Chapter 5
Quantum Mechanics
5 Quantum Mechanics
16 April 2018 2 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
5 Quantum Mechanics
16 April 2018 3 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
The Bohr theory of the atom has a number of severe limitations.
It applies only to hydrogen and one-electron ions such as He
+
and Li
+2
.
It cannot explain why certain spectral lines are more intense than others (that is,
why certain transitions between energy levels have greater probabilities of
occurrence than others).
It cannot account for the observation that many spectral lines actually consist of
several separate lines whose wavelengths differ slightly.
Perhaps most important, it does not permit us to obtain an understanding of how
individual atoms interact with one another to endow macroscopic aggregates of
matter with the physical and chemical properties we observe.
A more general approach to atomic phenomena is required.
Such an approach was developed in 1925 and 1926 by Erwin
Schrödinger, Werner Heisenberg, Max Born, Paul Dirac, and others
under the name of quantum mechanics.
5. 1 Quantum Mechanics
16 April 2018 4 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Classical mechanics is an approximation of quantum mechanics
The fundamental difference between classical (or Newtonian)
mechanics and quantum mechanics lies in what they describe.
In classical mechanics, the future history of a particle is completely
determined by its initial position and momentum together with the
forces that act upon it.
Quantum mechanics also arrives at relationships between observable
quantities, but the uncertainty principle suggests that the nature of an
observable quantity is different in the atomic level.
5. 1 Quantum Mechanics: Wave Function
16 April 2018 5 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
The quantity with which quantum mechanics is concerned is the wave
function Ψ of a body.
The linear momentum, angular momentum, and energy of the body are
quantities that can be established from Ψ.
The problem of quantum mechanics is to determine Ψ for a body when
its freedom of motion is limited by the action of external forces.
Wave functions are usually complex with both real and imaginary
parts
(Wave function)
where A and B are real functions.
A probability, however, must be a positive real quantity. The
probability density |Ψ|
2
for a complex is therefore taken as the product
of Ψ and its complex conjugate Ψ* which is Ψ*Ψ.
The complex conjugate of any function is obtained by replacing i by -i
wherever it appears in the function.
Since i
2
=-1; |Ψ|
2
=Ψ*Ψ is always a positive real quantity.
(Complex conjugate)
5. 1 Quantum Mechanics: Normalization
16 April 2018 6 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Since |Ψ|
2
is proportional to the probability density P of finding the
body described by Ψ, the integral of |Ψ|
2
over all space must be finite -
the body is somewhere.
If the particle does not exist.
It is usually convenient to have |Ψ|
2
be equal to the probability density
P of finding the particle described by Ψ, rather than merely be
proportional to P.
If |Ψ|
2
is to equal P, then it must be true that
A wave function that obeys Eq. (5.1) is said to be normalized.
Every acceptable wave function can be normalized by multiplying it by
an appropriate constant.
(5.1 Normaliation)
5.1 Quantum Mechanics: Well-Behaved Wave Functions
16 April 2018 7 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Only wave functions with the properties below can yield physically
meaningful results when used in calculations, so only such “well-
behaved” wave functions are admissible as mathematical
representations of real bodies.
To summarize:
1. Ψ must be continuous and single-valued everywhere.
2. Ψ/x, Ψ/y, Ψ/z must be continuous and single-valued
everywhere.
3. Ψ must be normalizable, which means that Ψ must go to 0 as
x , y , z  in order that Ψ|
2
dV over all space be a
finite constant.
(Probabilty)
For a particle restricted to motion in the x direction, the probability
of finding it between x
1
and x
2
is given by
5.2 The Wave Equation
16 April 2018 8 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
It can have a variety of solutions, including complex ones.
Schrödingers equation, which is the fundamental equation of quantum
mechanics in the same sense that the second law of motion is the
fundamental equation of Newtonian mechanics, is a wave equation in
the variable Ψ.
(5.3 Wave equation)
Solutions of the wave equation may be of many kinds, reflecting the
variety of waves that can occur.
Figure 5.1 Waves in the xy plane traveling in
the x direction along a stretched string lying
on the x axis.
All solutions must be of the form
where F is any function that can be
differentiated.
The solutions F(t-x/v) represent waves
traveling in the +x-direction,
and the solutions F(t+x/v) represent waves
traveling in the -x direction.
5.2 The Wave Equation
16 April 2018 9 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Let us consider the wave equivalent of a “free particle”, which is a
particle that is not under the influence of any forces and therefore
follow a straight path at constant speed.
This wave is described by the general solution of Eq. (5.3) for
undamped (that is, constant amplitude A), monochromatic (constant
angular frequency) harmonic waves in the x direction, namely
In this formula y is a complex quantity, with both real and imaginary
parts.
5.3 Schrödingers Equation: Time-Dependent Form
16 April 2018 10 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
A basic physical principle that cannot be derived from anything else.
In quantum mechanics, the wave function Ψ corresponds to the wave
variable y of wave motion in general.
However, Ψ, unlike y, is not itself a measurable quantity and may
therefore be complex. For this reason, we assume that for a particle
moving freely in the +x-direction is specified by
Replacing in the above formula by 2 and v by  gives
This is convenient since we already know what and are in terms of
the total energy E and momentum p of the particle being described by
Ψ. Because we have
(5.9 Free
particle)
Equation (5.9) describes the wave equivalent of an unrestricted particle
of total energy E and momentum p moving in the +x- direction.
The expression for the wave function Ψ given by Eq. (5.9) is correct
only for freely moving particles.
5.3 Schrödingers Equation: Time-Dependent Form
16 April 2018 11 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
However, we are most interested in situations where the motion of a
particle is subject to various restrictions.
An important concern, for example, is an electron bound to an atom by
the electric field of its nucleus.
What we must now do is obtain the fundamental differential equation
for Ψ, which we can then solve for in a specific situation.
This equation is Schrödingers equation.
We begin by differentiating Eq. (5.9) for Ψ twice with respect to x,
which gives
differentiating Eq. (5.9) once with respect to t gives
(5.10)
(5.11)
5.3 Schrödingers Equation: Time-Dependent Form
16 April 2018 12 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
At speeds small compared with that of light, the total energy E of a
particle is the sum of its kinetic energy p
2/
2m and its potential energy
U, where U is in general a function of position x and time
(5.12)
(5.14 Time dependent
Schrödinger’s equation in 1D)
The function U represents the influence of the rest of the universe on
the particle.
Of course, only a small part of the universe interacts with the particle
to any extent;
for instance, in the case of the electron in a hydrogen atom, only the
electric field of the nucleus must be taken into account.
Multiplying both sides of Eq. (5.12) by the wave function Ψ.
Now we substitute for E Ψ and p
2
Ψ from Eqs. (5.10) and (5.11) to
obtain the time dependent form of Schrödingers equation:
5.3 Schrödingers Equation: Time-Dependent Form
16 April 2018 13 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
where the particle’s potential energy U is some function of x, y,
z, and t.
Any restrictions that may be present on the particle’s motion
will affect the potential energy function U.
In three dimensions the time-dependent form of
Schrödingers equation is
Ernwin
Schrödinger
(18871961)
Nobel Prize in
Physics in 1933
Once U is known, Schrödingers equation may be solved for the wave
function Ψ of the particle, from which its probability density |Ψ|
2
may
be determined for a specified x, y, z, t.
Schrödingers equation cannot be derived from other basic principles of
physics; it is a basic principle in itself.
(5.15 Time dependent
Schrödinger’s equation in 3D)
5.4 Linearity and Superposition
16 April 2018 14 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Wave functions add, not probabilities.
An important property of Schrödingers equation is that it is linear in
the wave function: the equation has terms that contain and its
derivatives but no terms independent of or that involve higher powers
of or its derivatives.
As a result, a linear combination of solutions of Schrödingers equation
for a given system is also itself a solution.
If Ψ
1
and Ψ
2
are two solutions (that is, two wave functions that satisfy the
equation), then Ψ= a
1
Ψ
1
+a
2
Ψ
2
is also a solution, where a
1
and a
2
are constants.
Superposition principle.
We conclude that interference effects can occur for wave functions just
as they can for light, sound, water, and electromagnetic waves.
Let us apply the superposition principle to the diffraction of an electron
beam.
5.4 Linearity and Superposition
16 April 2018 15 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Figure 5.2a shows a pair of slits through which a parallel beam of
monoenergetic electrons pass on their way to a viewing screen.
Figure 5.2 (a) Arrangement of double-slit experiment. (b) The
electron intensity at the screen with only slit 1 open. (c) The electron
intensity at the screen with only slit 2 open. (d) The sum of the
intensities of (b) and (c). (e) The actual intensity at the screen with
slits 1 and 2 both open. The wave functions Ψ
1
and Ψ
2
add to
produce the intensity at the screen, not the probability densities |Ψ1|
2
and |Ψ2|
2
If slit 1 only is open, the result is the intensity variation shown in Fig. 5.2b that
corresponds to the probability density P=|Ψ
1
|
2
= Ψ
1
*
Ψ
1
If slit 2 only is open, as in Fig. 5.2c, the corresponding probability density is
P=|Ψ
2
|
2
= Ψ
2
*
Ψ
2
We might suppose that opening both slits would give an electron intensity variation
described by P
1
+ P
2
, as in Fig. 5.2d
However, this is not the case because in quantum mechanics wave functions add,
not probabilities.
Instead the result with both slits open is as shown in Fig. 5.2e, the same pattern of
alternating maxima and minima that occurs when a beam of monochromatic light
passes through the double slit of Fig. 2.4. Superposition of the wave functions.
5.5 Expectation Values
16 April 2018 16 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
How to extract information from a wave function.
Once Schrödingers equation has been solved for a particle in a given
physical situation, the resulting wave function Ψ(x, y, z, t) contains all
the information about the particle (that is permitted by the uncertainty
principle).
Let us calculate the expectation value <x> of the position of a particle
confined to the x axis that is described by the wave function Ψ(x, t).
This is the value of x we would obtain if we measured the positions
of a great many particles described by the same wave function at
some instant t and then averaged the results.
What is the average position x of a number of identical particles
distributed along the x axis in such a way that there are N
1
particles at
x
1
, N
2
particles at x
2
, and so on? The average position in this case is the
same as the center of mass of the distribution, and so
5.5 Expectation Values
16 April 2018 17 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
When we are dealing with a single particle, we must replace the
number N
i
of particles at x
i
by the probability P
i
that the particle be
found in an interval dx at x
i
.
This probability is P
i
=| Ψ
i
|
2
dx where Ψ
i
is the particle wave function
evaluated at x=x
i
. Making this substitution and changing the
summations to integrals, we see that the expectation value of the
position of the single particle is
If Ψ is a normalized wave function, the denominator of Eq. (5.18)
equals the probability that the particle exists somewhere between
x=- and x= therefore has the value 1. In this case
(5.19 Expectation value for position)
The same procedure as that followed above can be used to obtain the
expectation value G(x) of any quantity-for instance, potential energy
U(x)-that is a function of the position x of a particle described by a
wave function Ψ. The result is
(5.20 Expectation value for G)
5.5 Expectation Values
16 April 2018 18 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Example 5.2
A particle limited to the x axis has the wave function Ψ=ax between x=0
and x=1; Ψ= 0 elsewhere. (a) Find the probability that the particle can be
found between x=0.45 and x=0.55. (b) Find the expectation value <x> of
the particle’s position.
5.6 Operators
16 April 2018 19 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Another way to find expectation values.
A hint as to the proper way to evaluate <p> and <E> comes from
differentiating the free particle wave function Ψ=A e
(-i/h(Et-px))
with
respect to x and to t. We find that
(5.21)
(5.22)
An operator tells us what operation to carry out on the quantity that
follows it.
Thus, the operator E instructs us to take the partial derivative of what
comes after it with respect to t and multiply the result by .
It is customary to denote operators by using a caret, so that ˆp is the
operator that corresponds to momentum p and ˆE is the operator that
corresponds to total energy E.
From Eqs. (5.21) and (5.22) these operators are
(5.23 Momentum operator)
(5.24 Total-energy operator)
They are entirely
general results whose
validity is the same as
that of Schrödingers
equation.
5.6 Operators
16 April 2018 20 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Replace the equation E=KE+U for the total energy of a particle with
the operator equation we have
(5.25)
which is Schrödingers equation. Postulating Eqs. (5.23) and (5.24) is
equivalent to postulating Schrödingers equation.
5.6 Operators: Operators and Expectation Values
16 April 2018 21 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
(5.30 Expectation value of an operator)
Because p and E can be replaced by their corresponding operators in
an equation, we can use these operators to obtain expectation values
for p and E. Thus the expectation values for p and E are
Every observable quantity G characteristic of a physical system may
be represented by a suitable quantum-mechanical operator ˆG. To
obtain this operator, we express G in terms of x and p. If the wave
function Ψ of the system is known, the expectation value of G(x, p) is
5.7 Schrödingers Equation: Steady-State Form
16 April 2018 22 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
(5.31)
Eigenvalues and eigenfunctions.
In a great many situations, the potential energy of a particle does not
depend on time explicitly;
the forces that act on it, and hence U, vary with the position of the
particle only.
Then Schrödingers equation may be simplified by removing all
reference to t.
Ψ is now the product of a time-dependent function exp(-(iE/h)t) and a
position dependent function ψ.
Substituting the Ψ of Eq. (5.31) into the time-dependent form of
Schrödingers equation and dividing through the common exponential
factor, we find that
(5.32 Steady-state Schrödinger
equation in one dimension)
Equation (5.32) is the steady-state form of Schrödingers equation.
5.7 Schrödingers Equation: Steady-State Form & Eigenvalues and Eigenfunctions
16 April 2018 23 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
In three dimensions, the steady-state form of Schrödingers
equation is
An important property of Schrödingers steady-state equation is that, if
it has one or more solutions for a given system, each of these wave
functions corresponds to a specific value of the energy E.
Thus, energy quantization appears in wave mechanics as a natural
element of the theory.
The values of energy E
n
for which Schrödingers steady-state equation
can be solved are called eigenvalues and the corresponding wave
functions ψ
n
are called eigenfunctions.
The discrete energy levels of the hydrogen atom are an example of a
set of eigenvalues.
5.7 Schrödingers Equation: Operators and Eigenvalues
16 April 2018 24 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
An important example of a dynamical variable other than total energy
that is found to be quantized in stable systems is angular momentum L.
In the case of the hydrogen atom, eigenvalues of the magnitude of the
total angular momentum:
In the hydrogen atom, the electron’s position is not quantized. So
that we must think of the electron as being present in the vicinity of the
nucleus with a certain probability |ψ|
2
per unit volume but with no
predictable position or even orbit in the classical sense.
The condition that a certain dynamical variable G be restricted to the
discrete values G
n
(G be quantized) is that the wave functions ψ
n
of
the system be such that
(5.34 Eigenvalue equation)
where ˆG is the operator that corresponds to G and each G
n
is a real
number.
If measurements of G are made on a number of identical systems
(eigenfunction ψ
k
), each measurement will yield the single value
G
k
.
5.5 Expectation Values
16 April 2018 25 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Example 5.3
An eigenfunction of the operator d
2
/dx
2
is ψ=e
(2x)
. Find the
corresponding eigenvalue.
5.7 Schrödingers Equation: Operators and Eigenvalues
16 April 2018 26 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
The total-energy operator ˆE can also be written as
(5.35 Hamiltonian operator)
and is called the Hamiltonian operator. The steady-state
Schrödinger equation can be written simply as
(5.36 Schrödinger's equation)
Table 5.1 Lists the operators that correspond to various
observable quantities.
5.8 Particle in a Box
16 April 2018 27 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
We may specify the particle’s motion by saying that
it is restricted to traveling along the x axis between
x=0 and x=L by infinitely hard walls.
A particle does not lose energy when it collides with
such walls, so that its total energy stays constant.
Figure 5.4 A square potential well
with infinitely high barriers at each
end corresponds to a box with
infinitely hard walls.
Potential energy U of the particle is infinite on both
sides of the box, while U is a constant -say 0 for convenience- on the
inside (Fig. 5.4).
Because the particle cannot have an infinite amount of energy, it
cannot exist outside the box, and so its wave function ψ is 0 for x≤0
and x≥L.
How boundary conditions and normalization determine wave
functions.
The simplest quantum-mechanical problem is that of a particle trapped
in a box with infinitely hard walls.
5.8 Particle in a Box: Eigenvalues
16 April 2018 28 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Our task is to find what ψ is within the box (btw x=0 and x=L). Within
the box Schrödingers equation becomes
(5.37)
since U=0 there. Equation (5.37)
has the solution
(5.38)
A and B are constants to be
evaluated
This solution is subject to the boundary conditions:
ψ=0 for x=0 and for x=L. Since cos 0=1, the second term cannot
describe the particle because it does not vanish at x=0. Hence, B=0.
Since sin0=0, the sine term always yields ψ=0 at x=0, as required,
but ψ will be 0 at x=L only when
(5.39)
From Eq. (5.39), Energy of the particle can have only certain values
(eigenvalues). These eigenvalues, constituting the energy levels of the
system, are found by solving Eq. (5.39) for E
n
, which gives
(5.40 Particle in a box)
5.8 Particle in a Box: Eigenfunctions
16 April 2018 29 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
The wave functions (eigenfunctions) of a particle in a box whose
energies are E
n
are, from Eq. (5.38) with B=0
(5.41)
substituting Eq. (5.40) for E
n
gives
(5.42)
for the eigenfunctions corresponding to the
energy eigenvalues En.
With the help of the trigonometric identity sin
2
θ=1/2(1-cos2θ) we find
that
(5.43)
5.8 Particle in a Box: Normalization
16 April 2018 30 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
To normalize ψ we must assign a value to A such that |ψ
n
|
2
dx is equal
to the probability Pdx of finding the particle between x and x+dx,
rather than merely proportional to Pdx. If |ψ
n
|
2
dx is to equal P dx, then
it must be true that
(5.44)
Comparing Eqs. (5.43) and (5.44), we see that the wave
functions of a particle in a box are normalized if
(5.45)
The normalized wave functions of the particle are
therefore
(5.46 Particle in a box)
5.8 Particle in a Box
16 April 2018 31 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
The normalized wave functions ψ
1
, ψ
2
, and ψ
3
together
with the probability densities |ψ
1
|
2
,
2
|
2
, and |ψ
3
|
2
are
plotted in Fig. 5.5.
Although ψ
n
may be negative as well as positive,
n
|
2
is never negative.
Since ψ
n
is normalized, its value (
n
|
2
) at a given x is
equal to the probability density of finding the particle
there.
In every case |ψ
n
|
2
=0 at x=0 and x=L, the boundaries of
the box.
Figure 5.5 Wave
functions and
probability densities of
a particle confined to a
box with rigid walls.
5.8 Particle in a Box
16 April 2018 32 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Example 5.4
Find the probability that a particle trapped in a box L wide can be found
between 0.45L and 0.55L for the ground and first excited states.
Figure 5.6 The probability P
x1,x2
of finding a particle in the box of
Fig. 5.5 between x
1
=0.45L and
x
2
=0.55L is equal to the area
under the |ψ|
2
curves between
these limits.
5.8 Particle in a Box
16 April 2018 33 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Example 5.5
Find the expectation value <x> of the position of a particle trapped in a
box L wide.
5.9 Finite Potential Well
16 April 2018 34 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
The wave function penetrates the walls, which lowers the energy levels.
Potential energies are never infinite in the real world, and the box with
infinitely hard walls (particle in a box) has no physical counterpart.
However, potential wells with barriers of finite height certainly do
exist.
Figure 5.7 shows a potential well with square
corners that is U high and L wide and contains a
particle whose energy E is less than U
According to classical mechanics, when the
particle strikes the sides of the well, it bounces
off without entering regions I and III.
Figure 5.7 A square potential well
with finite barriers. The energy E of
the trapped particle is less than the
height U of the barriers.
In quantum mechanics, the particle also bounces back and forth, but
now it has a certain probability of penetrating into regions I and III
even though E<U.
5.9 Finite Potential Well
16 April 2018 35 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
In regions I and III Schrödingers steady-state equation is
which we can rewrite in the more convenient form
(5.53)
The solutions to
Eq. (5.53) are real
exponentials:
Both ψ
I
and ψ
III
must be finite everywhere. (e
-
0)
Since e
-ax
as x - and e
ax
as x , the
coefficients D and F must therefore be 0. Hence we have
These wave functions decrease exponentially inside the barriers at the
sides of the well.
In regions II (within the well). Schrödingers equation is the same as
Eq. (5.37) and its solution is
(5.59)
Here, ψ
II
=C at x=0 and ψ
II
=G at x =L, so both the sine and cosine
solutions of Eq. (5.59) are possible.
5.9 Finite Potential Well
16 April 2018 36 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
For either solution, both ψ and dψ/dx must be
continuous at x=0 and x=L:
The wave functions inside and outside each side of
the well must not only have the same value where
they join.
But also the same slopes, so they match up perfectly.
When these boundary conditions are taken into
account, the result is that exact matching only occurs
for certain specific values E
n
of the particle energy.
The complete wave functions and their probability
densities are shown in Fig. 5.8.
Figure 5.8 Wave functions
and probability densities
of a particle in a finite
potential well. The
particle has a certain
probability of being found
outside the wall.
5.11 Harmonic Oscillator
16 April 2018 37 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Its energy levels are evenly spaced.
Harmonic motion takes place when a system of some kind vibrates
about an equilibrium configuration. The system may be
an object supported by a spring,
floating in a liquid,
a diatomic molecule,
an atom in a crystal lattice … on all scales of size.
The condition for harmonic motion is the presence of a restoring force
that acts to return the system to its equilibrium configuration when it is
disturbed.
(Hooke’s law)
This relationship is customarily called Hooke’s law. From the second
law of motion, F=ma, we have
(5.62 Harmonicc oscillator)
5.11 Harmonic Oscillator
16 April 2018 38 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
(5.64 Frequency of harmonic oscillator)
There are various ways to write the solution to Eq. (5.62). A common
one is
ν is the frequency of the oscillations and A is their amplitude. The value
of , the phase angle, depends upon what x is at the time t=0 and on the
direction of motion then.
The importance of the simple harmonic oscillator in both classical
and modern physics lies not in the strict adherence of actual restoring
forces to Hooke’s law, which is seldom true, but in the fact that these
restoring forces reduce to Hooke’s law for small displacements x.
As a result, any system in which something executes small vibrations
about an equilibrium position behaves very much like a simple
harmonic oscillator.
5.11 Harmonic Oscillator
16 April 2018 39 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
The potential-energy function U(x) that corresponds to a Hooke’s law
force may be found by calculating the work needed to bring a particle
from x=0 to x=x against such a force. (see Figure 5.10)
Figure 5.10 The potential energy of a
harmonic oscillator is proportional to x
2
,
where x is the displacement from the
equilibrium position. The amplitude A of the
motion is determined by the total energy E of
the oscillator, which classically can have any
value.
Three quantum mechanical modifications to
this classical picture:
1.The allowed energies will not form a
continuous spectrum but instead a discrete
spectrum of certain specific values only.
2.The lowest allowed energy will not be E=0
but will be some definite minimum E=E
0
.
3.There will be a certain probability that the
particle can penetrate the potential well it is
in and go beyond the limits of -A and +A.
5.11 Harmonic Oscillator: Energy Levels
16 April 2018 40 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
Schrödingers equation for the harmonic oscillator is, with U=1/2kx
2
,
Figure 5.11 Potential wells and energy levels
of (a) a hydrogen atom, (b) a particle in a box,
and (c) a harmonic oscillator. In each case the
energy levels depend in a different way on the
quantum number n. Only for he harmonic
oscillator are the levels equally spaced.
(5.70 Energy levels of harmonic oscillator)
The energy of a harmonic oscillator is thus quantized in
steps of hν. Note that when n=0,
(5.71 Zero point energy)
This value is called the zero-point energy because a
harmonic oscillator in equilibrium with its surroundings
would approach an energy of E=E
0
and not E=0 as the
temperature approaches 0 K.
Figure 5.11 is a comparison of the energy levels of a
harmonic oscillator with those of a hydrogen atom and of
a particle in a box with infinitely hard walls.
The shapes of the respective potential-
energy curves are also shown.
The spacing of the energy levels is constant
only for the harmonic oscillator.
16 April 2018 41 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
5 Solved Problems
1. Show that Ae
i(kx-wt)
satisfies the time-dependent Schrödinger wave
equation.
16 April 2018 42 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
5 Solved Problems
2. The initial wavefunction of a particle is given as
ψ(x,0)=C e
(-|x|/x0),
where C and x
0
are constants.
Sketch of the function is given.
a) Find C in terms of x
0
such that ψ(x,0) is
normalized.
b) Calculate the probability that the particle will
be found in the interval -x
0
x
x
0
.
16 April 2018 43 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
5 Solved Problems
3. A small object of mass 1.00 mg is confined to move between two
rigid walls separated by 1.00 cm.
a) Calculate the minimum speed of the object.
b) If the speed of the object is 3.00 cm/s, find the corresponding
value of n.
16 April 2018 44 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
5 Solved Problems
4. An electron is trapped in a one-dimensional region of length
1.00x10
-10
m (a typical atomic diameter).
a) Find the energies of the ground state and first two excited states.
b) How much energy must be supplied to excite the electron from
the ground state to the second excited state?
c) From the second excited state, the electron drops down to the
first excited state. How much energy is released in this process?
d) In the first excited state, what is the probability of finding the
electron between x=0 and x=0.025 nm?
16 April 2018 45 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
5 Solved Problems
16 April 2018 46 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
5 Solved Problems
5. An atom can be viewed as a number of electrons
moving around a positively charged nucleus, where
the electrons are subject mainly to the Coulombic
attraction of the nucleus (which actually is partially
“screened” by the intervening electrons). The
potential well that each electron “sees” is sketched
in Figure. Use the model of a particle in a box to estimate the energy (in
eV) required to raise an atomic electron from the state n=1 to the state
n=2, assuming the atom has a radius of 0.100 nm.
16 April 2018 47 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
5 Solved Problems
6. A particle is known to be in the ground state of an infinite square
well with length L. Calculate the probability that this particle will be
found in the middle half of the well, that is, between x=L/4 and
x=3L/4.
16 April 2018 48 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
5 Solved Problems
7. An electron is bound to a region of space by a springlike force with
an effective spring constant of k = 95.7 eV/nm
2
.
a) What is its ground-state energy?
b) How much energy must be absorbed for the electron to jump
from the ground state to the second excited state?
16 April 2018 49 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
5 Quantum Mechanics
Additional Materials
16 April 2018 50 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
5.8 Particle in a Box
5.10 Tunnel Effect
16 April 2018 51 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
A particle without the energy to pass over a
potential barrier may still tunnel through it
Although the walls of the potential well of
Fig. 5.7 were of finite height, they were
assumed to be infinitely thick.
As a result, the particle was trapped
forever even though it could penetrate
the walls.
We next look at the situation of a particle
that strikes a potential barrier of height U,
again with E<U, but here the barrier has a
finite width (Fig. 5.9).
Figure 5.9 When a particle of energy E< U
approaches a potential barrier, according to
classical mechanics the particle must be reflected.
In quantum mechanics, the de Broglie waves that
correspond to the particle are partly reflected and
partly transmitted, which means that the particle
has a finite chance of penetrating the barrier.
What we will find is that the particle has a certain probability -not
necessarily great, but not zero either- of passing through the barrier
and emerging on the other side.
5.10 Tunnel Effect
16 April 2018 52 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
The particle lacks the energy to go over the top of the barrier, but it can
nevertheless tunnel through it, so to speak.
Not surprisingly, the higher the barrier and the wider it is, the less the
chance that the particle can get through.
The tunnel effect actually occurs, notably in the case of the alpha
particles emitted by certain radioactive nuclei.
An alpha particle whose kinetic energy is only a few MeV is able
to escape from a nucleus whose potential wall is perhaps 25 MeV
high.
The probability of escape is so small that the alpha particle might
have to strike the wall 1038 or more times before it emerges, but
sooner or later it does get out.
Tunneling also occurs in the operation of certain semiconductor diodes
in which electrons pass through potential barriers even though their
kinetic energies are smaller than the barrier heights.
5.10 Tunnel Effect: Scanning Tunneling Microscope
16 April 2018 53 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
The ability of electrons to tunnel through a potential
barner is used in an ingenious way in the scanning
tunneling microscope (STM) to study surfaces on an
atomic scale of size.
In an STM, a metal probe with a point so fine that its tip is a
single atom is brought close to the surface of a conducting or
semiconducting material.
Normally even the most loosely bound electrons in an atom on a
surface need several electron-volts of energy to escape -this is the
work function.
However, when a voltage of only 10 mV or so is applied between
the probe and the surface, electrons can tunnel across the gap
between them if the gap is small enough, a nanometer or two.
What is done is to move the probe across the surface in a series of
closely spaced back-and-forth scans.
5.10 Tunnel Effect: Scanning Tunneling Microscope
16 April 2018 54 MSE 228 Engineering Quantum Mechanics © Dr.Cem Özdoğan
The height of the probe is continually adjusted to give a constant
tunneling current, and the adjustments are recorded so that a map
of surface height versus position is built up.
Such a map is able to resolve individual atoms on a surface.
Actually, the result of an STM scan is not a true topographical
map of surface height but a contour map of constant electron
density on the surface. This means that atoms of different
elements appear differently.
Although many biological materials conduct electricity, they do so by the flow of
ions rather than of electrons and so cannot be studied with STMs.
The atomic force microscope (AFM) can be used on any surface, although with
somewhat less resolution than an STM.
In an AFM, the sharp tip of a fractured diamond presses gently against the
atoms on a surface.
A spring keeps the pressure of the tip constant, and a record is made of the
deflections of the tip as it moves across the surface.
The result is a map showing contours of constant repulsive force between the
electrons of the probe and the electrons of the surface atoms.