d ψ(x) ω2
2
= − 2 ψ(x)
dx v
In the second line, we have changed the partial derivative to an ordinary derivative since the
argument (i.e., ψ(x)) now is only a function of x.
Now, we know that ω = 2πν and v = νλ, where λ is the wavelength of the wave. (Here,
λ is the de Broglie wavelength since we are considering matter waves.) Substituting ω and v
in the above equation, we get

d2 ψ(x) 4π 2 ν 2
= − ψ(x)
dx2 ν 2 λ2
d2 ψ(x) 4π 2
= − ψ(x)
dx2 λ2
Now, from de Broglie’s hypothesis, the de Broglie wavelength λ is related to the mo-
mentum p of the particle through the relation (1.3). Substituting this relation in the above
equation, we get
d2 ψ(x) 4π 2 p2
= − ψ(x)
dx2 h2
Now, we know that the kinetic energy of the particle Ek in terms of its momentum p is
given by
p2
Ek =
2m
where m is the mass of the particle.
So we have p2 = 2mEk . Substituting this into the above equation, we get

d2 ψ(x) 8π 2 mEk
= − ψ(x)
dx2 h2
If E and V (x) are the total energy and the potential energy of the particle, respectively,
then we have the relation

E = Ek + V (x) =⇒ Ek = E − V (x)

Here E is a constant and V (x) is a function of position.


Substituting this Ek in the above equation, we get

d2 ψ(x) 8π 2 m(E − V (x))


= − ψ(x)
dx2 h2
Let us use the reduced Planck’s constant (h̄ = h/2π) here. We get

d2 ψ(x) 2m(E − V (x))


=− ψ(x)
dx2 h̄2

Department of Physics 17 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.5. SCHRODINGER’S TIME-INDEPENDENT WAVE EQUATION

After rearranging the terms, we get the following equation

h̄2 d2 ψ(x)
− + V (x)ψ(x) = Eψ(x) (1.15)
2m dx2
This is the Schrodinger’s time-independent wave equation.
When the wavefunction Ψ(x, t) is written like (1.14), i.e., Ψ(x, t) = ψ(x)e−iωt , it is called
a stationary state. It is called so because the probability density for this state does not
change with time. Let’s see how. Take the expression for probability density and substitute
(1.14):

P (x, t) = |Ψ(x, t)|2


= Ψ∗ (x, t)Ψ(x, t)
= ψ ∗ (x)eiωt ψ(x)e−iωt
= ψ ∗ (x)ψ(x)
=⇒ P (x) = |ψ(x)|2

In the last line, we have dropped t in P because now it is independent of time, i.e., we see
here that the probability density is now independent of time for a stationary state.
We can also write (1.15) in the following form

h̄2 d2
 
− + V (x) ψ(x) = Eψ(x)
2m dx2
Ĥψ(x) = Eψ(x) (1.16)

h̄2 d2
where Ĥ = − + V (x) is called the Hamiltonian operator.
2m dx2
Now, (1.16) is called an eigenvalue equation, just like the eigenvalue equations we see in
linear algebra such as AX = λX, where A is a given matrix, X is a column matrix (or a
vector) which is called the eigenvector of A, and λ is a scalar which is called the the eigenvalue
of A. Similarly, in (1.16) Ĥ is a given operator (analogous to matrix A), ψ(x) is called the
eigenfunction of Ĥ and E is called the eigenvalue of Ĥ.
In linear algebra, given an n × n matrix A, we can find the n eigenvalues λi and their
corresponding eigenvectors Xi of A using the charactersistic equation det(A − λI) = 0 (where
I is an n × n identity matrix). Then the eigenvalue equation for each eigenvector is AXi =
λi Xi , where i = 1, 2, . . . n. First we get the various eigenvalues corresponding to A, such
as λ1 , λ2 , λ3 , . . . , λn . Then corresponding to each of these eigenvalues, we get the respective
eigenvectors such as X1 , X2 , X3 , . . . , Xn .
We can do a similar thing for (1.16). The method is a little different (because we have an
operator here instead of an n×n matrix) but the end result is the same, i.e., given an operator
Ĥ, we can find both the eigenvalues (E1 , E2 , E3 , . . .) and the corresponding eigenfunctions
(ψ1 (x), ψ2 (x), ψ3 (x), . . .) of Ĥ.
To solve (1.16), we will need information about the potential energy V (x) in the corre-
sponding system. So knowing the system’s mass m and potential energy, we can solve the
above eigenvalue equation. This requirement of a mass and potential energy to solve an
equation is similar to the requirement of mass and forces to solve Newton’s second law of
motion (ma = F ). So Schrodinger’s equation can be regarded as the quantum mechanical
counterpart to the classical Newton’s second law.

Department of Physics 18 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.5. SCHRODINGER’S TIME-INDEPENDENT WAVE EQUATION

In the case of three dimensions, (1.15) takes the form


h̄2 2
− ∇ ψ(⃗r) + V (⃗r)ψ(⃗r) = Eψ(⃗r)
2m
where ∇2 is the three-dimensional counterpart of the one-diemensional second-order deriva-
tive operator called the Laplacian operator and is given by
∂2 ∂2 ∂2
∇2 = + +
∂x2 ∂y 2 ∂z 2

Expectation Value
Let’s say we are rolling a pair of dice and noting the sum of the two numbers. These readings
can range from 2, 3, 4, upto 12. Using a simple procedure, we can count the different possible
ways of getting these different readings. For example, there is only one way to get a 2 (1+1),
whereas there are three ways to get a 4 (1+3, 2+2, 3+1). The total number of all possible
combinations is 6 × 6 = 36. Then, we can say that 1/36 is the probability of getting a 2;
4/36 is the probability of getting a 4, and so on. In general, let’s say P (a) is the probability
of getting a, where a = 2, 3, 4, . . . 12. So P (2) = 1/36, etc.
Now, the mean value ⟨a⟩ of all the readings is given by the expression
12
X
⟨a⟩ = aP (a)
a=2

Solving this explicitly, we find that the mean value is equal to 7. This means that the
reading 7, on average, is relatively more likely to come up when physically rolling the dice.
That is why the mean value in this context is also called the expectation value. It is important
to note that if the expectation value is a certain reading, then it does not imply the other
readings won’t come up. For example, in our case, even though we got an expectation value
of 7, it does not mean that we won’t get any other number when physically rolling a dice.
The idea of expectation value in quantum mechanics is similar to the case as described
above. Instead of rolling a dice and taking a reading, we perform an experiment and make a
measurement. As the readings of a dice come up with different probabilities, so do the mea-
surement values of the experiment. For example, if we perform an experiment to determine
the position of a particle, the first measurement might yield some value with some probability
whereas the second measurement might yield some other value with its corresponding prob-
ability. (Note that the different measurements being made here are on different but similarly
prepared states of the particle. We are not making measurements on the same state one after
the other.)
So if P (x, t) is the probability of the particle to be at position x at time t, then we can
find the expectation value ⟨x⟩(t) of x as a function of time using a similar expression as
that mentioned above, but with a minor change. The above expression was for a discrete
quantity (a = 2, 3, 4, . . .), whereas here we have a continous quantity (for example, x = 10,
3.1, -22.678, 0.00001 or any real number). So we need to change the summation symbol into
an integral as follows Z ∞
⟨x⟩(t) = xP (x, t)dx
−∞
But we know that probailities are given by the square-modulus of the particle’s wave-
function. So we can finally write the expectation value of position as
Z ∞
⟨x⟩(t) = x|Ψ(x, t)|2 dx
−∞

Department of Physics 19 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.5. SCHRODINGER’S TIME-INDEPENDENT WAVE EQUATION

If the wavefunction is a stationary state, then the expectation value is independent of


time. For example, if we consider position, its expectation value for a stationary state is
given by Z ∞
⟨x⟩ = x|ψ(x)|2 dx
−∞

1.5.1 Application of Schrodinger’s time-independent wave equation


Particle in an infintely deep potential well (Particle in a 1D box)
Consider a particle of mass m present in one-dimensional space. We assume the potential
energy present in this space to be as follows
(
0 when 0 ≤ x ≤ a
V (x) =
∞ when x < 0 and x > a

The plot of V (x) against x is as shown in figure 1.3.

Figure 1.3: Infinitely deep potential well of width a

Considering each region of the potential energy plot seperately, we will solve the equation

h̄2 d2 ψ(x)
− + V (x)ψ(x) = Eψ(x) (1.17)
2m dx2
First, consider the region x < 0 and x > a. Here V = ∞. Substituting this V in (1.17),
we get
h̄2 d2 ψ(x)
− + ∞ψ(x) = Eψ(x)
2m dx2
This equation will hold only if we take ψ(x) = 0. (If we pick a non-zero value, then the
LHS will be infinite whereas the RHS will be finite since E is a constant.)
So we have ψ(x) = 0 in the region x < 0 and x > a. This implies that there is zero
probability of finding the particle in this region, which makes sense because no particle can
exist in an infinte potential region (if it did, the energy of the particle would be infinite,
which is impossible).

Department of Physics 20 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.5. SCHRODINGER’S TIME-INDEPENDENT WAVE EQUATION

Now consider the region 0 ≤ x ≤ a. Here V = 0. Substituting this value of V in (1.17),


we get
h̄2 d2 ψ(x)
− = Eψ(x)
2m dx2
d2 ψ(x) 2mE
2
= − 2 ψ(x)
dx h̄
d2 ψ(x)
= −k 2 ψ(x)
dx2
where
2mE
k2 = (1.18)
h̄2
So we have the following differential equation
d2 ψ(x)
+ k 2 ψ(x) = 0
dx2
The solution to this differential equation is of the form

ψ(x) = A sin(kx) + B cos(kx) (1.19)

where A and B are arbitrary constants which have to be determined. Remember that we
need to also determine k since it contains E, which is one of the things we’re trying to find
(eigenvalue of the eigenvalue equation (1.17)). We’ll find A, B, and k using the boundary
conditions and the wavefunction normalisation condition.
So what are the boundary conditions? To figure this out, we’ll have to use the property
of the wavefunction which states the wavefunction has to be continuous everywhere. Since
we already know that ψ(x) = 0 outside of the region we are considering now, ψ(x) will have
to be zero even at x = 0 and x = a for it to be continuous. (If it is not zero at these two
points, then there will be a discontinuity between the two regions which is not acceptable for
a wavefunction.) So our boundary conditions are ψ(0) = 0 and ψ(a) = 0.
Using the first boundary condition, we have

ψ(0) = A sin(0) + B cos(0)


0=0+B
=⇒ B = 0

So (1.19) now becomes


ψ(x) = A sin(kx) (1.20)
Using the second boundary condition, we have

ψ(a) = A sin(ka)
0 = Asin(ka)
=⇒ A = 0 or sin(ka) = 0

If A = 0, then from (1.20), we get ψ(x) = 0, which means the particle has zero probability
of being in this region (0 ≤ x ≤ a), which of course is not the case. So A ̸= 0, which implies

sin(ka) = 0
=⇒ ka = nπ

k=
a

Department of Physics 21 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.5. SCHRODINGER’S TIME-INDEPENDENT WAVE EQUATION

where n = 1, 2, 3, . . . is any positive integer. (Mathematically, n = 0 is also valid, but we


don’t consider it because it will give ψ(x) = 0.)
Since the value of k depends on n, we should write it as kn ,

kn = (1.21)
a
As we have determined k (now called kn ), we can determine the eigenvalue E (now called
En ). Substituting (1.21) in (1.18) (after changing k to kn ), we get
n2 π 2 2mEn
=
a 2
h̄2
n2 π 2 h̄2 n2 h2
=⇒ En = or (1.22)
2ma2 8ma2
These En (E1 , E2 , E3 , . . .) are called the allowed energy eigenvalues (energy levels) for the
given system. And we see that these energy levels are quantised.
Setting n = 1 in (1.22) gives E1 , which is called the ground state (or zero-point) energy
level
π 2 h̄2
E1 =
2ma2
We can also write all other energy levels in terms of E1 as follows

En = n2 E1

So E2 = 4E1 , E3 = 9E1 , E4 = 16E1 , etc.


Also note that the lowest energy (i.e., E1 ) in this system is not zero. This is due to
the Heisenberg’s uncertainity principle. It is a general result in quantum mechanics that
the lowest energy of any kind of system is non-zero. This lowest energy is also called the
zero-point energy. (If E is zero, then it means that ∆E is also zero, which is prohibited by
the uncertainty principle.)

Figure 1.4: Energy levels of a 1D box

The seperation between consecutive energy levels increases as n increases and is as shown
in figure 1.4.

Department of Physics 22 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.5. SCHRODINGER’S TIME-INDEPENDENT WAVE EQUATION

Now, remember that we still need to find A. (1.20) becomes


 nπx 
ψn (x) = A sin
a
(We changed ψ(x) to ψn (x) since it depends on n.)
We say that ψ1 (x), ψ2 (x), ψ3 (x), . . . are the energy eigenfunctions corresponding to the
energy eigenvalues E1 , E2 , E3 , . . . respectively.
The complete solution to this problem can be written as

0  nπx  when x < 0


ψn (x) = A sin when 0 ≤ x ≤ a (1.23)

 a
0 when x > a
To find A, we will use the normalisation condition
Z ∞
|Ψ(x, t)|2 dx = 1
−∞
But since we are using stationary states, the normalisation condition reduces to
Z ∞
|ψ(x)|2 dx = 1
−∞
Now, note that we have to split the integral into three parts corresponding to the three
regions in our solution (1.23), i.e.,
Z ∞ Z 0 Z a Z ∞
2 2 2
|ψn (x)| dx = |ψn (x)| dx + |ψn (x)| dx + |ψn (x)|2 dx
−∞ −∞ 0 a

Now, the first and the third integral on the RHS vanish because ψn (x) = 0 in those
regions. Therefore we are left with the following normalisation condition
Z a
|ψn (x)|2 dx = 1
0
Substituting ψn (x) here from (1.23), we get
Z a  nπx 
|A|2 sin2 dx = 1
0 a
Here, we use the trigonometric identity
1 − cos 2θ
sin2 θ =
2

a
|A|2
Z  
2nπx
1 − cos dx = 1
2 0 a
  a
2nπx

sin
|A|2 
x − a 
 =1
2  2nπ 
a 0
|A|2  a 
a− sin(2nπ) − 0 + 0 = 1
2 2nπ
a 2
|A| = 1
2 r
2
A=
a

Department of Physics 23 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.5. SCHRODINGER’S TIME-INDEPENDENT WAVE EQUATION

where we have used the fact that sin(2nπ) = 0 for any integer n.
So the complete solution to our eigenvalue equation (1.16) is

n2 π 2 h̄2
En = 2
(1.24)
2ma


 0 when x < 0
r
ψn (x) = 2  nπx 
(1.25)
sin when 0 ≤ x ≤ a


 a a
0 when x > a
r
2  nπx 
The wavefunction ψn (x) = sin is called the normalised wavefunction.
a a
The eigenfuction
r corresponding to n = 1 is called the ground state and is given by
2  πx 
ψ1 (x) = sin . The eigenfunction corresponding to n = 2 is called the first excited
a a r  
2 2πx
state and is given by ψ2 (x) = sin . Similarly we have the second, third, fourth
a a
etc., excited states for n = 3, 4, 5, etc., respectively.
In summary, each energy eigenvalue corresponds to one energy eigenfunction
r
π 2 h̄2 2  πx 
E1 = =⇒ ψ 1 (x) = sin
2ma2 a a
2
r
2
 
4π h̄ 2 2πx
E2 = =⇒ ψ2 (x) = sin
2ma2 a a
2
r
2
 
9π h̄ 2 3πx
E3 = =⇒ ψ3 (x) = sin
2ma2 a a
..
.

The plots of ψn (x) vs x and |ψn (x)|2 vs x for n = 1, 2, 3, 4 are shown in figure 1.5 and
figure 1.6 respectively.
From figure 1.6, we see that for n = 1, there is higher propbability for the particle to be
present near the center of the box than at the ends. But for n = 2, there is lower probability
for the particle to be near the center (it is infact zero exactly at the center) whereas the
probability is higher near x = a/4 and x = 3a/4. Similar observations can be made for
n = 3, 4 states.

1.5.2 Worked Examples

Q 1. An electron is bound in a one dimensional potential well of width 0.18 nm. Find the
energy value in eV of the second excited state.
Given:
a = 0.18 × 10−9 m = 1.8 × 10−10 m
me = 9.11 × 10−31 kg (because electron)
n = 3 (second excited state)
To find: E3 =?
We know that,
n2 h2
En =
8me a2

Department of Physics 24 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.5. SCHRODINGER’S TIME-INDEPENDENT WAVE EQUATION

Figure 1.5: ψn (x) vs. x Figure 1.6: |ψn (x)|2 vs. x

2
(3)2 × 6.625 × 10−34
E3 (in J) = = 1.673 × 10−17 J
8 × 9.11 × 10−31 × (1.8 × 10−10 )2
In terms of eV,

E3 (in J) 1.673 × 10−17


E3 (in eV) = = = 104.6 eV
e 1.6 × 10−19

E3 (in eV) = 104.6 eV

Department of Physics 25 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.6. MODEL AND PREVIOUS YEAR QUESTIONS

1.6 Model and Previous Year Questions

Q 1. State and explain de-Broglie’s hypothesis and derive the expression for de-Broglie
wavelength by analogy.
Q 2. What is wave packet? Give physical significance and properties of wave function.
Q 3. Define phase velocity and group velocity.
Q 4. State and explain Heisenberg’s Uncertainty principle.
Q 5. State and explain the Principle of Complementarity.
Q 6. State and explain Heisenberg’s uncertainty principle. Give its physical significance.
Show that electron cannot exist inside the nucleus.
Q 7. Setup one dimensional time-independent Schrodinger wave equation.
(or) Setup Schrodinger time independent wave equation in one dimension.
(or) Derive an expression for Schrödinger’s Time independent equation in one dimensional
form.
Q 8. Obtain the expression for energy eigen values using Schrodinger’s time independent
equation.
(or) Assuming the time independent Schrodinger’s wave equation discuss the solution for a
particle in a one dimensional potential well of infinite height and hence obtain the normalized
wave equation.
(or) Explain Eigen functions and Eigen Values and hence derive the eigen function of a
particle inside infinite potential well of width a using the method of normalization.
(or) Discuss the motion of a quantum particle in a one-dimensional infinite potential well of
width a and also obtain the eigen functions and energy eigen states.

1.7 Numericals
1.7.1 de Broglie Wavelength

Q 1. Calculate the wavelength associated with an electron having kinetic energy 100 eV.
Q 2. Calculate de Broglie wavelength associated with neutron of mass 1.674 × 10−27 kg with
(1/10)th the speed of light.
Q 3. Find the de Broglie wavelength of an electron accelerated through a potential difference
of 182 V and object of mass 1 kg moving with a speed of 1 m/s. Compare the results and
comment.
Q 4. Estimate thc potential difference through which an electron is needed to be accelerated
so that its de-Broglie wavelength becomes equal to 20 Å.

1.7.2 Heisenberg’s Uncertainty Principle

Q 1. The speed of electron is measured to within an uncertainty of 2.2 × 104 m/s in one
dimension. What is the minimum width required by the electron to be confined in an atom?
Q 2. An electron is confined to a box of length 1 nm, calculate the minimum uncertainty in
its velocity.

Department of Physics 26 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.7. NUMERICALS

Q 3. The inherent uncertainty in the measurement of time spent by Iridium-191 nuclei in


the excited state is found to be 1.4 × 10−10 s. Estimate the uncertainty that results in its
energy in eV in the excited state.
Q 4. An electron has a speed of 100 m/s. The inherent uncertainity in its measurement is
0.005 %. Calculate the corresponding uncertainity in the measurement of the position.
Q 5. In a measurement of position and velocity of an electron moving with a speed of
6 × 105 m/s, calculate highest accuracy with which its position could be determined, if the
inherent error in the measurement of its velocity is 0.01% for the speed stated.

1.7.3 Particle in a 1D box

Q 1. An electron is bound in a one dimensional infinite potential well of width 0.12 nm.
Find the energy value and de-Broglie wavelength in the first excited level.
Q 2. The first excited state energy of an electron in an infinite well is 240 eV. What will be
its ground state energy when the width of the potential well is doubled?
Q 3. A quantum particle confined to a one-dimensional box of width a is in its first ex-
cited state. What is the probability of finding the particle over an interval of a/2 marked
symmetrically at the center of the box.

Department of Physics 27 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju

You might also like