You are on page 1of 35

MINISTRY OF EDUCATION AND TRAINING

NONG LAM UNIVERSITY

FACULTY OF CHEMICAL ENGINEERING AND FOOD SCIENCE

Course: Physics 2

Module 4

QUANTUM MECHANICS

Instructor: Dr. Nguyen Thanh Son

Academic year: 2023-2024


1
Module 4: Quantum mechanics

4.1 Effects of quantum


4.2 Wave-particle duality of matter
4.3 Introduction to quantum mechanics
4.3.1 Introduction to quantum mechanics
4.3.2 Schrödinger equation
4.3.3 Some examples

2
4.1 Effects of quantum

♦ Quantum effects

• There are several classes of phenomena that appear under quantum mechanics and have no analogues in

classical physics. In other words, there are classes of phenomena that classical physics cannot account for,

sometimes referred to as "quantum effects", including

- the quantization (discretization) of certain physical quantities,

- the wave-particle duality,

- the uncertainty principle,

- the quantum entanglement.

4.1.1. Quantization of certain physical quantities

Quantization first arose in the mathematical formulae of Max Planck. Examples of quantized observables

include angular momentum, the total energy of a bound system, and the energy contained in an

electromagnetic wave of a given frequency.

Quantization of energy

• Max Planck discovered a constant that when multiplied by the frequency of any electromagnetic wave

gives the energy of the wave. This constant is referred to by the letter h in mathematical formulae and is

called Planck’s constant. This discovery is a cornerstone of modern physics. By measuring the energy in a

discrete (non-continuous) portion of the wave, one recognized that the wave takes on the appearance of

chunks or packets of energy. These chunks of energy resemble particles. So energy is said to be

quantized because it only comes in discrete chunks instead of a continuous range of energies.

• If you look at the spectrum of light emitted by energetic atoms (such as the orange-yellow light from

sodium vapor street lights, or the blue-white light from mercury vapor lamps), you will notice that it is

3
composed of individual lines of different colors. These lines represent the discrete energy levels of the

electrons in those excited atoms. When an electron in a high energy state jumps down to a lower one, the

atom emits a light photon whose energy corresponds to the exact energy difference of these two levels

(conservation of energy). The bigger the energy difference, the more energetic the photon will be, and the

closer its color will be to the violet end of the visible spectrum. If electrons were not restricted to discrete

energy levels, the spectrum from an excited atom would be a continuous spread of colors from red to violet

with no individual lines.

• According to the hydrogen atom model of Niels Bohr, electrons move along a circular trajectory around

the proton like the earth around the sun, as shown in Figure 34.

Figure 34: Trajectory of an electron in a hydrogen atom, as used in the Bohr model.

• The spectrum of electromagnetic radiation from an excited hydrogen gas was yet another experimental

finding, which was difficult to explain since the spectrum is discrete rather than continuous. The emitted

wavelengths were associated with a set of discrete energy levels En described by

13.6 eV
En = − , with n = 1, 2,... (142)
n2

4
n is called the quantum number.

• Each value of n corresponds to a quantum state of the atom; n = 1 corresponding to the ground state (a

very stable state, lowest energy level); n = 2, 3, …. ∞ corresponding to excited states (unstable states,

higher energy levels).

• From Equation (142), we see that for any quantum number, n, there exists a single orbit with an electron

energy level, En, which can be calculated.

• When electrons transfer between energy levels, they either absorb energy (excite) or emit energy (de-

excite).

• When an electron gains energy, by absorbing a photon, it rises to a higher energy level (excitation), as

shown in the figure below.

• When an electron loses energy, by emitting a photon, it falls to a lower energy level (de-excitation), as

shown in the figure below.

5
• Energy of the emitted (or absorbed) photon Eph equals the energy difference that is released when an

electron makes a transition between a higher energy Ei and a lower energy Ej, as shown in Figure 35.

Eph = ∆E = Ei – Ej = hfij = hc/λij with i > j (143)

• From (142) and (143), we obtain

Eph = ∆E = Ei – Ej = hfij = hc/λij = 13.6(1/i2 – 1/j2) (eV) (143’)

with i, j are integers and i > j.

Ei

Figure 35: hfij Eph = hfij


Ej

Figure 36: Energy levels and possible electronic

transitions in a hydrogen atom. Shown are the first

six energy levels, as well as three possible

transitions involving the lowest energy level (n = 1).

6
• The maximum photon energy emitted from a hydrogen atom is 13.6 eV. This energy is called one

Rydberg. The electron transitions and the resulting photon energies are further illustrated by Figure 36.

• The figure below shows the visible portion of spectral emission lines of a hydrogen gas. Each color line

corresponds to a discrete energy.

• It is the fact that electrons can only exist at discrete energy levels which prevent them from spiraling into

the nucleus, as classical physics predicts. And it is this quantization of energy, along with some other

atomic properties that are quantized, gives quantum mechanics its name.

Quantization of electronic angular momentum

• Rutherford proposed that electrons orbit about the nucleus of an atom. One problem with this model is

that, classically, orbiting electrons experience a centripetal acceleration, and accelerating charges lose

energy by radiating; a stable electronic orbit is thus classically forbidden. Bohr nevertheless assumed

stable electronic orbits with the electronic angular momentum quantized as

L = mvr = nħ (144)

where n = 1, 2, 3, ... is order of electron level or the quantum number and ħ = h/(2π) is Dirac’s

constant.

• Bohr’s model has electrons in orbits around a central nucleus, but allows only certain orbits for

electrons. Treating the orbit of an electron as a continuous wave, the path length (or circumference, 2πr)

must be, according to de Broglie’s relation, equal to a whole number of wavelengths, i.e. 2πr = nλ. Later

we will see λ = h/mv, according to de Broglie’s relation. As a result, for a stable orbit, angular momentum

must be a multiple of h/2π = ħ. Angular momentum of electrons is thus quantized, as expressed in

Equation (144).

7
• The two below figures show the Bohr model of the hydrogen atom, where a particle wave fits exactly

onto an orbit, where n = 4 and n = 6, respectively.

n=6

• Quantization of angular momentum means that the radius of the orbit and the energy will be quantized as

well. Bohr assumed that the discrete lines seen in the spectrum of the hydrogen atom were due to

transitions of an electron from one allowed orbit/energy to another. He further assumed that the energy for

a transition is acquired or released in the form of a photon as proposed by Einstein, so that

Eph = ∆E = Ei – Ej = hfij = hc/λij (145)

corresponding to a transition between a higher energy level Ei and a lower energy level Ej. Where fij is the

frequency and λij is the wavelength of photon emitted (or absorbed) in the transition.

8
• Equation (145) expresses the Bohr frequency condition. This condition, along with Bohr's expression for

the allowed energy levels, gives a good match to the observed hydrogen atom spectrum. However, it works

only for atoms with one electron.

Example a What is the wavelength of the photon emitted when the electron in the hydrogen atom

makes a transition from the n = 3 state to the n = 2 state? Given h = 6.63x10 -34 J.s. Ans. 653 nm

Solution

Using (143) and (145) with j = 2 and i = 3 leads to λ = 653 nm.

Example b For an electron in a hydrogen atom in the n = 2 state, compute: (a) the angular
-34
momentum and (b) the total energy. Given h = 6.63x10 J.s.
2
Ans. (a) L = h/π kg.m /s; (b) E = –3.4 eV
2
Solution
(a) Using (144) L = nħ = nh/(2π) with n = 2 leads to L = h/π kg.m2/s.
(b) Using (142) with n = 2 leads to E = –3.4 eV.
2

Example c What is the frequency of the photon absorbed when the hydrogen atom makes the
transition from the ground state to the n = 4 state? Given h = 6.63x10=34 J.s.
Ans. f41 = 3.038 x 1015 Hz = 3038 THz
Solution
Using (143) and (145) with j = 1 and i = 4 leads to f41 = 3.038 x 1015 Hz.

Quantization of electronic charge

• The values of electric charge of charged particles are also quantized. Mathematically, we have

q = ne (146)

where q is the symbol used to represent electronic charge, n is an integer (n = 0, ±1, ±2, ±3, …), and e is

the elementary charge, 1.60 x 10-19 coulombs.

4.1.2 Uncertainty principle

• This principle underlies the phenomenon that simultaneous measurements of two or more observables

may possess a fundamental limitation on accuracy. This implies that position and momentum can never be

simultaneously measured with arbitrary precision, even in principle: as the precision of the position

measurement improves, the precision of the momentum measurement decreases, and vice versa. Those

9
variables for which it holds (e.g., momentum and position, or energy and time) are canonically conjugate

variables in classical physics.

• This principle that states that the position and velocity of an object cannot both be measured exactly

at the same time, and that the concepts of exact position and exact velocity together have no meaning

in nature. Articulated by Werner Heisenberg in 1927, it applies only at the small scales of atoms and

subatomic particles and is not noticeable for macroscopic objects, such as moving vehicles. Any attempt to

measure the velocity of a subatomic particle precisely will displace the particle in an unpredictable way,

thus invalidating any simultaneous measurement of its position. This displacement is a result of the wave

nature of particles (see wave-particle duality). The principle also applies to other related pairs of variables,

such as energy and time.

• According to this principle, a particle cannot have a well-defined position and simultaneously a well-

defined speed (or momentum). This principle states that certain pairs of physical properties (like

position/speed or time/energy) cannot have simultaneously well-defined values. In addition, performing

a measurement on a quantum system inescapably affects the system. The measurement of one

characteristic of a quantum entity inherently affects the values of other characteristics of that entity. This is

not due to a simple lack of subtlety in the design of experiments, but is a fundamental attribute of all

quantum entities.

• The uncertainly principle does not limit the precision with which either the particle's momentum or its

position may be measured, but only the precision with which both may be measured simultaneously. This

uncertainty is expressed mathematically as

(∆x) (∆px) ≥ ħ/2 = h/4π (147)

where ∆x indicates the uncertainty (that is, variance) in the measurement of position, ∆px the uncertainty in

the momentum measurement and ħ = h/(2π). It means that x and p x are canonically conjugate variables.

10
• Similarly, we also have

(∆E) (∆t) ≥ ħ/2 = h/4π (148)

where ∆E indicates the uncertainty in the measurement of energy, ∆t the uncertainty in the time

measurement. Equation (148) again means that E and t are also canonically conjugate variables.

Example A proton is confined to a nucleus whose diameter is 5.5x10 -15 m. If this distance is

considered to be the uncertainty in the position of the proton, what is the minimum uncertainty in its

momentum? Given h = 6.63x10-34 J.s. Ans. 0.96x10 -20 kg.m/s

Solution
Using (147) (∆x) (∆p ) ≥ ħ/2 = h/4π, we have
x

∆px ≥ (ħ/2)/∆x or ∆p xMin = (ħ/2)/∆x


With ∆x = 5.5x10-15 m ∆pxMin = 0.96x10-20 kg.m/s.

4.1.3 Wave-particle duality

• It has been shown that, under certain experimental conditions, microscopic objects like atoms or

electrons exhibit particle-like behavior, such as scattering. ("Particle-like" in the sense of an object that can

be localized to a particular region of space.) Under other conditions, the same type of objects exhibit wave-

like behavior, such as interference, diffraction. We can observe only one type of property at a time, never

both at the same time.

• The wave-particle duality will be explained in great detail in Section 4.2.

4.1.4 Quantum entanglement

• In some cases, the wavefunction of a system composed of many particles cannot be separated into

independent wavefunctions, one for each particle. In that case, the particles are said to be "entangled". If

quantum mechanics is correct, entangled particles can display remarkable and counter-intuitive properties.

For example, a measurement made on one particle can produce, through the collapse of the total

wavefunction, an instantaneous effect on other particles with which it is entangled, even if they are far

apart.

11
• Consider a system of two interacting particles. It is possible that one particle interacts with the other in

such a way that the quantum states of the two particles of the system form a single entangled state. The

definition of an entangled state is that the state of one particle is not entirely independent of the state of the

other: one particle’s state is dependent on another particle’s state in some way. Because of this

dependency, it is a mistake to consider either state in isolation from the other. Rather we should combine

the states and treat the result as a single entangled state.

• Quantum entanglement, also called the quantum non-local connection, is a property of a quantum state

of a system of two or more objects in which the quantum states of the constituting objects are linked

together so that one object can no longer be adequately described without full mention of its counterpart

(counterparts) - even if the individual objects are spatially separated in a space-like manner. The property

of entanglement was understood in the early days of quantum theory, although not by that name. This

interconnection leads to non-classical correlations between observable physical properties of remote

systems.

♦ Principal findings and predictions

• The principle findings and predictions of quantum mechanics are:

- Light, along with all electromagnetic radiation, is not emitted in a continuous stream of

energy, but in very small units of predetermined size, called quanta - from which the theory derives its

name.

- There is a fixed relationship between the wavelength of light, and the amount of energy

contained in a quantum of the light.

- Not only light, but also energy and a number of other fundamental entities (such as electric

charge) are quantized. Indeed, space and time themselves are generally thought to be quantized, too,

although the details are still obscure.

- Classical physics holds that light is comprised of waves, i.e. traveling disturbances in the

electromagnetic field, but light also (most paradoxically) appears to have characteristics of particles, i.e.

entities which are of fixed size and form. It is for this reason that the quanta of electromagnetic waves are

also called photons - i.e. light particles - using the -on ending traditionally reserved for particles.
12
- Not only do things usually thought of as waves have particle-like aspects, but things usually

thought of as particles (e.g. electrons) also have wave-like aspects; this wave-particle duality is now seen

as an inherent aspect of all quantum entities.

- This raises the question: what is a particle anyway? The naive model, that it is something like

a small ball, is clearly - once again - the result of the incorrect assumption that the world at the quantum

level looks the same as it does at the level of reality we experience, only much smaller.

- Many processes at the quantum level are only seemingly deterministic; i.e. while their

behavior, when measured in large numbers, follows some law (as in our coin-flipping example), individual

events are not predictable. For example, given a large amount of a radioactive element, it is possible to

accurately predict how many of those atoms will decay in a particular amount of time. It is, however,

impossible to predict if, and when, any particular atom will decay.

- Quantum mechanics also appears to indicate that among many attributes of a quantum entity,

there may be an attribute, such as spin, that does not have a fixed, definite value until it is measured. In

other words, that attribute (or, to be precise, its value) in some sense does not exist until it is measured.

This particular point has been a source of much debate since the 1920s, which continues to these days.

♦ Correspondence principle

• The correspondence principle states that the predictions of quantum mechanics reduce to those of

classical physics when a system becomes large. This "large system" limit is known as the classical or

correspondence limit.

• One can therefore start from an established classical model of a particular system, and attempt to guess

the underlying quantum model that gives rise to the classical model in the correspondence limit.

4.2. Wave-particle duality of matter

♦ Waves and particles

• At macroscopic scales we are used to two broad types of phenomenon: waves and particles. Briefly,

particles are localized phenomena which transport both mass and energy as they move, while waves are

de-localized phenomena (that is they are spread out in space) which carry energy but no mass as they

13
move. Physical objects that one can touch are particle-like phenomena (e.g. cricket balls), while ripples on

a lake (for example) are waves (note that there is no net transport of water; hence no net transport of mass).

• Quantum mechanics acknowledges the fact that particles exhibit wave properties. For instance, particles

can produce interference patterns and can penetrate or "tunnel" through potential barriers. Neither of these

effects can be explained using Newtonian mechanics. Photons, on the other hand, can behave as particles

with well-defined energy. These observations blur the classical distinction between waves and particles.

• Two specific experiments demonstrate the particle-like behavior of light, namely the photoelectric effect

and blackbody radiation. Both can only be explained by treating photons as discrete particles whose energy

is proportional to the frequency of the light. The emission spectrum of an excited hydrogen gas

demonstrates that electrons confined to an atom can only have discrete energies. Niels Bohr explained the

emission spectrum by assuming that the wavelength of an electron wave is inversely proportional to the

electron’s linear momentum.

• The particle and wave pictures are both simplified forms of the wave packet, a localized wave consisting

of a combination of plane waves with different wavelength. As the range of wavelength is compressed to a

single value, the wave packet becomes a plane wave with a single frequency and yields the wave picture.

As the range of wavelength is increased, the size of the wave packet is reduced, yielding a localized

particle.

♦ The photoelectric effect

• The photoelectric effect demonstrates the quantized nature of light: when applying monochromatic light

to a metal in vacuum, one finds that electrons are released from the metal. This observation confirms the

notion that electrons are confined to the metal, but can escape when provided sufficient energy, for

instance in the form of light. However, the surprising fact is that when illuminating with light of long

wavelengths (typically larger than 400 nm), no electrons are emitted from the metal even if the light

intensity is increased.

• On the other hand, one easily observes electron emission at ultra-violet wavelengths for which the

number of electrons emitted does vary with the light intensity. A more detailed analysis reveals that the

14
maximum kinetic energy of the emitted electrons varies linearly with the inverse of the wavelength, for

wavelengths shorter than the maximum wavelength (threshold wavelength).

• The experimental set-up to measure the photoelectric effect is shown in Figure 37.

Figure 37: Experimental set-up to measure the photoelectric effect.

• The experimental apparatus consists of two metal electrodes within a vacuum chamber. Light is incident

on one of two electrodes to which an external voltage is applied. The external voltage is adjusted so that the

electric current due to the photo-emitted electrons becomes zero. This voltage corresponds to the maximum

kinetic energy, EkinMax, of the electrons in units of electron volt, called the stopping potential and denoted

by V0.

• The maximum kinetic energy is measured for different wavelengths and is plotted versus the inverse of

the wavelength, as shown in Figure 38. The resulting graph is a straight line.

15
Figure 38: Maximum kinetic energy, EkinMax, of electrons emitted from a metal upon illumination with
photon energy, Eph. The maximum kinetic energy EkinMax is plotted versus the inverse of the
wavelength of the light.

• Albert Einstein explained this experiment by postulating that the energy of light is quantized. He assumed

that light consists of individual particles called photons, so that the maximum kinetic energy of the emitted

electrons is

EkinMax = K.E = (1/2)mev0max2 = Eph – qeΦM = Eph – φ (149a)

where Eph is the energy of the incident photon, me is the rest mass of electron (me = 9.1x10 -31 kg) and

φ = qeΦM is the minimum energy required to extract the electrons from the metal, called the photoelectric

workfunction of the metal and ΦM quantifies the potential which the electrons have to overcome in order to

leave the metal, v0max is the maximum velocity of emitted electron.

• The maximum kinetic energy EkinMax is related to the stopping potential V0 by the relation

EkinMax = K.E = (1/2)mev0max2 = eV0 (149b)

V0 quantifies the potential which stops the photoelectric current completely, is positive and measured in

volts.

• The energy of the photon Eph is given by

hc
E ph = hf = (150)
λ

where h is Planck's constant, c the speed of light in vacuum, f the frequency of the light, and λ the

16
wavelength of the light.

• Thus the condition for the photoelectric effect to occur is

hf ≥ φ (151)

Let φ = hf0 (152)

• The condition for the photoelectric effect to occur becomes

f ≥ f0 (153)

fo is called the threshold (cutoff) frequency of the metal.

• Using f = c/λ and letting f0 = c/λ0 (153’) , Equation (153) becomes

λ ≤ λ0 (154)

λ0 is called the threshold wavelength of the metal. φ, f0 and λ0 depend on the nature of the metal of

interest.

• Combining Equation (149a) and Equation (150), we have

EkinMax = Eph – φ = hf – φ = hc/λ – φ (155)

• EkinMax thus depends on the frequency of light falling on the metal surface, but not on the intensity of the

incident light.

Example a (27.6, Serway and Faughn) A sodium-vapor lamp has a power output of 1000 W.

Using 589.3 nm as the average wavelength of this light source, calculate the number of photons emitted per

second. Ans. Eph = 3.38x10-19 J; N = P.t/Eph = 2.96x1021

Example b Sodium has a photoelectric workfunction of 2.36 eV. If a scientist illuminates a piece

of sodium with a 450 nm wavelength light, what are the:

(a) maximum kinetic energy of the emitted electrons, and

(b) emitted electron’s maximum speed?

17
Given: 1 eV = 1.6 x 10 -19 J, me = 9.1x10 -31 kg and h = 6.63 x 10 -34 J.s.

Ans. (a) EkinMax = 0.4 eV (b) v0Max = 0.275 x 106 m/s

Solution

(a) Using (155 EkinMax = Eph – φ = hc/λ – φ EkinMax = 0.4 eV.

(b) From EkinMax = (1/2)mev0max2 v0max = 0.275 x 10 6 m/s.

Example A 600-nm light falls on a metal surface and electrons with the maximum kinetic energy of

0.17 eV are emitted. Determine (a) the work function and (b) the threshold (cutoff) frequency of the metal.

Given h = 6.63x10 -34 J.s. Ans. a. 1.89 eV; b. 459 THz.

Solution

(a) Using (155) EkinMax = Eph – φ = hc/λ – φ, we find φ = 1.89 eV.

(b) Using (152) φ = hf0 f0 = 0.459 x 1012 Hz = 459 THz.

♦ Wave-particle duality of matter

• While other light-related phenomena such as the interference of two coherent light beams demonstrate the

wave characteristics of light, it is the photoelectric effect, which demonstrates the particle-like behavior of

light.

• These experiments lead to the particle-wave duality concept, namely that particles observed in an

appropriate environment behave as waves, while waves can also behave as particles. This concept

applies to all waves and particles. For instance, coherent electron beams also yield interference patterns

similar to those of light beams.

• It is the wave-like behavior of particles, which led to de Broglie wavelength. Since particles have wave-

18
like properties, for each particle there is an associated wavelength, called de Broglie wavelength and given

by

h
λ=
p (156)

where λ is the wavelength, h is Planck's constant and p is the particle’s linear momentum.

• By combining equations (150) and (156), we can also show that for light photons

Eph = pc (157)

• Einstein showed that the linear momentum of a photon is

Pph = h/λ (158)

• This can be easily done as follows. Assuming Eph = hf for a photon and λf = c for an electromagnetic

wave, we obtain

Eph = hc/λ (159)

• Now we use Einstein's relativity result, Eph = mc2, to find that for a photon

λ = h/(mphc) (160)

which is consistent with Equation (156) with p ph = mphc being the photon’s linear momentum and vph = c

the photon’s velocity.

Note that m here refers to the relativistic mass, not the rest mass, since the rest mass of a photon is

zero.

• Since light can behave both as a wave (it can be diffracted and it has a wavelength) and as a particle (it

contains packets of energy hf), de Broglie reasoned in 1924 that matter can also exhibit this wave-particle

duality. He further reasoned that matter would obey Equation (156) as light. In 1927, Davisson and

Germer observed diffraction patterns by bombarding metals with electrons, confirming de Broglie's

proposition.

19
• De Broglie's equation offers a justification for Bohr's assumption (Equation 144). If we think of an

electron as a wave, then for the electron’s orbit to be stable the wave must complete an integral number of

wavelengths during its orbiting. Otherwise, it would interfere destructively with itself. This condition may

be written as

2πr = nλ

• If we use the de Broglie relation (156) and p = mv, this can be rewritten as

mvr = nħ

which is identical to Bohr's equation (144).

Example a The radius of the nth orbit of the electron of a hydrogen atom is given by rn = n2a0,

where a0 = 0.0539 nm. Find the speed of the electron in the first (ground-state) Bohr orbit of the hydrogen

atom. Ans. v1 = 2.15x106 m/s

Example b For an electron having a de Broglie wavelength of 0.167 nm (appropriate for

interacting with crystal lattice structures that are about this size):

(a) Calculate the electron’s velocity, assuming it is nonrelativistic;

(b) Calculate the electron’s kinetic energy in eV.

Ans. (a) 4.36×10 6 m/s; (b) 54.0 eV

20
Solution

(a) From p = h/λ (158) and p = mv (nonrelativistic approximation) v = h/(mλ).

Substituting leads to v = 4.36×106 m/s.

(b) KE = (1/2)mv2 (nonrelativistic approximation) KE = 54.0 eV.

Example c

(a) What is the wavelength of a photon that has a momentum of 5.00 x 10 −29 kg.m/s?

(b) Find its energy in eV.

Given h = 6.63x10 -34 J.s and c = 3.0x108 m/s. Ans (a) 13.3 µm; (b) 9.38 x 10 −2 eV

Solution

(a) Using (156) λ = h/p λ = 13.3 x 10 −6 m = 13.3 µm.

(b) Using (157) Eph = pc Eph = 9.38 x 10−2 eV.

4.3 Introduction to quantum mechanics

4.3.1 Introduction to quantum mechanics

• Quantum mechanics is the description of motion and interaction of particles at small scales where the

discrete nature of the physical world becomes important. Quantum mechanics represents a fundamental

break with classical physics, in which energies and angular momenta were regarded as continuous

quantities that could change by arbitrary amounts.

• Quantum mechanics is a more fundamental theory than Newtonian mechanics and classical

electromagnetism, in the sense that it provides accurate and precise descriptions for many phenomena that

these "classical" theories simply cannot explain on the atomic and subatomic levels. It is necessary to use

quantum mechanics to understand the behavior of systems at atomic length scales and smaller. For

21
example, if Newtonian mechanics governed the workings of an atom, electrons would rapidly travel

toward and collide with the nucleus. However, in the natural world the electron normally remains in a

stable orbit around a nucleus, seemingly defying the classical electromagnetism theory.

• As mentioned earlier, Einstein postulated that electromagnetic radiation could exist only in discrete units,

called photons. This was followed by Bohr's postulate that the angular momentum of the electron orbiting

the nucleus of a hydrogen atom was quantized, and led to a formula that correctly predicted the observed

line spectrum of the hydrogen atom. Another early development was the de Broglie wavelength and the

concept of wave-particle duality.

• Quantum mechanics (from the Latin quantus, "how much") is a theory in physics that explains and

predicts behaviors of matter and energy at very small scales, which is often unusual and sometimes

extremely counter-intuitive, since it is deeply in conflict with the ideas most people have of how the

physical world works. It is perhaps the most important building block in the revolution of physics (1900 -

1925 period) which erased the limitations of classical physics and created the physics of today.

• Quantum mechanics is a fundamental branch of theoretical physics that replaces Newtonian mechanics

and classical electromagnetism at atomic and subatomic levels. It is the underlying mathematical

framework of many fields of physics and chemistry, including condensed matter physics, atomic physics,

molecular physics, computational chemistry, quantum chemistry, particle physics, and nuclear physics.

Along with general relativity, quantum mechanics is one of the pillars of modern physics.

• The discovery that waves could be measured in particle-like small packets of energy called quanta led to

the branch of physics that deals with atomic and subatomic systems which we today call quantum

mechanics.

• Quantum mechanics uses complex number wavefunctions (sometimes referred to as orbitals in the case of

atomic electrons), and more generally, elements of a complex vector space to explain observed effects.

These are related to classical physics largely through probability.

22
• Probability in the context of quantum mechanics has to do with the likelihood of finding a system

in a particular state at a certain

time, for example, finding an

electron in a particular region

around the nucleus at a particular

time. Therefore, electrons cannot

be pictured as localized particles

in space but rather should be

thought of as "clouds" of negative

charge spread out over the entire

orbit. Examples are shown in the

figure on the right.

• These clouds represent the

regions around the nucleus where the probability of "finding" an electron is the largest. This

probability cloud obeys a quantum mechanical principle called Heisenberg's uncertainty principle, which

states that there is an uncertainty in the classical position of any subatomic particle, including the

electron; so instead of describing where an electron or other particle is, the entire range of possible values

is used, describing a probability distribution.

At the left, the old view of a simple atom. At the right, the modern view.

• As a result, in normal atoms with electrons in stationary states, the probability of the electron being

within the nucleus (or somewhere else in atom within similarly small volume) is nearly zero, according to

the uncertainty principle (it is nearly zero as the nucleus has a volume and is not a point). Therefore,
23
quantum mechanics, translated to Newton's equally deterministic description, leads to a probabilistic

description of nature.

• Another exemplar that led to quantum mechanics was the study of electromagnetic waves such as light.

When it was found in 1900 by Max Planck that the energy of waves could be described as consisting of

small packets or quanta, Albert Einstein exploited this idea to show that an electromagnetic wave such as

light could be described by particles called photons with discrete energy dependent on the wave frequency.

• Those findings mentioned above led to a theory of unity between subatomic particles and

electromagnetic waves called the wave-particle duality in which particles and waves were neither one nor

the other, but had certain properties of both. While quantum mechanics describes the world of very small

systems, it is also needed to explain certain "macroscopic quantum systems" such as superconductors and

superfluids.

• Generally, quantum mechanics does not assign definite values to observables. Instead, it makes

predictions about probability distributions; that is, the probability of obtaining each of the possible

outcomes from measuring an observable.

• The position and momentum of the particle are observables. The uncertainty principle of quantum

mechanics states that both the position and the momentum cannot simultaneously be known with infinite

precision at the same time.

4.3.2 Schrödinger equation

4.3.2.1 Schrödinger equation

• In quantum mechanics, there is wave-particle duality so the properties of a particle can be described as

those of a wave. Therefore, the particle’s quantum state can be represented by a wavefunction.

• Wavefunctions can change as time progresses. An equation, known as the Schrödinger equation,

describes how wavefunctions change in time, a role similar to that of Newton's second law in classical

mechanics.

• The Schrödinger equation, applied to a free particle, predicts that the center of a wave packet will move

through space with a constant velocity, like a classical particle with no forces acting on it. However, the

24
wave packet will also spread out as time progresses, which means that the position becomes more

uncertain.

• Some wavefunctions produce probability distributions that are constant in time. Many systems that are

treated dynamically in classical mechanics are described by such "static" wavefunctions. For example, a

single electron in an unexcited atom is pictured classically as a particle moving in a circular trajectory

around the atomic nucleus, whereas in quantum mechanics (QM), it is described by a static, spherically

symmetric wavefunction.

• The time evolution of wavefunctions is deterministic in the sense that, given a wavefunction at an initial

time, it makes a definite prediction of what the wavefunction will be at any later time. During a

measurement, the change of the wavefunction into another one is not deterministic, but rather

unpredictable, i.e., random.

• The approach suggested by Schrodinger was to postulate a function which would vary in both time and

space in a wave-like manner (the so-called wavefunction) and which would carry within it information

about a particle or system. The time-dependent Schrodinger equation allows us to deterministically predict

the behavior of the wavefunction over time, once we know its environment.

• The information concerning environment is in the form of the potential field V(x) which would be

experienced by the particle according to classical mechanics.

• Whenever we make a measurement on a quantum system, the results are dictated by the wavefunction at

the time at which the measurement is made. It turns out that for each possible quantity, we might want to

measure (an observable), there is a set of special wavefunctions (known as eigenfunctions) which will

always return the same value for the observable (an eigenvalue).

• Wavefunctions have a number of important properties, the key one being: For every quantum system,

there exists a wavefunction, from which all possible predictions of the system can be obtained.

In this context, "possible predictions" include the energy of the system, the linear momentum of

particles in the system, etc. Thus, if we know the wavefunction, we can extract from that wavefunction all

the information that can possibly be obtained.

25
• The Schrödinger equation is the fundamental equation of physics for describing quantum mechanical

behavior. It is also often called the Schrödinger wave equation, and is a partial differential equation that

describes how the wavefunction of a physical system evolves over time.

• We know that the total energy of a system is

(161)

Schrödinger converted this expression into a wave equation by defining a wavefunction, Ψ, and

multiplied each term in Equation (161) by that wavefunction:

(162)

ℏ 2∂ 2
• To incorporate the de Broglie wavelength of the particle, he introduced the operator, − , which
∂x2

provides the square of the linear momentum when applied to a plane wave propagating in the x direction:

(163)

where k is the wavenumber, which equals 2π /λ. Without claiming that this is an actual proof, we now

simply replace the linear momentum squared, p2, in Equation (162) by this operator yielding the time-

independent Schrödinger equation:

(164)

We have assumed a one - dimensional case.

• To illustrate the use of Schrödinger equation, we will present two solutions of Schrödinger equations, one

for an infinite quantum well and one for a harmonic oscillator. Prior to that, we discuss the physical

interpretation of the wavefunction.

26
4.3.2.2 Physical interpretation of the wavefunction

• The use of a wavefunction to describe a particle, as in the Schrödinger equation, is consistent with the

particle-wave duality concept. However, the physical meaning of the wavefunction does not naturally

follow.

• The value of Ψ2 for a particular object at a certain place and time is proportional to the probability

of finding the object at that place at that time. For this reason the quantity Ψ 2 is called the probability

density function of the object. A large value of Ψ2 means that the object is likely to be found at the

specified place and time; a small value of Ψ2 means that the object is less likely to be found at that place

and time.

• Quantum theory postulates that the product of the wavefunction Ψ (x) and its complex conjugate Ψ*(x)

is the probability density function, P(x), associated with the particle of interest.

(165)

• This probability density function P(x) integrated over a specific volume provides the probability that the

particle described by the wavefunction is within that volume. The probability function is frequently

normalized to indicate that the probability of finding the particle somewhere equals 100%. This

normalization enables one to calculate the magnitude of the wavefunction, using

(166)

• This probability density function P(x) can then be used to find all properties of the particle by using the

quantum operators. For example, to find the expected value <f (x)> of a function f(x) for the particle

described by the wavefunction Ψ, one calculates

+∞
< f ( x) > = ∫ Ψ ( x)F ( x )Ψ * ( x)dx (167)
−∞

where F(x) is the quantum operator associated with the function f(x) of interest. A list of quantum operators

corresponding to a selection of common classical variables is provided in Table 4.1.

27
Table 4.1: Selected classical variables and their corresponding quantum operators. Here i = (–1)1/2 is the
purely imaginary number.
• For example, if we let f(x) = x then F(x) = x. If we already know the function Ψ(x) that describes the

particle’s quantum state, we can find the expected value <x> by using Equation (167) and obtain

<x>= ∫ Ψ( x) xΨ ( x)dx
*
(168)
−∞

4.3.3 Some examples

4.3.3.1 The infinite quantum well

• The one-dimensional infinite quantum well represents one of the simplest quantum mechanical structures.

We use it here to illustrate some specific properties of quantum mechanical systems. The potential in an

infinite well is zero between x = 0 and x = Lx and is infinite on either side of the well; thus there is no

chance of finding the particle outside the well. The potential and the first five possible energy levels an

electron can occupy are shown in Figure 39.

Figure 39: Potential energy of an infinite quantum well, with width Lx. Also indicated are the lowest five
energy levels of the electron in the well.

28
• The energy levels in an infinite quantum well are calculated by solving the Schrödinger wave equation

(164) with the potential, V(x), as shown in Figure 39. As a result, one solves the following equation within

the well.

(169)

• The general solution to this differential equation is

(170)

where the coefficients A and B must be determined by applying the boundary conditions.

• Since the potential is infinite on both sides of the well, the probability of finding an electron outside the

well and at the well boundary equals zero. Therefore, the wave function must be zero on both sides of the

infinite quantum well or

(170)

• These boundary conditions imply that the coefficient B must be zero and the argument of the sine function

must equal a multiple of pi at the edges of the quantum well or

(171)

where the subscript n was added to the energy, En, to indicate the energy corresponding to a specific value

of n. The resulting values of the energy, En, are then equal to

(172)

Or En = [h2/(8mLx2)]n2 (172’)

Here En varies as n2, whereas in hydrogen En ∝ 1/ n2. The integer n is a quantum number

with which a state is labeled. The corresponding wave function is

29
(173)

where the coefficient A was determined by requiring that the probability of finding the electron in the well

equals unity (normalization condition) or

(174)

The asterisk denotes the complex conjugate.

• Note that the lowest possible energy (in the ground state, n = 1), namely E1, is not zero although the

potential is zero within the well. Only discrete energy values are obtained as eigenvalues of the Schrödinger

equation. The energy difference between two adjacent energy levels increases as the energy increases. An

electron occupying one of the energy levels can have a positive or negative spin (s = ½ or s = -½). Both

quantum numbers, n and s, are the only two quantum numbers needed to describe this system.

• The wavefunctions corresponding to each energy level are shown in Figure 40 (left). Each wavefunction

has been shifted by the corresponding energy and scaled with an arbitrary magnitude as is commonly done.

The probability density function, P(x) = Ψ(x)*Ψ(x), provides the probability of finding an electron at a

certain location in the well. These probability density functions are shown in Figure 40 (right) for the first

five energy levels. For instance, for n = 2 the electron is least likely to be in the middle of the well and at

the edges of the well and is most likely to be one quarter of the well width away from either edge.

Figure 40: (Left) Energy levels, wavefunctions and (right) probability density functions in an infinite

30
quantum well. The figure is calculated for a 10 nm wide well containing an electron with mass
m0. The wavefunctions and the probability density functions have an arbitrary magnitude (i.e.
they are not normalized) and are shifted by the corresponding electron energy.

• Note that classically the particle is equally likely to be found anywhere in the well, but this is not the

case here.

• The lowest energy for a particle in an infinite well is E1 = h2/(8mLx2) for n = 1 (the ground state). This is

called the zero-point energy.

4.3.3.2 The harmonic oscillator

• The next problem of interest is the harmonic oscillator, which is characterized by a quadratic potential,

V(x) = (1/2)kx2, yielding the following Schrödinger equation

(175)

where k is the spring constant which relates to the restoring force of the equivalent classical problem of a

mass connected to a spring for which

(176)

• The Schrödinger equation can be solved by assuming the solution to be of the form

(177)

where γ2 = mk/ħ2, which reduces Equation (175) to

(178)

• A polynomial of order n satisfies this equation if

(179)

so that the minimal energy is E1 = ħω/2 (n = 1, the ground state) while all the energy levels are separated

31
from each other by an energy ħ ω. The polynomials which satisfy Equation (178) are knows as the Hermite

polynomials of order n, Hn, so that the wavefunction Ψ(x) is

(180)

• As an example, we present the wavefunctions for the first three bound states in Figure 41.

Figure 41: Electron wavefunctions for the first three bound states of a harmonic oscillator.

♦ Applications of quantum mechanics

• Quantum mechanics has had enormous success in explaining many of the features of our world. The

individual behavior of the subatomic particles that make up all forms of matter - electrons, protons,

neutrons, and so forth - can often only be satisfactorily described by using quantum mechanics. Quantum

mechanics has strongly influenced on the string theory, a candidate for a theory of everything. It is also

related to statistical mechanics.

• Quantum mechanics is important for understanding how individual atoms combine covalently to form

chemicals or molecules. The application of quantum mechanics to chemistry is known as quantum

chemistry. The (relativistic) quantum chemistry can in principle mathematically describe most of

chemistry.

32
• Quantum mechanics can provide quantitative insight into ionic and covalent bonding processes by

explicitly showing which molecules are energetically favorable to others, and by approximately how

much. Most of the calculations performed in computational chemistry rely on quantum mechanics.

• Much of modern technology operates at a scale where quantum effects are significant. Examples include

lasers, transistors, electron microscopes, and magnetic resonance imagings (MRI). The study of

semiconductors led to the invention of diodes and transistors, which are indispensable for modern

electronics.

• Researchers are currently seeking robust methods of directly manipulating quantum states. Efforts are

being made to develop quantum cryptography, which will allow guaranteed secure transmission of

information.

• A more distant goal is the development of quantum computers, which are expected to perform certain

computational tasks exponentially faster than classical computers. Another active research topic is

quantum teleportation, which deals with techniques to transmit quantum states over arbitrary distances.

Example a An electron is confined to a 1 micron thin layer of silicon. Assuming that the

semiconductor can be adequately described by a one-dimensional quantum well with infinite walls,

calculate the lowest possible energy within the material in units of electron volt. If the energy is interpreted

as the kinetic energy of the electron, what is the corresponding electron velocity? (The effective mass of

electrons in silicon is 0.26 m0, where m0 = 9.11 x 10 -31 kg is the free electron rest mass).

Ans. E1 = 1.45 µeV; v1 = 1.399 x 106 m/s

Solution

For one-dimensional quantum well with infinite walls, we have

En = [h2/(8mLx2)]n2 (172’)

where n = 1, 2, …: quantum number.

Here Lx = 1 µm = 1.0 x 10-6 m and m = 0.26 m0 = 0.26 x 9.11 x 10-31 kg.

33
For the lowest energy (the ground state), we have n = 1.

Substituting leads to E1 = 1.45 µeV.

Using E1 = KE1 = (1/2)mv12, we find that v1 = 1.399 x 106 m/s.

Example b An electron is trapped in a one-dimensional infinite potential well of width 4.0×10 -10

m. Find the wavelength of photon emitted by the electron as it makes a transition from n = 2 state to the

ground state (n = 1) in the well.

Ans. λ21 = 176 nm

Solution

• Applying the general formula Eph = ∆E = Ei – Ej = hc/λij (145) with i = 2 and j = 1, we have

E2 – E1 = hc/λ21.

• For one-dimensional infinite potential, we have En = [h2/(8mLx2)]n2 (172’)

E1 = [h2/(8mLx2)]12 and E2 = [h2/(8mLx2)]2 2.

Combining those equations leads to hc/λ21 = [h2/(8mLx2)](4 – 1). Plugging numbers gives

λ21 = 176 nm.

REFERENCES

1) Halliday, David; Resnick, Robert; Walker, Jearl (1999), Fundamentals of Physics 7th ed. John Wiley &
Sons, Inc.
2) Feynman, Richard; Leighton, Robert; Sands, Matthew (1989), Feynman, Lectures on Physics. Addison-
Wesley.
3) Serway, Raymond; Faughn, Jerry (2003), College Physics 7th ed. Thompson, Brooks/Cole.

34
4) Sears, Francis; Zemansky Mark; Young, Hugh (1991), College Physics 7th ed. Addison-Wesley.
5) Beiser, Arthur (1992), Physics 5th ed. Addison-Wesley Publishing Company.
6) Jones, Edwin; Childers, Richard (1992), Contemporary College Physics 7th ed. Addison-Wesley.
7) Alonso, Marcelo; Finn, Edward (1972), Physics 7th ed. Addison-Wesley Publishing Company.
8) Michels, Walter; Correll, Malcom; Patterson, A. L. (1968), Foundations of Physics 7th ed. Addison-
Wesley Publishing Company.
9) Hecht, Eugene (1987), Optics 2th ed. Addison-Wesley Publishing Company.
10) Eisberg, R. M. (1961), Modern Physics, John Wiley & Sons, Inc.
11) Reitz, John; Milford, Frederick; Christy Robert (1993), Foundations of Electromagnetic Theory, 4th
ed. Addison-Wesley Publishing Company.
12) Giambattista Alan; Richardson, B. M; Richardson, R. C. (2004), College Physics, McGraw-Hill.
13) Websites:
http://en.citizendium.org/wiki/Quantum_mechanics
http://www.crystalinks.com/quantumechanics.html
http://www.answers.com/topic/quantum-mechanics
http://en.wikipedia.org/wiki/Quantum_mechanics
http://newton.ex.ac.uk/research/qsystems/people/jenkins/mbody/mbody2.html
http://scienceworld.wolfram.com/physics/QuantumMechanics.html
http://www.answers.com/topic/quantum-entanglement
http://www.ipod.org.uk/reality/reality_entangled.asp
http://en.citizendium.org/wiki/Category:Physics_Content
http://ecee.colorado.edu/~bart/book/book/chapter1/ch1_2.htm#fig1_2_2.

35

You might also like