EEE 540 - Nanoelectronics Technology
Autumn 2022
Lecture – 2
Quantum
Mechanics of
Semiconductors
Faculty
Dr. Mustafa Habib Chowdhury
IUB Dr. Mustafa H. Chowdhury 1
Why Quantum Mechanics ?
EEE540_L2
• Quantum Mechanics is the new physics. To be a good engineer, you really need to
have some fundamental understanding of basic quantum mechanics.
• Quantum mechanical properties become important when we start working with small
objects, which typically have the dimensions of nanometers or smaller and quantum
mechanics plays a key role the operation of modern electronic devices.
• The operation of the fundamental electronic devices, such as transistors and diodes, is
described by a combination classical mechanics and quantum mechanical principles.
• In the year 2016, the latest generation MOSFET transistor in production, which is the
key transistor building block of most electronics, has a gate length on the order of 10
nm.
• Other devices are even smaller, and have critical dimensions of a few nanometers or
less, and thus are strongly in- fluenced by the principles of quantum mechanics.
• Electrons and holes, the basic charge carriers in electronics are quantum mechanical
entities.
• While virtually all electronic devices are influenced by quantum mechanics, some
devices operate virtually totally on quantum mechanical principles. These devices
include Flash Memory Sticks, Solar Cells and solid state Lasers.
IUB Dr. Mustafa H. Chowdhury 2
Brief History of Quantum Mechanics
EEE540_L2
• From a historical perspective, quantum mechanics started to come into
being in the early 20th century.
•Then, for the next one hundred years, most discovery in physical
science has been guided by quantum mechanical issues.
• Quantum Mechanics arose to satisfy the need for a new science to
describe the observations that were being discovered about one hundred
years ago.
• Two of these key areas of the discovery were the Photo Electric Effect
and the Light Emission Spectra of Atoms.
• These discoveries opened up the fundamental quantum mechanical
properties that small particles, like electrons, exhibit both wavelike and
particle-like behavior, and that light, long thought of as a wave, can
exhibit both particle and wave-like properties as well.
IUB Dr. Mustafa H. Chowdhury 3
Solvay Conference 1927
EEE540_L2
The most famous conference was the fifth Solvay Conference on Physics, which was held from 24 to 29 October 1927; the subject was Electrons and Photons and the
world's most notable physicists met to discuss the newly formulated quantum theory. The leading figures were Albert Einstein and Niels Bohr. Seventeen of the 4
29 attendees
were or became Nobel Prize winners, including Marie Curie who, alone among them, had won Nobel Prizes in two separate scientific disciplines. Attendees Albert Einstein,
Niels Bohr, Werner Heisenberg, Paul Dirac, and Erwin Schrödinger would be listed among the top ten greatest physicists of all-time.
Introduction
EEE540_L2
The triumph of modern physics is the triumph of quantum mechanics. Even the
simplest experimental observation that the resistivity of a metal depends linearly
on the temperature can only be explained by quantum physics.
One of the most important discoveries in physics has been the wave–particle
duality of nature.
The electron, which we have so far considered to be a particle and hence to be
obeying Newton’s second law (F = ma), can also exhibit wave-like properties
quite contrary to our intuition.
An electron beam can give rise to diffraction patterns and interference fringes,
just like a light wave. Interference and diffraction phenomena displayed by
light can only be explained by treating light as an electromagnetic wave.
But light can also exhibit particle-like properties in which it behaves as if it were
a stream of discrete entities (“photons”), each carrying a linear momentum and
each interacting discretely with electrons in matter (just like a particle colliding
with another particle).
IUB Dr. Mustafa H. Chowdhury 5
Photons
EEE540_L2
A ray of light is considered to be an electromagnetic (EM) wave with a given
frequency, as depicted in Figure 3.1.
The electric and magnetic fields, Ey and Bz, of this wave are perpendicular to
each other and to the direction of propagation x. The electric field Ey at position
x at time t may be described by:
IUB Dr. Mustafa H. Chowdhury 6
Photons
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 7
Photons
EEE540_L2
Understanding the wave nature of light is fundamental to understanding
interference and diffraction, two phenomena that we experience with sound
waves almost on a daily basis.
Figure 3.2 illustrates how the interference of secondary waves from the two
slits S1 and S2 gives rise to the dark and bright fringes (called Young’s fringes)
on a screen placed at some distance from the slits.
IUB Dr. Mustafa H. Chowdhury 8
Photons
EEE540_L2
• At point P on the screen, the waves emanating from S1 and S2 interfere
constructively, if they are in phase.
• This is the case if the path difference between the two rays is an integer multiple
of the wavelength λ, or:
where n is an integer. If the two waves are out of phase by a path difference of λ∕2, or:
• then the waves interfere destructively and the intensity at point P vanishes.
• Thus, in the y direction, the observer sees a pattern of bright and dark fringes.
IUB Dr. Mustafa H. Chowdhury 9
Photons
EEE540_L2
• When X-rays are incident on a crystalline material, they give rise to typical
diffraction patterns on a photographic plate, as shown in Figure 3.3a and b, which
can only be explained by using wave concepts.
• For simplicity, consider two waves, 1 and 2, in an X-ray beam. The waves are
initially in phase, as shown in Figure 3.3c.
IUB Dr. Mustafa H. Chowdhury 10
Photons
EEE540_L2
• Suppose that wave 1 is “reflected” from the first plane of atoms in the crystal,
whereas wave 2 is “reflected” from the second plane.
• After reflection, wave 2 has traveled an additional distance equivalent to 2d sin θ
before reaching wave 1.
• The path difference between the two waves is 2d sin θ , where d is the separation
of the atomic planes.
• For constructive interference, this must be nλ, where n is an integer.
• Otherwise, waves 1 and 2 will interfere destructively and will cancel each other.
• Waves reflected from adjacent atomic planes interfere constructively to
constitute a diffracted beam only when the path difference between the waves is
an integer multiple of the wavelength, and this will only be the case for certain
directions.
• Therefore, the condition for the existence of a diffracted beam is:
IUB Dr. Mustafa H. Chowdhury 11
Photons
EEE540_L2
• The condition expressed in Equation 3.3, for observing a diffracted beam, forms
the whole basis for identifying and studying various crystal structures (the science
of crystallography).
• The equation is referred to as Bragg’s law, and arises from the constructive
interference of waves.
• Aside from exhibiting wave-like properties, light can behave like a stream of
“particles” of zero rest-mass.
• The only way to explain a vast number of experiments is to view light as a stream
of discrete entities or energy packets called photons, each carrying a quantum of
energy hf, and momentum h∕λ, where h is a universal constant that can be
determined experimentally, and f is the frequency of light.
• This photonic view of light is drastically different than the simple wave picture
and must be examined closely to understand its origin.
IUB Dr. Mustafa H. Chowdhury 12
Basics of Quantum Mechanics
Autumn 2022
- Photoelectric Effect -
A Photocell is Used to Study the Photoelectric Effect
IUB Dr. Mustafa H. Chowdhury 13
The Photoelectric Effect
Autumn 2022
• If monochromatic light is incident on a clean surface of a material, then
under certain conditions, electrons (photoelectrons) are emitted from the
surface.
• According to classical physics, if the intensity of the light is large
enough, the work function of the material will be overcome and an
electron will be emitted from the surface independent of the of the
incident frequency. This result is NOT observed !!!
• The observed effect is that, at constant incident intensity, the maximum
kinetic energy of the photoelectron varies linearly with frequency with a
limiting frequency v = v0 below which NO photoelectron is produced.
• If the incident intensity varies at a constant frequency, the rate of
photoelectron emission changes, but the maximum kinetic energy of the
generated photoelectron remains the same.
IUB Dr. Mustafa H. Chowdhury 14
The Photoelectric Effect
Autumn 2022
IUB Dr. Mustafa H. Chowdhury 15
The Photoelectric Effect
Autumn 2022
Larger light intensity means larger number of photons at a
given frequency (Energy)
IUB Dr. Mustafa H. Chowdhury 16
The Photoelectric Effect
EEE540_L2
• Consider a quartz glass vacuum tube with two metal electrodes, a photocathode
and an anode, which are connected externally to a voltage supply V (variable and
reversible) via an ammeter, as schematically illustrated in Figure 3.4.
• When the cathode is illuminated with light, if the frequency f of the light is
greater than a certain critical value f0, the ammeter registers a current I, even
when the anode voltage is zero (i.e., the supply is bypassed). When light strikes
the cathode, electrons are emitted with sufficient kinetic energy to reach the
opposite electrode.
IUB Dr. Mustafa H. Chowdhury 17
The Photoelectric Effect
EEE540_L2
• Applying a positive voltage to the anode helps to collect more of the electrons
and thus increases the current, until it saturates because all the photoemitted
electrons have been collected.
• The current, then, is limited by the rate of supply of photoemitted electrons.
• If, on the other hand, we apply a negative voltage to the anode, we can “push”
back the photoemitted electrons and hence reduce the current I.
• Figure 3.5a shows the dependence of the photocurrent on the anode voltage, for
one particular frequency of light.
IUB Dr. Mustafa H. Chowdhury 18
The Photoelectric Effect
EEE540_L2
• Recall that when an electron traverses a voltage difference V, its potential energy
changes by eV (potential difference is defined as work done per unit charge).
• When a negative voltage is applied to the anode, the electron has to do work to
get to this electrode, and this work comes from its kinetic energy just after
photoemission.
• When the negative anode voltage V is equal to V0, which just “extinguishes” the
current I, we know that the potential energy “gained” by the electron is just the
kinetic energy lost by the electron, or:
• where v is the velocity and KEm is the kinetic energy of the electron just after
photoemission.
• Therefore, we can conveniently measure the maximum kinetic energy KEm of an
emitted electron.
IUB Dr. Mustafa H. Chowdhury 19
The Photoelectric Effect
EEE540_L2
• For a given frequency of light, increasing the intensity of light I requires the
same voltage V0 to extinguish the current; that is, the KEm of the emitted
electrons is independent of the light intensity I. This is quite surprising.
• However, increasing the intensity does increase the saturation current. Both of
these effects are noted in the I–V results shown in Figure 3.5a.
• Since the magnitude of the saturation photocurrent depends on the light
intensity I, whereas the KE of the emitted electron is independent of I, we are
forced to conclude that only the number of electrons ejected depends on the light
intensity.
• Furthermore, if we plot KEm (from the V0 value) against the light frequency f for
different electrode metals for the cathode, we find the typical behavior shown in
Figure 3.6.
• This shows that the KE of the emitted electron depends on the frequency of
light.
IUB Dr. Mustafa H. Chowdhury 20
The Photoelectric Effect
EEE540_L2
• The experimental results shown in Figure 3.6 can be summarized by a statement
that relates the KEm of the electron to the frequency of light and the electrode
metal, as follows:
• where h is the slope of the straight line and is independent of the type of metal,
whereas f0 depends on the electrode material for the photocathode (e.g., f01, f02,
etc.).
IUB Dr. Mustafa H. Chowdhury 21
The Photoelectric Effect
EEE540_L2
• Equation 3.4 is essentially a succinct statement of the experimental observations of
the photoelectric effect as exhibited in Figure 3.6.
• The constant h is called Planck’s constant, which, from the slope of the straight
lines in Figure 3.6, can be shown to be about 6.6 × 10−34 J s.
• The successful interpretation of the photoelectric effect was first given in 1905
by Einstein, who proposed that light consists of “energy packets,” each of which
has the magnitude hf.
• We can call these energy quanta photons.
• When one photon strikes an electron, its energy is transferred to the electron. The
whole photon becomes absorbed by the electron.
• But, an electron in a metal is in a lower state of potential energy (PE) than in
vacuum, by an amount Φ, which we call the work function of the metal, as
illustrated in Figure 3.7.
• The lower PE is what keeps the electron in the metal; otherwise, it would “drop
out.”
IUB Dr. Mustafa H. Chowdhury 22
The Photoelectric Effect
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 23
The Photoelectric Effect
EEE540_L2
• This lower PE is a result of the Coulombic attraction interaction between the
electron and the positive metal ions. Some of the photon energy hf therefore
goes toward overcoming this PE barrier.
• The energy that is left (hf − Φ) gives the electron its KE. The work function Φ
changes from one metal to another. Photoemission only occurs when hf is
greater than Φ.
• This is clearly borne out by experiment, since a critical frequency f0 is needed to
register a photocurrent.
• When f is less than f0, even if we use an extremely intense light, no current
exists because no photoemission occurs, as demonstrated by the experimental
results in Figure 3.6.
• Inasmuch as Φ depends on the metal, so does f0 . Therefore, in Einstein’s
interpretation hf0 = Φ.
• The measurement of f0 constitutes one method of determining the work function of
a metal.
IUB Dr. Mustafa H. Chowdhury 24
The Photoelectric Effect
EEE540_L2
• In the photonic interpretation of light, we still have to resolve the meaning of the
intensity of light, because the classical intensity in Equation 3.2 is obviously not
acceptable.
• Increasing the intensity of illumination in the photoelectric experiment
increases the saturation current, which means that more electrons are emitted
per unit time.
• We therefore infer that the cathode must be receiving more photons per unit time
at higher intensities.
• By definition, “intensity” refers to the amount of energy flowing through a unit
area per unit time.
• The number of photons crossing a unit area per unit time is defined as the photon
flux density, and denoted by Γph.
• The flow of energy through a unit area per unit time, the light intensity, is the
product of this photon flux density and the energy per photon, that is:
IUB Dr. Mustafa H. Chowdhury 25
The Photoelectric Effect
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 26
The Photoelectric Effect
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 27
The Photoelectric Effect
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 28
The Photoelectric Effect
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 29
The Photoelectric Effect
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 30
Compton Scattering
EEE540_L2
• When an X-ray strikes an electron, it is deflected, or “scattered.” In addition, the
electron moves away after the interaction, as depicted in Figure 3.9.
• The wavelength of the incoming and scattered X-rays can readily be measured.
The frequency f ′ of the scattered X-ray is less than the frequency f of the
incoming X-ray.
IUB Dr. Mustafa H. Chowdhury 31
Compton Scattering
EEE540_L2
• When the KE of the electron is determined, we find that:
• Since the electron now also has a momentum pe, then from the conservation of
linear momentum law, we are forced to accept that the X-ray also has a
momentum. The Compton scattering experiments show that the momentum of
the photon is related to its wavelength by:
• We see that a photon not only has an energy hf, but also a momentum p, and it
interacts as if it were a discrete entity like a particle. Therefore, when discussing
the properties of a photon, we must consider its energy and momentum as if it
were a particle.
• We should mention that the description of the Compton effect shown in Figure 3.9
is, in fact, the inference from a more practical experiment involving the scattering
of X-rays from a metal target. A collimated monochromatic beam of X-rays of
wavelength λ0 strikes a conducting target, such as graphite, as illustrated in Figure
3.10a.
IUB Dr. Mustafa H. Chowdhury 32
Compton Scattering
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 33
Compton Scattering
EEE540_L2
• A conducting target contains a large number of nearly “free” electrons (conduction
electrons), which can scatter the X-rays.
• The scattered X-rays are detected at various angles θ with respect to the original
direction, and their wavelength λ′ is measured.
• The result of the experiment is therefore the scattered wavelength λ′ measured at
various scattering angles θ, as shown in Figure 3.10b.
• It turns out that the λ′ versus θ results agree with the conservation of linear
momentum law applied to an X-ray photon colliding with an electron with the
momentum of the photon given precisely by Equation 3.7.
• The photoelectric experiment and the Compton effect are just two convincing
experiments in modern physics that force us to accept that light can have
particle-like properties.
• We already know that it can also exhibit wave-like properties, in such experiments
as Young’s interference fringes. We are then faced with what is known as the
wave–particle dilemma.
• How do we know whether light is going to behave like a wave or a particle?
IUB Dr. Mustafa H. Chowdhury 34
Compton Scattering
EEE540_L2
• The properties exhibited by light depend very much on the nature of the
experiment.
• Some experiments will require the wave model, whereas others may use the
particulate interpretation of light.
• We should perhaps view the two interpretations as two complementary ways of
modeling the behavior of light when it interacts with matter, accepting the fact that
light has a dual nature.
• Both models are needed for a full description of the behavior of light.
• The expressions for the energy and momentum of the photon, E = hf and p = h∕λ,
can also be written in terms of the angular frequency ω (= 2πf ) and the wave
number k, defined as k = 2π∕λ. If we define ħ = h∕2π, then:
IUB Dr. Mustafa H. Chowdhury 35
Compton Scattering
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 36
Compton Scattering
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 37
The Electron as a Wave
EEE540_L2
• Can electrons exhibit wave-like properties? This depends on the experiment and
on the energy of the electrons.
• When the interference and diffraction experiments in Figures 3.2 and 3.3 are
repeated with an electron beam, very similar results are found to those
obtainable with light and X-rays.
• When we use an electron beam in Young’s double-slit experiment, we observe
high- and low-intensity regions (i.e., Young’s fringes), as illustrated in Figure
3.12.
IUB Dr. Mustafa H. Chowdhury 38
The Electron as a Wave
EEE540_L2
• When an energetic electron beam hits a polycrystalline aluminum sheet, it
produces diffraction rings on a fluorescent screen as shown in Figures 3.13a and b
just like X-rays do on a photographic plate.
• When we bring a magnet to the screen, the electrons moving toward the screen
experience a force that would bend their paths, which results in a distorted
diffraction pattern as shown in Figure 3.13c.
IUB Dr. Mustafa H. Chowdhury 39
The Electron as a Wave
EEE540_L2
• An X-ray diffraction pattern, on the other hand, would be unaffected by a
magnetic field.
• If we analyze the diffraction pattern obtained with an electron beam, for
example Figure 3.13b, we would find that the electrons obey the Bragg
diffraction condition 2d sin θ = nλ just as much as the X-ray waves.
• Since we know the interatomic spacing d and we can measure the angle of
diffraction 2θ, we can readily evaluate the wavelength λ associated with
the wave-like behavior of the electrons.
• Furthermore, from the accelerating voltage V in the electron tube, we can
also determine the momentum of the electrons, because the kinetic energy
gained by the electrons, (p2∕2me), is equal to eV.
• Simply by adjusting the accelerating voltage V, we can therefore study
how the wavelength of the electron depends on the momentum.
IUB Dr. Mustafa H. Chowdhury 40
The Electron as a Wave
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 41
The Electron as a Wave
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 42
Time-Independent Schrödinger Equation
EEE540_L2
• The experiments in which electrons exhibit interference and diffraction
phenomena show quite clearly that, under certain conditions, the electron can
behave as a wave; in other words, it can exhibit wave-like properties.
• There is a general equation that describes this wave-like behavior and, with the
appropriate potential energy and boundary conditions, will predict the results of
the experiments.
• The equation is called the Schrödinger equation and it forms the foundations of
quantum theory.
• Its fundamental nature is analogous to the classical physics assertion of Newton’s
second law, F = ma, which of course cannot be proved.
• As a fundamental equation, Schrödinger’s has been found to successfully
predict every observable physical phenomenon at the atomic scale.
• Without this equation, we will not be able to understand the properties of
electronic materials and the principles of operation of many semiconductor
devices.
IUB Dr. Mustafa H. Chowdhury 43
Time-Independent Schrödinger Equation
EEE540_L2
• A traveling electromagnetic wave resulting from sinusoidal current oscillations, or
the traveling voltage wave on a long transmission line, can generally be described
by a traveling-wave equation of the form:
• where E(x) = E0 exp( jkx) represents the spatial dependence, which is separate
from the time variation.
• We assume that no transients exist to upset this perfect sinusoidal propagation. We
note that the time dependence is harmonic and therefore predictable.
• For this reason, in AC circuits we put aside the exp(−jωt) term until we need the
instantaneous magnitude of the voltage.
IUB Dr. Mustafa H. Chowdhury 44
Time-Independent Schrödinger Equation
EEE540_L2
• In 1926, Max Born suggested a probability wave interpretation for the wave-like
behavior of the electron.
• is a plane traveling wavefunction for an electric field; experimentally, we measure
and interpret the intensity of a wave, namely ∣E(x, t)∣2.
• There may be a similar wave function for the electron, which we can represent by
a function Ψ(x, t).
• According to Born, the significance of Ψ(x, t) is that its amplitude squared
represents the probability of finding the electron per unit distance.
• Thus, in three dimensions, if Ψ(x, y, z, t) represents the wave property of the
electron, it must have one of the following interpretations:
IUB Dr. Mustafa H. Chowdhury 45
Time-Independent Schrödinger Equation
EEE540_L2
• If we are just considering one dimension, then the wavefunction is Ψ(x, t), and
∣Ψ(x, t)∣2 dx is the probability of finding the electron between x and (x + dx) at
time t.
• We should note that since only ∣Ψ∣2 has meaning, not Ψ, the latter function need
not be real; it can be a complex function with real and imaginary parts.
• For this reason, we tend to use Ψ* Ψ, where Ψ* is the complex conjugate of Ψ,
instead of ∣Ψ∣2 , to represent the probability per unit volume.
• To obtain the wavefunction Ψ(x, t) for the electron, we need to know how the
electron interacts with its environment. This is embodied in its potential energy
function V = V(x, t), because the net force the electron experiences is given by:
IUB Dr. Mustafa H. Chowdhury 46
Time-Independent Schrödinger Equation
EEE540_L2
• If the PE of the electron is time independent, which means that V = V(x) in one
dimension, then the spatial and time dependences of Ψ(x, t) can be separated, just
as in Equation 3.19, and the total wavefunction Ψ(x, t) of the electron can be
written as:
• where ψ(x) is the electron wavefunction that describes only the spatial behavior,
and E is the energy of the electron.
• The temporal behavior is simply harmonic, by virtue of exp(−jEt∕ħ), which
corresponds to exp(−jωt) with an angular frequency ω = E∕ħ.
• The fundamental equation that describes the electron’s behavior by determining
ψ(x) is called the time-independent Schrödinger equation. It is given by the
famous equation:
IUB Dr. Mustafa H. Chowdhury 47
Time-Independent Schrödinger Equation
EEE540_L2
• This is a second-order differential equation. It should be re-emphasized that the
potential energy V in Equation 3.21a depends only on x.
• If the potential energy of the electron depends on time as well, that is, if V = V(x,
t), then in general Ψ(x, t) cannot be written as ψ(x) exp(−jEt∕ħ).
• Instead, we must use the full version of the Schrödinger equation, which is
discussed in more advanced textbooks/courses.
• In three dimensions, there will be derivatives of ψ with respect to x, y, and z.
• We use the calculus notation (∂ψ∕∂x), differentiating ψ(x, y, z) with respect to x but
keeping y and z constant. Similar notations ∂ψ∕∂y and ∂ψ∕∂z are used for
derivatives with respect to y alone and with respect to z alone, respectively. In
three dimensions, Equation 3.21a becomes:
IUB Dr. Mustafa H. Chowdhury 48
Time-Independent Schrödinger Equation
EEE540_L2
• This Equation 3.21b is a fundamental equation, called the time-independent
Schrödinger equation, the solution of which gives the steady-state behavior of
the electron in a time-independent potential energy environment described by V
= V(x, y, z).
• By solving Equation 3.21b, we will know the probability distribution and the
energy of the electron.
• Once ψ(x, y, z) has been determined, the total wavefunction for the electron is
given by Equation 3.20 so that:
• The time-independent Schrödinger equation can be viewed as a “mathematical
crank.”
• We input the potential energy of the electron and the boundary conditions, turn the
crank, and get the probability distribution and the energy of the electron under
steady-state conditions.
IUB Dr. Mustafa H. Chowdhury 49
Time-Independent Schrödinger Equation
EEE540_L2
• Two important boundary conditions are often used to solve the Schrödinger
equation.
• First, as an analogy, when we stretch a string between two fixed points and put it
into a steady-state vibration, there are no discontinuities or kinks along the string.
We can therefore intelligently guess that because ψ(x) represents wave-like
behavior, it must be a smooth function without any discontinuities.
• The first boundary condition is that Ψ must be continuous, and the second is
that dΨ∕dx must be continuous. In the steady state, these two conditions translate
directly to ψ and dψ∕dx being continuous.
• Since the probability of finding the electron is represented by ∣ψ∣2, this function
must be single-valued and smooth, without any discontinuities, as illustrated in
Figure 3.15.
IUB Dr. Mustafa H. Chowdhury 50
Time-Independent Schrödinger Equation
EEE540_L2
• The enforcement of these boundary conditions results in strict
requirements on the wavefunction ψ(x), as a result of which only certain
wavefunctions are acceptable.
• These wavefunctions are called the eigenfunctions (characteristic
functions) of the system, and they determine the behavior and energy of
the electron under steady-state conditions.
• The eigenfunctions ψ(x) are also called stationary states, inasmuch as we
are only considering steady-state behavior.
• It is important to note that the Schrödinger equation is generally
applicable to all matter, not just the electron.
• For example, the equation can also be used to describe the behavior of a
proton, if the appropriate potential energy V(x, y, z) and mass (mproton) are
used.
• Wavefunctions associated with particles are frequently called matter
waves.
IUB Dr. Mustafa H. Chowdhury 51
Infinite Potential Well: A Confined Electron
EEE540_L2
• Consider the behavior of the electron when it is confined to a certain region, 0 < x
< a. Its PE is zero inside that region and infinite outside, as shown in Figure 3.16.
IUB Dr. Mustafa H. Chowdhury 52
Infinite Potential Well: A Confined Electron
EEE540_L2
• The electron cannot escape, because it would need an infinite PE. Clearly the
probability ∣ψ∣2 of finding the electron per unit volume is zero outside 0 < x < a.
• Thus, ψ = 0 when x ≤ 0 and x ≥ a, and ψ is determined by the Schrödinger
equation in 0 < x < a with V = 0. Therefore, in the region 0 < x < a:
IUB Dr. Mustafa H. Chowdhury 53
Infinite Potential Well: A Confined Electron
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 54
Infinite Potential Well: A Confined Electron
EEE540_L2
• We have already seen this relationship, when we defined k as 2π∕λ (wavenumber)
for a free traveling wave.
• So the constant k here is a wavenumber-type quantity even though there is no
distinct traveling wave. Its value is determined by the boundary condition at x = a
where ψ = 0, or:
• We notice immediately that k, and therefore the energy of the electron, can only
have certain values; they are quantized by virtue of n being an integer. Here, n is
called a quantum number. For each n, there is a special wavefunction:
IUB Dr. Mustafa H. Chowdhury 55
Infinite Potential Well: A Confined Electron
EEE540_L2
• We still have not completely solved the problem, because A has yet to be
determined.
• To find A, we use what is called the normalization condition. The total
probability of finding the electron in the whole region 0 < x < a is unity, because
we know the electron is somewhere in this region.
• Thus, ∣ψ∣2 dx summed between x = 0 and x = a must be unity, or:
IUB Dr. Mustafa H. Chowdhury 56
Infinite Potential Well: A Confined Electron
EEE540_L2
• We can now summarize the behavior of an electron in a one-dimensional PE well.
• Its wavefunction and energy, shown in Figure 3.16, are given by Equations 3.29
and 3.28, respectively. Both depend on the quantum number n.
• The energy of the electron increases with n2, so the minimum energy of the
electron corresponds to n = 1. This is called the ground state, and the energy of
the ground state is the lowest energy the electron can possess.
• Note also that the energy of the electron in this potential well cannot be zero,
even though the PE is zero. Thus, the electron always has KE, even when it is in
the ground state.
IUB Dr. Mustafa H. Chowdhury 57
Infinite Potential Well: A Confined Electron
EEE540_L2
• The node of a wavefunction is defined as the point where ψ = 0 inside the well.
• It is apparent from Figure 3.16 that the ground wavefunction ψ1 with the lowest
energy has no nodes, ψ2 has one node, ψ3 has two nodes, and so on. Thus, the
energy increases as the number of nodes increases in a wavefunction.
• It may seem surprising that the energy of the electron is quantized; that is, it can
only have finite values, given by Equation 3.28.
• The electron cannot be made to take on any value of energy, as in the classical
case. If the electron behaved like a classical particle, then an applied force F could
impart any value of energy to it, because F = dp∕dt (Newton’s second law), or p =
∫ F dt.
• By applying a force F for a time t, we can give the electron a KE of:
IUB Dr. Mustafa H. Chowdhury 58
Infinite Potential Well: A Confined Electron
EEE540_L2
• As a increases to macroscopic dimensions, a → ∞, the electron is completely free
and ΔE → 0.
• Since ΔE = 0, the energy of a completely free electron (a = ∞) is continuous. The
energy of a confined electron, however, is quantized, and ΔE depends on the
dimension (or size) of the potential well confining the electron.
• In general, an electron will be “contained” in a spatial region of three dimensions,
within which the PE will be lower (hence the confinement).
• We must then solve the Schrödinger equation in three dimensions. The result is
three quantum numbers that characterize the behavior of the electron.
• Examination of the wavefunctions ψn in Figure 3.16 shows that these are either
symmetric or antisymmetric with respect to the center of the well at x = (1/2)a.
IUB Dr. Mustafa H. Chowdhury 59
Infinite Potential Well: A Confined Electron
EEE540_L2
• The symmetry of a wavefunction is called its parity.
• Whenever the potential energy function V(x) exhibits symmetry about a certain
point C, for example, about x = (1/2)a in Figure 3.16, then the wavefunctions
have either even parity (such as ψ1, ψ3, . . . That are symmetric) or have odd
parity (such as ψ2, ψ4, . . . that are antisymmetric) with respect to C.
• An electron in an infinite PE well would normally occupy the lowest state that
corresponds to ψ1 with an energy E1. The next possible state is ψ2 at an energy E2.
• The electron in the ground state at E1 can be excited to E2 by the absorption
of a photon of exactly the energy E2 − E1, which corresponds to a radiation
frequency f12 such that hf12 = E2 − E1.
• An electromagnetic radiation that is incident on the quantum well and has the
right frequency f12 will be absorbed by the electron at E1, which will be excited to
the energy level E2.
IUB Dr. Mustafa H. Chowdhury 60
Infinite Potential Well: A Confined Electron
EEE540_L2
• In this particular case, as apparent from Figure 3.16, the excitation of the electron
by the absorption of a photon from ψ1 to ψ2 involves a change in the parity of the
wavefunction, from even to odd. This observation turns out to be generally true.
• Whenever an electron in a quantum well absorbs or emits electromagnetic
radiation, its parity must change.
• Those transitions in which the parity does not change have a low probability
(but not zero) of occurrence and are usually called forbidden transitions.
• For example, suppose the electron is in a state ψ3 at an energy level E3. It can emit
a photon and decay down from E3 to E2 or E1.
• It will transit down to E2 because the parity of its wavefunction will change, and
this transition has a much higher probability.
• The transition from ψ3 to ψ1 is “forbidden” and the probability of its occurrence
is low. From E2, the electron will decay down to E1.
IUB Dr. Mustafa H. Chowdhury 61
Infinite Potential Well: A Confined Electron
EEE540_L2
Planck's constant h = 6.62607015 × 10−34 J⋅s OR 4.135667696 × 10−15 eV⋅s
Reduced Planck's constant ħ = 1.054571817 × 10−34 J⋅s OR 6.582119569 × 10−16 eV⋅s
Mass of an electron me = 9.1 ×10−31 kg
1 eV=1.602×10−19 J
1 J = 6.242×1018 eV
IUB Dr. Mustafa H. Chowdhury 62
Infinite Potential Well: A Confined Electron
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 63
Infinite Potential Well: A Confined Electron
EEE540_L2
Planck's constant h = 6.62607015 × 10−34 J⋅s OR 4.135667696 × 10−15 eV⋅s
Reduced Planck's constant ħ = 1.054571817 × 10−34 J⋅s OR 6.582119569 × 10−16 eV⋅s
Mass of an electron me = 9.1 ×10−31 kg
1 eV=1.602×10−19 J
1 J = 6.242×1018 eV
IUB Dr. Mustafa H. Chowdhury 64
Infinite Potential Well: A Confined Electron
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 65
Infinite Potential Well: A Confined Electron
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 66
Heisenberg’s Uncertainty Principle
EEE540_L2
• The wavefunction of a free electron corresponds to a traveling wave with a single
wavelength λ, as shown in Example 3.6.
• The traveling wave extends over all space, along all x, with the same amplitude, so
the probability distribution function is uniform throughout the whole of space.
• The uncertainty Δx in the position of the electron is therefore infinite.
• However, the uncertainty Δpx in the momentum of the electron is zero, because λ
is well-defined, which means that we know px exactly from the de Broglie
relationship, px = h∕λ.
IUB Dr. Mustafa H. Chowdhury 67
Heisenberg’s Uncertainty Principle
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 68
Heisenberg’s Uncertainty Principle
EEE540_L2
• The product of the position and momentum uncertainties is simply h. This
relationship is fundamental; and it constitutes a limit to our knowledge of the
behavior of a system.
• We cannot exactly and simultaneously know both the position and momentum of
a particle along a given coordinate.
• In general, if Δx and Δpx are the respective uncertainties in the simultaneous
measurement of the position and momentum of a particle along a particular
coordinate (such as x), the Heisenberg uncertainty principle states that:
• We are therefore forced to conclude that as previously stated, because of the
wave nature of quantum mechanics, we are unable to determine exactly and
simultaneously the position and momentum of a particle along a given
coordinate.
• There will be an uncertainty Δx in the position and an uncertainty Δpx in the
momentum of the particle and these uncertainties will be related by Heisenberg’s
uncertainty relationship in Equation 3.31.
IUB Dr. Mustafa H. Chowdhury 69
Heisenberg’s Uncertainty Principle
EEE540_L2
• These uncertainties are not in any way a consequence of the accuracy of a
measurement or the precision of an instrument.
• Rather, they are the theoretical limits to what we can determine about a system.
They are part of the quantum nature of the universe.
• In other words, even if we build the most perfectly engineered instrument to
measure the position and momentum of a particle at one instant, we will still be
faced with position and momentum uncertainties Δx and Δpx such that Δx Δpx > ħ.
• There is a similar uncertainty relationship between the uncertainty ΔE in the
energy E (or angular frequency ω) of the particle and the time duration Δt
during which it possesses the energy (or during which its energy is measured).
• We know that the kx part of the wave leads to the uncertainty relation ΔxΔpx > ħ or
Δx Δk ≥ 1 (px = hkx) .
• By analogy we should expect a similar relationship for the ωt part, or Δω Δt ≥ 1.
• This hypothesis is true, and since E = ħω, we have the uncertainty relation for the
particle energy and time:
IUB Dr. Mustafa H. Chowdhury 70
Heisenberg’s Uncertainty Principle
EEE540_L2
• Note that the uncertainty relationships in Equations 3.31 and 3.32 have been
written in terms of ħ, rather than h, as implied by the electron in an infinite
potential energy well (Δx Δpx ≥ h).
• In general, there is also a numerical factor of ½ multiplying ħ in Equations 3.31
and 3.32 which comes about when we consider a Gaussian spread for all possible
position and momentum values.
• It is important to note that the uncertainty relationship applies only when the
position and momentum are measured in the same direction (such as the x
direction).
• On the other hand, the exact momentum, along, say, the y direction and the
exact position, along, say, the x direction can be determined exactly, since Δx
Δpy need not satisfy the Heisenberg uncertainty relationship (in other words, Δx
Δpy can be zero).
IUB Dr. Mustafa H. Chowdhury 71
Heisenberg’s Uncertainty Principle
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 72
Heisenberg’s Uncertainty Principle
EEE540_L2
Planck's constant h = 6.62607015 × 10−34 J⋅s OR 4.135667696 × 10−15 eV⋅s
Reduced Planck's constant ħ = 1.054571817 × 10−34 J⋅s OR 6.582119569 × 10−16 eV⋅s
Mass of an electron me = 9.1 ×10−31 kg
IUB Dr. Mustafa H. Chowdhury 73
Heisenberg’s Uncertainty Principle
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 74
Heisenberg’s Uncertainty Principle
EEE540_L2
Planck's constant h = 6.62607015 × 10−34 J⋅s OR 4.135667696 × 10−15 eV⋅s
Reduced Planck's constant ħ = 1.054571817 × 10−34 J⋅s OR 6.582119569 × 10−16 eV⋅s
IUB Dr. Mustafa H. Chowdhury 75
Heisenberg’s Uncertainty Principle
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 76
Confined Electron in a Finite Potential Energy Well
EEE540_L2
• When the electron is contained in a finite PE well as shown in Figure 3.17a, due to
the confinement, the electron energy is again quantized but the energy values are
not given by the simple expression in Equation 3.28 for an infinite PE well.
• For the infinite well, the electron wavefunction ψ(x) abruptly terminates at x = 0
and x = a as in Figure 3.16; ψ(x) = 0 outside the well.
IUB Dr. Mustafa H. Chowdhury 77
Confined Electron in a Finite Potential Energy Well
EEE540_L2
• This may seem contrary to the boundary condition that dψ∕dx should be continuous (see
Figure 3.15). However, the infinite PE well is an exceptional case because V = ∞ means
that only ψ = 0 outside the well can satisfy the Schrödinger equation.
IUB Dr. Mustafa H. Chowdhury 78
Confined Electron in a Finite Potential Energy Well
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 79
Confined Electron in a Finite Potential Energy Well
EEE540_L2
• We are looking for electron energies inside the well, that is, E < Vo, which
means α is positive. Each of Equations 3.35 and 3.38a, and 3.38b has two
constants that we need to find through boundary conditions and
requirements on the wavefunction.
• In the present case, ψ(x) cannot be zero at the boundaries, ψ(x) exists
both inside and outside the well, and it must be continuous, single valued
and have a continuous slope, that is dψ∕dx must be continuous. (See
Figure 3.15.)
• Figure 3.17b and c show the wavefunctions and the energies of the
electron derived by continuing the mathematical steps above further.
• Within the well, we have harmonic-type solutions somewhat similar to
before but ψ is not zero at the boundaries.
IUB Dr. Mustafa H. Chowdhury 80
Confined Electron in a Finite Potential Energy Well
EEE540_L2
• The potential energy V(x) is symmetric about x = a∕2, which means that the
wavefunctions must be either even or odd parity as in Figure 3.17. Outside the
well, ψ decreases exponentially as we move away from the well.
• The waveforms in I, II, and III need to be joined smoothly and provide the overall
wavefunction. The energy E of the electron is quantized because only certain
energies give the right k and α for the wavefunctions in Equations 3.35 and 3.38a
and b to satisfy the Schrödinger Equations 3.34 and 3.37.
• In addition, not all solutions exist inasmuch as if we were to impart sufficient
energy to the electron such that E > Vo, the electron would become free. The
number of solutions and the energy values depend on the width a and depth of
the well, Vo.
• The example in Figure 3.17a has only three solutions with the three wavefunctions
ψ1, ψ2, and ψ3 shown in Figure 3.17b. Notice that the wavefunctions penetrate
into the barriers as exponentially decaying functions.
IUB Dr. Mustafa H. Chowdhury 81
Confined Electron in a Finite Potential Energy Well
EEE540_L2
• For example, in region III, the wavefunction ψIII ∝ exp[−α(x−a)]. The quantity 1∕α
is a measure of the extent of penetration of the electron into the barrier, and is
called the penetration depth.
• As a simple example, consider a finite well that has a width of 2 nm and a PE
barrier Vo of 0.50 eV as shown in Figure 3.17c.
• If this were an infinite PE well, the first three levels would be 0.094 eV, 0.38 eV
and 0.85 eV.
• For this finite PE well, only three solutions exist that correspond to E1 = 0.057 eV,
E2 = 0.22 eV, and E3 = 0.45 eV. Notice that the energies are significantly different
and lower (Why?).
• Finite PE wells play an important role in confining charge carriers in today’s
optoelectronic devices.
• One particular optoelectronic application is Terahertz emitters. Electrons are
injected into the well and they move from one level to the next, for example from
E3 to E2. By choosing the width a and the height Vo, the emitted radiation from E3
to E2 or E2 to E1 can be made to be in the terahertz range.
IUB Dr. Mustafa H. Chowdhury 82
Confined Electron in a Finite Potential Energy Well
EEE540_L2
Planck's constant h = 6.62607015 × 10−34 J⋅s OR 4.135667696 × 10−15 eV⋅s
Reduced Planck's constant ħ = 1.054571817 × 10−34 J⋅s OR 6.582119569 × 10−16 eV⋅s
IUB Dr. Mustafa H. Chowdhury 83
Confined Electron in a Finite Potential Energy Well
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 84
Tunneling Phenomenon: Quantum Leak
EEE540_L2
• To understand the tunneling phenomenon, let us examine the thrilling events
experienced by the roller coaster shown in Figure 3.19a.
IUB Dr. Mustafa H. Chowdhury 85
Tunneling Phenomenon: Quantum Leak
EEE540_L2
• Consider what the roller coaster can do when released from rest at a height A. The
conservation of energy means that the carriage can reach B and at most C, but
certainly not beyond C and definitely not D and E.
• Classically, there is no possible way the carriage will reach E at the other side of
the potential barrier D. An extra energy corresponding to the height difference, D
− A, is needed. Anyone standing at E will be quite safe.
• Ignoring frictional losses, the roller coaster will go back and forth between A and
C.
IUB Dr. Mustafa H. Chowdhury 86
Tunneling Phenomenon: Quantum Leak
EEE540_L2
• Now, consider an analogous event on an atomic scale.
• An electron moves with an energy E in a region x < 0 where the potential energy
PE is zero; therefore, E is solely kinetic energy.
• The electron then encounters a potential barrier of “height” Vo, which is greater
than E at x = 0. The extent (width) of the potential barrier is a.
• On the other side of the potential barrier, x > a, the PE is again zero as shown in
Figure 3.19b. What will the electron do?
IUB Dr. Mustafa H. Chowdhury 87
Tunneling Phenomenon: Quantum Leak
EEE540_L2
• Classically, just like the roller coaster, the electron should bounce back and thus be
confined to the region x < 0, because its total energy E is less than Vo.
• In the quantum world, however, there is a distinct possibility that the electron
will “tunnel” through the potential barrier and appear on the other side; it will
leak through.
• To show this, we need to solve the Schrödinger equation for the present choice of
V(x).
• Remember that the only way the Schrödinger equation will have the solution
ψ(x) = 0 is if the PE is infinite, that is, V = ∞.
• Therefore, within any zero or finite PE region, there will always be a solution
ψ(x) and there always will be some probability of finding the electron.
• We can divide the electron’s space into three regions, I, II, and III, as indicated in
Figure 3.19b.
• We can then solve the Schrödinger equation for each region, to obtain three
wavefunctions: ψI(x), ψII(x), and ψIII(x).
IUB Dr. Mustafa H. Chowdhury 88
Tunneling Phenomenon: Quantum Leak
EEE540_L2
• In regions I and III, ψ(x) must be traveling waves, as there is no PE (the electron is
free and moving with a kinetic energy E).
• In zone II, however, E − Vo is negative, so the general solution of the
Schrödinger equation is the sum of an exponentially decaying function and an
exponentially increasing function.
IUB Dr. Mustafa H. Chowdhury 89
Tunneling Phenomenon: Quantum Leak
EEE540_L2
• Equation 3.41 follows from substituting ψI(x) and ψIII(x) into the Schrödinger
equation in regions I and III, respectively. Equation 3.42 for α2 follows from
substituting ψII(x) into the Schrödinger equation in region II. Both k2 and α2, and
hence k and α, in Equations 3.40a to c are positive numbers.
• This means that exp(jkx) and exp(−jkx) represent traveling waves in opposite
directions, and exp(−αx) and exp(αx) represent an exponential decay and rise,
respectively.
• We see that in region I, ψI(x) consists of the incident wave A1exp( jkx) in the +x
direction, and a reflected wave A2exp(−jkx), in the −x direction. Furthermore,
because the electron is traveling toward the right in region III, there is no reflected
wave, so C2 = 0.
• We must now apply the boundary conditions and the normalization condition to
determine the various constants A1, A2, B1, B2, and C1. In other words, we must
match the three waveforms in Equations 3.40a to c at their boundaries (x = 0 and x
= a) so that they form a continuous single-valued wavefunction.
IUB Dr. Mustafa H. Chowdhury 90
Tunneling Phenomenon: Quantum Leak
EEE540_L2
• With the boundary conditions enforced onto the wavefunctions ψI(x), ψII(x), and
ψIII(x), all the constants can be determined in terms of the amplitude A1 of the
incoming wave.
• The relative probability that the electron will tunnel from region I through II to
III is defined as the transmission coefficient T, and this depends very strongly on
both the relative PE barrier height (Vo − E) and the width a of the barrier.
• The final result that comes out from a tedious application of the boundary
conditions is:
IUB Dr. Mustafa H. Chowdhury 91
Tunneling Phenomenon: Quantum Leak
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 92
Tunneling Phenomenon: Quantum Leak
EEE540_L2
• We can now summarize the entire tunneling affair as follows.
• When an electron encounters a potential energy barrier of height Vo
greater than its energy E, there is a finite probability that it will leak
through that barrier.
• This probability depends sensitively on the energy and width of the
barrier.
• For a wide potential barrier, the probability of tunneling is
proportional to exp(−2αa), as in Equation 3.45.
• The wider or higher the potential barrier, the smaller the chance of
the electron tunneling.
IUB Dr. Mustafa H. Chowdhury 93
Tunneling Phenomenon: Quantum Leak
EEE540_L2
• One of the most remarkable technological uses of the tunneling effect is in the
scanning tunneling microscope (STM), which elegantly maps out the surfaces of
solids.
IUB Dr. Mustafa H. Chowdhury 94
Tunneling Phenomenon: Quantum Leak
EEE540_L2
• A conducting probe is brought so close to the surface of a solid that electrons can
tunnel from the surface of the solid to the probe, as illustrated in Figure 3.20.
• When the probe is far removed, the wavefunction of an electron decays
exponentially outside the material, by virtue of the potential energy barrier being
finite (the work function is ∼10 eV).
• When the probe is brought very close to the surface, the wavefunction penetrates
into the probe and, as a result, the electron can tunnel from the material into the
probe.
• Without an applied voltage, there will be as many electrons tunneling from the
material to the probe as there are going in the opposite direction from the probe
to the material, so the net current will be zero.
• On the other hand, if a positive bias is applied to the probe with respect to the
material, as shown in Figure 3.20, an electron tunneling from the material to the
probe will see a lower potential barrier than one tunneling from the probe to the
material.
IUB Dr. Mustafa H. Chowdhury 95
Tunneling Phenomenon: Quantum Leak
EEE540_L2
• Consequently, there will be a net current from the probe to the material and this current will
depend very sensitively on the separation a of the probe from the surface, by virtue of
Equation 3.45.
• Because the tunneling current is extremely sensitive to the width of the potential barrier, the
tunneling current is essentially dominated by electrons tunneling to the probe atom nearest
to the surface.
• Thus, the probe tip has an atomic dimension. By scanning the surface of the material
with the probe and recording the tunneling current the user can map out the surface
topology of the material with a resolution comparable to the atomic dimension.
• The probe motion along the surface, and also perpendicular to the surface, is controlled by
piezoelectric transducers to provide sufficiently small and smooth displacements.
IUB Dr. Mustafa H. Chowdhury 96
Tunneling Phenomenon: Quantum Leak
EEE540_L2
• Figure 3.21 shows an STM image of a graphite surface, on which the hexagonal
carbon rings can be clearly seen.
• Notice that the scale is 0.2 nm (2 Å). The contours in the image actually represent
electron concentrations within the surface since it is the electrons that tunnel from
the graphite surface to the probe tip.
IUB Dr. Mustafa H. Chowdhury 97
Tunneling Phenomenon: Quantum Leak
EEE540_L2
• The astute reader will notice that not all the carbon atoms in a hexagonal ring are
at the same height; three are higher and three are lower.
• The reason is that the exact electron concentration on the surface is also influenced
by the second layer of atoms underneath the top layer.
• The overall effect makes the electron concentration change (alternate) from one
atomic site to a neighboring site within the hexagonal rings.
• STM was invented by Gerd Binning and Heinrich Rohrer at the IBM Research
Laboratory in Zurich, for which they were awarded the 1986 Nobel prize.
IUB Dr. Mustafa H. Chowdhury 98
Tunneling Phenomenon: Quantum Leak
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 99
Tunneling Phenomenon: Quantum Leak
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 100
Tunneling Phenomenon: Quantum Leak
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 101
Tunneling Phenomenon: Quantum Leak
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 102
Tunneling Phenomenon: Quantum Leak
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 103
Tunneling Phenomenon: Quantum Leak
EEE540_L2
IUB Dr. Mustafa H. Chowdhury 104