You are on page 1of 14

A RESEARCH ASSINGMENT ON THE BEHAVIOUR OF A PARTICLE IN A

BOX MODEL IN RELATION TO QUANTUM MECHANICS

NAMES ADM.NO.
NAMES

ODONGO MAXWEL PS17/00075/19


OTANA WALTER PS17/00073/19
STANLEY K. MARUHI PS17/00006/19
SHARON WASONGA PS17/00082/19
MERCY JELAGAT PS17/00055/19
KIPKIRUI AMOS PS17/00040/19
SARAH LILLY ATIENO PS17/00072/19
PAUL KAZUNGU PS17/00003/19
BENSON APOLLO OTIENO PS17/00081/19

ASSIGNMENT/TASK; Using Quantum Mechanics, describe a particle in a box


model. (15marks)

may, 2021
ABSTRACT

We investigate why the particle-in-a-box (PB) model works well for


calculating the absorption wavelengths of cyanine dyes and why it does not
work for conjugated polyenes. The PB model is immensely useful in the
classroom, but owing to its highly approximate character there is little reason
to expect that it can yield quantitative agreement with experimental data.
However, we show that a more realistic behavior of the potential energy
leaves the highest occupied and the lowest unoccupied PB energy levels
unmodified as long as the variations in the potential are not too large (i.e., as
long as first-order perturbation theory is valid). The result explains the
success of the PB model in modeling absorption wavelengths of conjugated
delocalized chains with small bond-length alternation such as cyanine dyes.
In a next step, a suitable perturbation is applied to model conjugated
polyenes that exhibit pronounced bond-length alternation and have a
different behavior of the longest wavelength absorption as a function of
chain length. In this case, the perturbation directly affects the highest
occupied PB energy levels in a way that agrees with experimental
observations.

The recently proposed particle-in-a-box model of one-dimensional excitons


in conjugated polymers is applied in calculations of optical absorption and
electro absorption spectra. It is demonstrated that for polymers of long
conjugation length a superposition of single excitation resonances produces a
line shape characterized by a square-root singularity in agreement with
experimental spectra near the absorption edge. The effects of finite
conjugation length on both absorption and electro absorption spectra are
analyzed.

A simple two-particle model of excitons in conjugated polymers is proposed


as an alternative to usual highly computationally demanding quantum
chemical methods. In the two-particle model, the exciton is described as an
electron-hole pair interacting via Coulomb forces and confined to the
polymer backbone by rigid walls. Furthermore, by integrating out the
transverse part, the two-particle equation is reduced to one-dimensional
form. It is demonstrated how essentially exact solutions are obtained in the
cases of short and long conjugation length, respectively. From a linear
combination of these cases an approximate solution for the general case is
obtained. As an application of the model the influence of a static electric
field on the electron-hole overlap integral and exciton energy is considered
INTRODUCTION

In quantum mechanics, the particle in a box model (also known as the infinite

potential well or the infinite square well) describes a particle free to move in a small

space surrounded by impenetrable barriers. The model is mainly used as a

hypothetical example to illustrate the differences between classical and quantum

systems. In classical systems, for example, a particle trapped inside a large box can

move at any speed within the box and it is no more likely to be found at one position

than another. However, when the well becomes very narrow (on the scale of a few

nanometers), quantum effects become important. The particle may only occupy

certain positive energy levels. Likewise, it can never have zero energy, meaning that

the particle can never "sit still". Additionally, it is more likely to be found at certain

positions than at others, depending on its energy level. The particle may never be

detected at certain positions, known as spatial nodes. The solutions to the problem

give possible values of E and ψψ that the particle can possess. E represents allowed

energy values and ψ(x)ψ(x) is a wavefunction, which when squared gives us the

probability of locating the particle at a certain position within the box at a given

energy level.

The history of quantum mechanics began with a number of different scientific

discoveries: the 1838 discovery of cathode rays by Michael Faraday; the 1859–60

winter statement of the black-body radiation problem by Gustav Kirchhoff; the 1877

suggestion by Ludwig Boltzmann that the energy states of a physical system could

be discrete; the discovery of the photoelectric effect by Heinrich Hertz in 1887; and

the 1900 quantum hypothesis by Max Planck that any energy-radiating atomic system

can theoretically be divided into a number of discrete "energy elements" ε (Greek


letter epsilon) such that each of these energy elements is proportional to

the frequency ν with which each of them individually radiate energy, as defined by

the following formula:

where h is a numerical value called Planck's constant.

Then, Albert Einstein in 1905, in order to explain the photoelectric effect previously

reported by Heinrich Hertz in 1887, postulated consistently with Max Planck's quantum

hypothesis that light itself is made of individual quantum particles, which in 1926

came to be called photons by Gilbert N. Lewis. The photoelectric effect was observed

upon shining light of particular wavelengths on certain materials, such as metals,

which caused electrons to be ejected from those materials only if the light quantum

energy was greater than the work function of the metal's surface.

The phrase "quantum mechanics" was coined (in German, Quantenmechanik) by the

group of physicists including Max Born, Werner Heisenberg, and Wolfgang Pauli, at

the University of Göttingen in the early 1920s, and was first used in Born's 1924

paper "Zur Quantenmechanik".[1] In the years to follow, this theoretical basis slowly

began to be applied to chemical structure, reactivity, and bonding.

In 1926, Erwin Schrodinger was awarded Nobel Price for developing the

Schrodinger equation.
THEORY

To solve the problem for a particle in a 1-dimensional box, we must follow our Big,

Big recipe for Quantum Mechanics:

1. Define the Potential Energy, V

2. Solve the Schrödinger Equation

3. Define the wavefunction

4. Define the allowed energies

Step 1: Define the Potential Energy V

A particle in a 1D infinite potential well of dimension LL.

The potential energy is 0 inside the box (V=0 for 0<x<L) and goes to infinity at

the walls of the box (V=∞ for x<0 or x>L). We assume the walls have infinite

potential energy to ensure that the particle has zero probability of being at the

walls or outside the box. Doing so significantly simplifies our later mathematical

calculations as we employ these boundary conditions when solving the

Schrödinger Equation.

Step 2: Solve the Schrödinger Equation


The time-independent Schrödinger equation for a particle of mass m moving in one

direction with energy E is

 ћ is the reduced Planck Constant where  

 m is the mass of the particle

 ψ(x) is the stationary time-independent wavefunction

 V(x) is the potential energy as a function of position

 E is the energy, a real number

This equation can be modified for a particle of mass m free to move parallel to the x-

axis with zero potential energy (V = 0 everywhere) resulting in the quantum

mechanical description of free motion in one dimension:

This equation has been well studied and gives a general solution of:

ψ(x)=Asin(kx)+Bcos(kx)………………..(3)

where A, B, and k are constants.

Step 3: Define the wave function

The solution to the Schrödinger equation we found above is the general solution for a

1-dimensional system. We now need to apply our boundary conditions to find the

solution to our particular system. According to our boundary conditions, the


probability of finding the particle at x=0 or x=L is zero. When x=0, sin(0)=0

and cos(0)=1; therefore, B must equal 0 to fulfill this boundary condition giving:

ψ(x)=Asin(kx)………………(4)

We can now solve for our constants (A and k) systematically to define the

wavefunction.

Solving for k

Differentiate the wave function with respect to x:

dψ/dx=kAcos(kx)……………………(5)

d2ψ/dx2=−k2Asin(kx)……………….(6)

Since ψ(x)=Asin(kx), the

If we then solve for k by comparing with the Schrödinger equation above, we find:

k=(8π2mE/h2)^1/2………………….(8)

Now we plug k into our wave function:

ψ=Asin(8π^2mE/h^2)^1/2 x ……………..(9).

Solving for A

To determine A, we have to apply the boundary conditions again. Recall that

the probability of finding a particle at x = 0 or x = L is zero.

When x=L:
0=Asin(8π^2mE/h^2)^1/2L………..(10)

This is only true when

(8π^2mE/h^2)^1/2 L=nπ…………(11)

where n = 1,2,3…

Plugging this back in gives us:

ψ=Asin nπ/L x ………………….(12)

To determine A, recall that the total probability of finding the particle inside the box

is 1, meaning there is no probability of it being outside the box. When we find the

probability and set it equal to 1, we are normalizing the wave function.

For our system, the normalization looks like:

Using the solution for this integral from an integral table, we find our normalization

constant, A:

Which results in the normalized wavefunction for a particle in a 1-dimensional box:


Step 4: Determine the Allowed Energies

Solving for E results in the allowed energies for a particle in a box:

This is an important result that tells us:

1. The energy of a particle is quantized and

2. The lowest possible energy of a particle is NOT zero. This is called the zero-

point energy and means the particle can never be at rest because it always has

some kinetic energy.

This is also consistent with the Heisenberg Uncertainty Principle: if the particle had zero

energy, we would know where it was in both space and time.


APPLICATION

Because of its mathematical simplicity, the particle in a box model is used to

find approximate solutions for more complex physical systems in which a

particle is trapped in a narrow region of low electric potential between two high

potential barriers. These quantum well systems are particularly important

in optoelectronics, and are used in devices such as the quantum well laser,

the quantum well infrared photo-detector and the quantum-confined Stark

effect modulator. It is also used to model a lattice in the Kronig-Penney

model and for a finite metal with the free electron approximation.


CONCLUSION

The wavefunction for a particle in a box at the n=1n=1 and n=2n=2 energy levels look

like this:

The probability of finding a particle a certain spot in the box is determined by

squaring ψψ. The probability distribution for a particle in a box at

the n=1n=1 and n=2n=2 energy levels looks like this:

Notice that the number of nodes (places where the particle has zero probability of

being located) increases with increasing energy n. Also note that as the energy of the

particle becomes greater, the quantum mechanical model breaks down as the energy

levels get closer together and overlap, forming a continuum. This continuum means

the particle is free and can have any energy value. At such high energies, the classical
mechanical model is applied as the particle behaves more like a continuous wave.

Therefore, the particle in a box problem is an example of Wave-Particle Duality

FINDINGS

 The energy of a particle is quantized. This means it can only take on discrete

energy values.

 The lowest possible energy for a particle is NOT zero (even at 0 K). This

means the particle always has some kinetic energy.

 The square of the wave function is related to the probability of finding the

particle in a specific position for a given energy level.

 The probability changes with increasing energy of the particle and depends on

the position in the box you are attempting to define the energy for

 In classical physics, the probability of finding the particle is independent of

the energy and the same at all points in the box


ACKNOWLEDGEMENT

To make this work a success, we want to acknowledge every effort of the

contributors.

Firstly, much acknowledgements to our dear lecturer Mr Omanga

Kenyanya for enriching us with this vast information and skills that enabled

us to successfully undertake this task, lots of thanks.

To the entire team of students who participated in the discussion and

research to bring content to the table, making us succeed as a team.

Lastly, much thanks to the university administration for providing research

materials in the library like text books, that formed our base for the

knowledge and reference.


REFERENCES

1. Chang, Raymond. Physical Chemistry for the Biosciences. Sansalito, CA:

University Science, 2005.

2.  Zory, Peter (1993). Quantum Well Lasers. San Diego: Academic Press

Unlimited.

3.  "Quantum Dots : a True "Particle in a Box" System". PhysicsOpenLab. 20

November 2015. Retrieved  5 November 2016.

4. Hanle, P.A. (December 1977), "Erwin Schrodinger's Reaction to Louis de Broglie's Thesis on

the Quantum Theory.", Isis, 68 (4): 606–09, doi:10.1086/351880

5. Max Born, My Life: Recollections of a Nobel Laureate, Taylor & Francis, London, 1978. ("We

became more and more convinced that a radical change of the foundations of physics was

necessary, i.e., a new kind of mechanics for which we used the term quantum mechanics.

This word appears for the first time in physical literature in a paper of mine...")

6. D. Edwards, The Mathematical Foundations of Quantum Field Theory: Fermions, Gauge

Fields, and Super-symmetry, Part I: Lattice Field Theories, International J. of Theor. Phys., Vol.

20,

You might also like