Doc-20230723-Wa0005 230830 225908

You might also like

You are on page 1of 12

Paper: Nanoscience and Nanotechnology

Unit-1: Module-3

In this module we will learn the following:


- Importance of particle-in-a-box systems
- Wavefunctions and energies of particle in a one-dimensional system
- Wavefunctions and energies of particle in a two-dimensional system
- Wavefunctions and energies of particle in a three-dimensional system

1. Particle in a Box
Quantum mechanics describes the particle in a box model as a particle which is can freely move within a
small space surrounded by impenetrable barriers. This model is also termed as infinite potential well or an
infinite square well, mainly due to its structure. This model is specifically employed as in theoretical
investigations to demonstrate the differences between the classical and quantum systems. For instance, in
case of classical system, a ball entrapped within a large box is free to move inside the box at any speed,
and has equal probability of being found at any point inside the box. As the well becomes narrow and
reaches the nanoscale dimensions, quantum confinement effects assume importance. As a result, the
particle is no longer to free to move anywhere within the box, but can only occupy specific energy levels.
Additionally, it cannot have zero energy at any moment, implying that the particle is never at rest.
Furthermore, depending upon the energy levels, the probability of finding the particle at certain positions
is more than at other places. Thus, the particles may not be detectable at some specific positions. These
positions which have zero probability of particle occupation are termed as spatial nodes.
In quantum mechanics, very few problems are solvable analytically without using approximations.
Particle in a box model is one such problem. It is a very simple model which provides essential insights
into the quantum effects without going for complex mathematical derivations. It can be used to explain
the appearance of energy quantization (or discrete energy levels) within complex quantum systems, e.g.,
atoms and molecules. This model is very frequently employed as an approximation for more complex
quantum systems.
1.1 Particle in an One Dimensional System
One dimensional (1D) system is the simplest form of the particle in a box model. In this model, the
particle can move only in backward and forward directions along a straight line and is surrounded by
impenetrable barriers at either end. The walls of a 1D box can be visualized as regions of space which
have infinitely large potential energy. On the contrary, a constant zero potential energy is assumed to
present within the box. This implies that the particle is free to move within the box and no forces act upon
it in this region. Further, owing to infinitely large potential energy at the walls, very large forces are
applied on the particles by the walls of the box. These forces repel the particle when it touches the walls
of the box. Thus the particle cannot escape the potential well. Potential energy, in case of 1D box, can be
calculated as:
0, 𝑤ℎ𝑒𝑛 0 < 𝑥 < 𝑎
𝑉(𝑥) = {
∞, 𝑒𝑙𝑠𝑒𝑤ℎ𝑒𝑟𝑒
The Schrodinger equation in the region is:

ħ2 𝑑 2
[− + 𝑉(𝑥)] 𝛹 = 𝐸𝛹
2𝑚 𝑑𝑥 2

with the boundary conditions shown in Figure 1.

Figure 1 Schematics of a one dimensional particle in a box system. The barriers outside the
box have infinitely large potential, whereas the interior of the box has a constant, zero
potential.

The solution to Schrodinger equation in the region of interest (𝑉 = 0) is:

𝑑2 𝛹 2𝑚𝐸
2
=− 2 𝛹
𝑑𝑥 ħ
whereupon letting 𝑘 2 = 2𝑚𝐸/ħ2 yields the following second-order differential equation:

𝑑2 𝛹
+ 𝑘 2𝛹 = 0
𝑑𝑥 2
General solutions to this are linear combinations of plane waves

𝛹(𝑥) = 𝐴𝑒 𝑖𝑘𝑥 + 𝐵𝑒 −𝑖𝑘𝑥


where 𝐴 and 𝐵 are constant coefficients to be determined.
Let us now apply the boundary conditions to specify 𝐴 and 𝐵. Namely, the first says that the
wavefunction must gor to zero at the origin (i.e., 𝛹(0) = 0). Therefore, we get:

𝛹(0) = 𝐴 + 𝐵 = 0
Or 𝐵 = −𝐴. This gives the complex wavefunction:
𝛹(𝑥) = 𝐴(𝑒 𝑖𝑘𝑥 − 𝑒 −𝑖𝑘𝑥 )

which can be rewritten using Euler’s relation:

𝛹(𝑥) = 2𝑖𝐴 sin 𝑘𝑥


Next, we apply the second boundary condition, i.e., 𝛹(𝑎) = 0 to get:

𝛹(𝑎) = 2𝑖𝐴 sin 𝑘𝑎 = 0


At this point, to obtain a solution, we have either 𝐴 = 0 (the trivial solution) or 𝑘𝑎 = 𝑛𝜋, with 𝑛 an
integer. We pursue the second solutions, since 𝐴 = 0 just means that the wavefunction is zero.
Consequently, 𝑘 values are restricted to:

𝑘 = 𝑛𝜋/𝑎
It is interesting to note here a factor-of-two difference between the 𝑘 value in this case and that in the free
particle case.
Therefore,
𝑛𝜋𝑥
𝛹(𝑥) = 2𝑖𝐴 sin
𝑎
or, more generally,
𝑛𝜋𝑥
𝛹(𝑥) = 𝑁 sin
𝑎
Where 𝑁 is a normalization constant to be determined. This is the general form of wavefunction for a
particle in a 1D box.
Normalization
The probabilistic interpretation of wavefunctions means that they must be normalized (i.e., the particle
must be located somewhere within the box with unity probability):
𝑎 𝑎
𝑛𝜋𝑥
1 = ∫ 𝛹(𝑥)∗ 𝛹(𝑥)𝑑𝑥 = ∫ 𝑁 2 𝑠𝑖𝑛2 𝑑𝑥
0 0 𝑎

We therefore solve the above integral to find 𝑁 by invoking the trigonometric identity 𝑠𝑖𝑛2 𝑚 =
1
(1 − cos 2𝑚). This gives:
2

𝑁2 𝑎
𝑎
2𝑛𝜋𝑥 𝑁2
1= (𝑥|0 − ∫ cos 𝑑𝑥 ) = 𝑎
2 0 𝑎 2

Solving for 𝑁 gives:

2
𝑁=√
𝑎

Wavefunctions
Normalized wavefunctions for a particle in a 1D box are therefore:
2 𝑛𝜋𝑥
𝛹(𝑥) = √ sin
𝑎 𝑎

Figure 2 shows the first three wavefunctions along with their accompanying probability densities.

Figure 2 The first three wavefunctions for a particle in a 1D box (a) and their associated probability
densities (b).

Energies
Associated energies arise from the equivalence obtained earlier:
𝑛𝜋
𝑘=
𝑎

2𝑚𝐸 𝑛𝜋
√ 2 =
ħ 𝑎

yielding
𝑛 2 ℎ2
𝐸=
8𝑚𝑎2
where 𝑛 = 1,2,3, … Note that the quantum number 𝑛 is never zero, as this would violate the Heisenberg
Uncertainty Principle. Finally, we can extend these results to two and three dimensions by simply varying
the form of the Laplacian and assuming separable solutions.
Figure 3 shows the first five energy levels of the particle in a box along with their corresponding
quantum numbers. Note the absence of degeneracies (i.e., multiple levels having the same energy) in the
one-dimensional box energy levels. Apparently, the states are more closely spaced together at lower
energies but become more distant as 𝐸 increases.
Figure 3 Energy levels for a particle in a one-dimensional box.

1.2 Particle in a Two-Dimensional Box: Degeneracies


For simplicity, we shall solve the problem of a particle confined to a two-dimensional box in (𝑥, 𝑦) plane
of a Cartesian coordinate system. The wavefunctions and energies have 𝑥 and 𝑦 dependencies.
Additionally, an important consequence of working in higher dimensions is increased likelihood of
finding degeneracies. Implying that there will be increased propensity for finding states that have the
same energy. This stems from built-in symmetries of the system.
The Schrodinger equation in this case is:

ħ2 2
[− ∇ + 𝑉(𝑥, 𝑦)] 𝛹(𝑥, 𝑦) = 𝐸𝛹(𝑥, 𝑦)
2𝑚

where, in two dimensions, the Laplacian takes the form:

𝜕2 𝜕2
∇2 = +
𝜕𝑥 2 𝜕𝑦 2
As a consequence, we have:

ħ2 𝜕 2 𝜕2
[− ( 2 + 2 ) + 𝑉(𝑥, 𝑦)] 𝛹(𝑥, 𝑦) = 𝐸𝛹(𝑥, 𝑦)
2𝑚 𝜕𝑥 𝜕𝑦

We assume the potentials and the boundary conditions as shown in Figure 4. Namely, we have a
symmetric box with a dimension 𝑎 along each axis and with a potential that is zero inside the box but
infinite outside it. The particle is therefore confined to the interior of the box.
Figure 4 Potential for a particle in a two-dimensional box.
Additionally, its wavefunctions go to zero at the edges. We therefore have for the potential:
0 𝑊ℎ𝑒𝑛 0 < 𝑥 < 𝑎 𝑎𝑛𝑑 0 < 𝑦 < 𝑎
𝑉(𝑥, 𝑦) = {
∞ 𝑒𝑙𝑠𝑒𝑤ℎ𝑒𝑟𝑒
with the accompanying boundary conditions:
𝛹(0, 𝑦) = 0 for 0 ≤ 𝑦 ≤ 𝑎,
𝛹(𝑎, 𝑦) = 0 for 0 ≤ 𝑦 ≤ 𝑎,
𝛹(𝑥, 0) = 0 for 0 ≤ 𝑥 ≤ 𝑎,
𝛹(𝑥, 𝑎) = 0 for 0 ≤ 𝑥 ≤ 𝑎.
We now solve the Schrodinger equation in the region of interest, where 𝑉(𝑥, 𝑦) = 0:

ħ2 𝜕 2 𝜕2
− ( 2 + 2 ) 𝛹(𝑥, 𝑦) = 𝐸𝛹(𝑥, 𝑦)
2𝑚 𝜕𝑥 𝜕𝑦

For simplicity, let us assume a separable solution of the form:


𝛹(𝑥, 𝑦) = 𝛹𝑥 (𝑥)𝛹𝑦 (𝑦) ≡ 𝛹𝑥 𝛹𝑦

which is basically a product of two independent wavefunctions, each with its own 𝑥 or 𝑦 dependence.
There is a benefit in choosing this type of solution, because the result ultimately comprises two separate
one-dimensional particle-in-a-box problems. Since we have already solved this model, the two-
dimensional problem simplifies to a great extent.
Introducing 𝛹 = 𝛹𝑥 𝛹𝑦 into the Schrodinger equation gives:
ħ2 𝑑2 𝛹𝑥 ħ2 𝑑2 𝛹𝑦
𝛹𝑦 (− ) + 𝛹𝑥 (− ) = 𝐸𝛹𝑥 𝛹𝑦
2𝑚 𝑑𝑥 2 2𝑚 𝑑𝑦 2

whereupon dividing both sides by 𝛹𝑥 𝛹𝑦 yields

1 ħ2 𝑑2 𝛹𝑥 1 ħ2 𝑑2 𝛹𝑦
(− ) + (− ) =𝐸
𝛹𝑥 2𝑚 𝑑𝑥 2 𝛹𝑦 2𝑚 𝑑𝑦 2

This relation remains valid for all values of 𝑥 and 𝑦. Consequently, each term on the left-hand side must
equal a constant; these constants turn out to be the energy contributions from 𝑥 and 𝑦. We thus separate
the total energy term into contributions, one from 𝑥 and one from 𝑦:
𝐸 = 𝐸𝑥 + 𝐸𝑦

This in turn means that we can separate the full equation into two smaller ones:

1 ħ2 𝑑2 𝛹𝑥
(− ) = 𝐸𝑥
𝛹𝑥 2𝑚 𝑑𝑥 2

1 ħ2 𝑑2 𝛹𝑦
(− ) = 𝐸𝑦
𝛹𝑦 2𝑚 𝑑𝑦 2

Rearranging both terms yields:


ħ2 𝑑2 𝛹𝑥
− = 𝐸𝑥 𝛹𝑥
2𝑚 𝑑𝑥 2
ħ2 𝑑2 𝛹𝑦
− = 𝐸𝑦 𝛹𝑦
2𝑚 𝑑𝑦 2
Which are two one-dimensional particle-in-a-box problems. This demonstrates how assuming a separable
solution greatly simplifies the original two-dimensional problem. Now both these one-dimensional
problems can be solved as shown above.
The obtained independently normalized wavefunctions are:

2 𝑛𝑥 𝜋𝑥
𝛹𝑥 = √ sin
𝑎 𝑎

2 𝑛𝑦 𝜋𝑦
𝛹𝑦 = √ sin
𝑎 𝑎

with the corresponding energies:


𝑛𝑥2 ℎ2
𝐸𝑥 =
8𝑚𝑎2
𝑛𝑦2 ℎ2
𝐸𝑦 =
8𝑚𝑎2
In both cases, 𝑛𝑥 and 𝑛𝑦 are integers reflecting the quantization of the particle.
Wavefunctions
When everything is put together, the total wavefunction for the particle in two-dimensional square box is:
2 𝑛𝑥 𝜋𝑥 𝑛𝑦 𝜋𝑦
𝛹(𝑥, 𝑦) = sin sin
𝑎 𝑎 𝑎
where 𝑛𝑥 and 𝑛𝑦 are independent integers. Figure 5 shows the first few wavefunctions of a particle in a
two-dimensional box. The associated probability densities are shown in Figure 6.

Figure 5 First few wavefunctions of a particle in a two-dimensional box.

Figure 6 Probability densities associated with first few wavefunctions of a particle in a two-
dimensional box.
Energies
The corresponding energies are:

ℎ2
𝐸= (𝑛2 + 𝑛𝑦2 )
8𝑚𝑎2 𝑥
where 𝑛𝑥 = 1,2,3, … and 𝑛𝑦 = 1,2,3, …

Figure 7 illustrate the energies of the particle in a two-dimensional box. Since the box is symmetric, there
exist a number of degeneracies, where different states end up having the same energy.

Figure 7 Energy levels for a particle in a two-dimensional box.


As can be seen from the above figure, the state characterized by the quantum numbers 𝑛𝑥 = 1, 𝑛𝑦 = 2
possesses the same energy as the state with quantum numbers 𝑛𝑥 = 2, 𝑛𝑦 = 1. Note that these
degeneracies can be ‘lifted’ by making the box asymmetric. For instance, if we were to elongate one side
of the box, (e.g., y direction), to a length 𝑏, the particle’s wavefunctions and energies would become:
2 𝑛𝑥 𝜋𝑥 𝑛𝑦 𝜋𝑦
𝛹(𝑥, 𝑦) = sin sin
√𝑎𝑏 𝑎 𝑏
and

ℎ2 𝑛𝑥2 𝑛𝑦2
𝐸= ( + )
8𝑚 𝑎2 𝑏 2

1.3 Particle in a Three-Dimensional Box: Degeneracies


Let us now consider the particle trapped inside a three-dimensional box having equal lengths on all sides
(i.e., a cube). The relevant potential and boundary conditions are shown in Figure 8, where a box with a
length 𝑎 on all three sides and a potential of zero (infinity) inside (outside).
Figure 8 Potential for a particle in a three-dimensional box.
The boundary conditions and the procedure explained for two-dimensional and one-dimensional systems
can be applied in this case in similar manner to obtain the resulting particle wavefunctions and energies
as:

2 3/2 𝑛𝑥 𝜋𝑥 𝑛𝑦 𝜋𝑦 𝑛𝑧 𝜋𝑧
𝛹(𝑥, 𝑦, 𝑧) = ( ) sin sin sin
𝑎 𝑎 𝑎 𝑎
and

ℎ2
𝐸= (𝑛2 + 𝑛𝑦2 + 𝑛𝑧2 )
8𝑚𝑎2 𝑥
In both the expressions, 𝑎 is the width of the box, while 𝑛𝑥 = 1,2,3, …, 𝑛𝑦 = 1,2,3, …, and 𝑛𝑧 = 1,2,3, …
are quantum numbers reflecting the confinement of the particle.
Figure 9 illustrates the first few energies of the particle in a three-dimensional box. Again, because of the
system’s high degree of symmetry, there are significant degeneracy, with multiple states having the same
energy.
Figure 9 Energy levels for a particle in a three-dimensional box.
Questions

1. Find the energies and normalized wavefunctions of a particle in a symmetric one-


1 1
dimensional box with barriers at 𝑥 = − 2 𝑎 and 𝑥 = 2 𝑎. Assume the free electron mass
𝑚0 .
2. Verify by spacing of energy levels shown in Figure 3 by using 𝐸(𝑛 + 1) − 𝐸(𝑛).

You might also like