You are on page 1of 22

Waves: part 3

5 Waves: part 3
In this section of the course we look in more detail at electromagnetic (e/m) radiation and the
electromagnetic spectrum. We will see how diffraction experiments confirm the wave-like nature of
electromagnetic radiation but we will also look at the experimental evidence that shows that e/m
radiation has particle-like properties. Thus, we introduce the curious concept of wave-particle
duality. In this context we will look briefly at the class of waves known as matter waves. Black body
radiation will be discussed before we conclude by returning to the topic of diffraction, and how it
can be used as tool for analysing e/m spectra and so obtaining information about the nature and
properties of an object.

5.1 Properties of electromagnetic radiation


When we introduced the topic of waves (section 3) we identified mechanical waves (e.g. sound
waves, water waves, waves on a string) as being distinct from electromagnetic waves (e.g. light,
radio waves, X-rays). The main difference between these classes of wave is that mechanical waves
can only propagate through a medium, such as air, water, string etc. Unlike mechanical waves,
however, e/m waves do not require a medium through which to travel. Electromagnetic radiation
can travel through materials such as air, glass or water but it can also travel though a vacuum (e.g.
light from a distant star can travel through space to reach us).

The reason that e/m radiation can travel through a vacuum is because of the nature of the oscillating
quantities that constitute the wave. Electromagnetic waves are transverse waves, in other words the
simple harmonic oscillation at each point on the wave happens in a direction perpendicular to the
direction of travel of the wave. However, in the case of e/m waves there are actually two oscillating
quantities, an electric field and a magnetic field 11. These fields are always perpendicular to one
another and, as stated above, perpendicular to the direction in which the wave is travelling. Both
fields vary sinusoidally, just like the transverse waves discussed in previous sections. In addition, the
electric and magnetic fields oscillate with the same frequency and are always in phase with each
other (see Figure 5.1).

Figure 5.1 A representation of an electromagnetic wave, showing the oscillating electric (red) and magnetic (blue) fields.
The fields are perpendicular to one another and to the direction of travel of the wave. The two fields are always in phase.

We know that electric and magnetic fields can exist in a vacuum, so it is no surprise that an e/m
wave can travel through a vacuum.

11
It was James Clerk Maxwell who showed that a beam of light is a travelling wave of electric and magnetic
fields. He explained that the varying magnetic field induces (via Faradays law) a varying electric field which, in
turn, induces (via Maxwells law of induction) a varying magnetic field, and so on. In this way the fields
continually generate each other and the resulting oscillations travel as an electromagnetic wave.

37
Waves: part 3

Like all waves, the speed of an e/m wave is given by /k, or alternatively by f. However, because
this is an e/m wave, its speed (in vacuum) is given the symbol c, rather than v. In common with
other waves, the speed of e/m waves depends on the properties of the medium through which it is
travelling. So what about when light travels in a vacuum? What determines c in a vacuum? An
expression for the speed of light in a vacuum comes out of Maxwells equations:

1
c= 5.1
0 0

where 0 and 0 are the permeability and permittivity of free space, respectively. Note that all
electromagnetic waves, including visible light, have the same speed (c = 3.00 108 m s1) in vacuum.

An interesting point to note, and one which was formalised by Einstein in his 1905 paper on special
relativity, is that 0 and 0 are fundamental and universal constants that do not depend on the
inertial frame of reference in which they are measured (in other words, their values do not change
according to the relative speed of the observer). The consequence of this is that the speed of light
(or any e/m wave) is the same, regardless of the frame of reference from which it is measured. This
is a unique property of e/m waves.

5.2 The electromagnetic spectrum


There is no inherent upper or lower limit to the wavelength (or frequency) range of electromagnetic
waves. This range in known as the electromagnetic spectrum and stretches from long wavelength
(low frequency) radio waves to short wavelength (high frequency) gamma rays. The division of the
spectrum into smaller ranges allows us to name particular regions (for example; microwave, infrared,
visible, ultraviolet, X-ray) and to characterise the waves in terms of their wavelength, frequency,
energy, and how they interact with matter and objects (see Figure 5.2). It should be noted that the
limits of the various sub-divisions are generally not well defined.

Figure 5.2 The electromagnetic spectrum

38
Waves: part 3

We know that, in a given medium, the velocity of the e/m wave is constant (and in a vacuum it
equals c). So, as illustrated in Figure 5.2, the frequency is inversely proportional to wavelength. Also,
as we would expect from our understanding of the intensity of travelling waves, we can see that the
energy associated with an e/m wave is proportional to the frequency of the wave (remember,
intensity is the rate of energy transfer per unit area). So, the higher the frequency, the greater the
energy. This leads to a further classification of e/m radiation; into ionising and non-ionising radiation.
Electromagnetic radiation at frequencies above about 1016 Hz (the ultraviolet part of the spectrum)
is sufficiently energetic that it can strip electrons from atoms that absorb it, hence the term ionising
radiation. This is a significant point because, particularly in living tissue, this has the potential to do
serious damage to material that is subject to the radiation.

Not all of the properties of e/m waves are related quite so simply as wavelength, frequency and
energy. Note, for example, that the ability of e/m radiation to penetrate the Earths atmosphere is
not a simple function of frequency. Due to the absorption characteristics of different gases and
water vapour in the atmosphere, there are certain wavelength windows which are transparent to
e/m waves, separated by ranges in which radiation of that wavelength cannot penetrate the
atmosphere. Similarly, and very significantly from a technological point of view, there are certain
well-defined wavelengths of light (from 0.7 to 1.4 m) that, due to the particular absorption
characteristics of silica, can travel through glass (in the form of glass fibres) without any significant
loss of intensity.

The different characteristics of radiation from the entire range of the electromagnetic spectrum
have now been utilised in some form of technological application; from radio waves (600 kHz to
110 Mhz) to mobile communications (900 Mhz to 2.4 Ghz) or wireless data transmission (2.4 GHz);
from medical imaging and security scanning using mm waves (0.1 10 THz) to visible and infrared
optical communication (1014 1015 Hz); and from medical, crystallographic and security scanning
using X-rays (1018 1019 Hz) to medical imaging techniques such as PET (positron emission
tomography) that utilise the gamma ray region (> 1020 Hz) of the spectrum. In addition to man-made
sources of e/m radiation across the spectrum, Table 5.1 includes some examples of natural sources
at the different frequencies.

Table 5.1 Classification of the electromagnetic spectrum, with examples of wavelength/frequency ranges and some
examples of typical sources (both natural and man-made).

39
Waves: part 3

5.3 The wave nature of electromagnetic radiation


Diffraction is a wave phenomenon. In sections 4.3 and 4.4 we stated that diffraction was the change
in direction of a 2-D or 3-D wave as it passed close to an object. In addition, by using Huygens
construction it is possible to determine the position of successive wavefronts as they interact with
an object and so one can predict the outcome of, for example, the diffraction of waves as they pass
through a narrow gap. In this way it is easy to show that waves spread out after passing through a
gap (or slit) and that the narrower the gap (in comparison to the wavelength) the more the waves
are spread out.

In the following section we are going to consider an experiment that relies on the diffraction of light
and the resulting interference of the diffracted waves to produce a visible interference pattern that
is easily interpreted and fully explained by classical wave theory. In doing so, we will therefore show
that light (e/m radiation) behaves like a wave.

5.4 Youngs double slit experiment


Thomas Young, in 1801, devised an experiment to prove that light is a wave 12 by demonstrating that
light undergoes interference, just like waves of all other types (water waves, sound waves etc). He
used an arrangement of slits to diffract light such that the spreading out of the wavefronts would
cause overlapping of the waves and so result in interference.

Copyright2011 John Wiley & Sons, Inc. All rights reserved.


Figure 5.3 Young's double-slit interference experiment. The interference results from the diffraction of light by slits S1
and S2 in screen B. The interference pattern is visible if a viewing screen (C) is used to intercept the light. Note that this
schematic is not drawn to scale.

12
This was contrary to what most other scientists thought at the time

40
Waves: part 3

Figure 5.3 shows the basic set-up (not to scale) of Youngs double-slit experiment. Incident
monochromatic light (light of a single wavelength) 13 is diffracted by slit S0, which acts as a point
source of circular wavefronts. As these wavefronts reach screen B, the light passing through slits S 1
and S 2 is diffracted. Light from each slit overlaps and interferes with light from the other slit. In
accordance with what we know about interference, where the overlapping waves are in phase (i.e.
where wavefronts from slits S 1 and S 2 intersect) we get constructive interference. Where the
overlapping waves are exactly out of phase we get fully destructive interference. If we use a viewing
screen to observe the pattern created by the interference, we see alternating maxima (bright bands,
or fringes) and minima (dark bands or fringes) extending across the screen.

Figure 5.4 Double-slit interference pattern showing the alternating maxima and minima known as interference fringes

The Youngs double-slit interference experiment produces the expected (according to wave theory)
interference pattern. But what, exactly, determines the positions of the fringes? The analysis is
relatively straightforward and is illustrated using the set-up shown in Figure 5.5.

Figure 5.5 Double-slit interference experiment (not to scale) (a) waves from the two slits combine at point P on the
viewing screen. (b) Where the distance from the slits to the screen is >> D we can approximate the two rays as being
parallel

13
Thomas Young would not have had a truly monochromatic source. It is likely that he used sunlight, projected
from a small hole in a window shutter. Although sunlight is not monochromatic it does have a distinct peak
wavelength (of about 555 nm). Young was not only able to observe an interference pattern but also to make a
measurement of this wavelength, obtaining a reasonable value of 570 nm

41
Waves: part 3

In the arrangement shown in Figure 5.5, compared to Youngs original set-up, we have done away
with the first screen containing the single slit. The purpose of this was to provide a source of
coherent waves incident upon the two slits (coherent means in-phase and with the same frequency).
The wave sources must have a constant phase difference in order to produce a meaningful
interference pattern. Nowadays the experiment is simplified by the use of a laser source, which
provides a convenient source of monochromatic, coherent waves in a collimated beam that
illuminates the double slits directly.

Referring to Figure 5.5; light passing through slits S 1 and S 2 is in phase and so rays arriving at point Q
on the central axis, midway between the slits, must be in phase because the light from each slit has
travelled the same distance. Hence, constructive interference occurs and the centre of the
interference pattern is a bright fringe.

Now let us consider an arbitrary point P on the screen, separated by an angle from the central axis.
Light arriving at P from S2 has travelled further than light coming from S1. The resulting intensity at P
when the two rays combine depends on the phase difference and, for light from a coherent source,
the phase difference depends only on the path difference between the two rays. If the path
difference happens to be exactly one wavelength (or 2, or 3, or 4 etc) then the rays will again be
in phase and constructive interference will produce a bright fringe at P. But if the path difference
happens to be half a wavelength (or 3/2, 5/2, 7/2 etc) then the rays will be completely out of
phase and fully destructive interference will give a dark fringe at P. If the path difference is
somewhere in between, the phase difference will be intermediate, the interference will be
intermediate, and the resulting intensity on the screen will be somewhere in-between a maximum
and a minimum.

So the location of the bright (or dark) fringes depends on the path difference between rays from the
two slits, which we will call L. This path difference is illustrated in Figure 5.5(b). Note that we can
treat the rays from S1 and S2 to P as being parallel because the distance from the slits to the screen
(typically > 1 m) is very much greater than the distance D between the slits (typically < 1 mm).

It is simple to see that the path length difference is given by

L =D sin

For a bright fringe we saw that L must be either zero or an integer number of wavelengths. This
means the condition for a bright fringe (either side of the central maximum) is

n
=sin = for n 1, 2,3,... 5.2
D

For a dark fringe, L must be an odd multiple of half wavelengths. So the condition for a dark fringe
(either side of the central maximum) is

1
n +
sin =
2
= for n 1, 2,3,... 5.3
D

42
Waves: part 3

The analysis above represents an ideal double-slit experiment. We have implicitly assumed that the
two slits have a slit width much less than the wavelength of the light, so that the diffracted
wavefronts from each individual slit spread out uniformly over a wide angular spread, as from a
point source.

We will not do it here, but it can be shown (using nothing more than classical wave theory, as we
have done for the double slits) that diffraction at a single slit also produces an interference pattern
with a central, bright maximum and alternating dark and light fringes on either side 14. If the slit
width is very much narrower than the wavelength of the light the width of the central maximum is
Copyright2011 John Wiley & Sons, Inc. All rights reserved. effectively spread out over the entire
width of the viewing screen, and
beyond. So in this case, the
diffraction from each individual slit
of a double-slit experiment would
not affect the appearance of the
final interference pattern.

In reality though, the slits of a


double-slit experiment have a width
that is not negligible. This means
that the diffraction pattern from
each individual slit is much narrower
and the central maximum, plus one
or more minima on either side of it,
may fall within the width of the
viewing screen. Hence the
interference pattern that is actually
seen is a combination of the
diffraction pattern from each
individual slit and the pattern caused
by the interference between the two
slits (see Figure 5.6c).

Figure 5.6(a) the intensity plot expected in a double-slit interference


experiment with slits of negligible width (b) the intensity plot for
diffraction by a single slit of width a (not negligible) (c) the intensity plot
for two slits of width a. The curve of (b) acts as an envelope, limiting the
intensity of the double-slit fringes in (a). Note that the first minima of the
diffraction pattern of (b) eliminate the double-slit fringes that would occur
near 12 in(c)

14
For a single slit of width a, we consider the slit to be a line of radiating sources and use geometrical
arguments to calculate the effect of the path differences between pairs of rays. The same reasoning about the
phase difference and constructive/destructive interference applies, as for the double slit case. The simplest
equation that can be derived looks like the condition for bright fringes in double-slit diffraction
i.e. sin = n/a. However, it is important to note that in the single-slit case, this is actually the condition for a
dark fringe

43
Waves: part 3

Our analysis of Youngs double-slit interference experiment, and the fact that the derived equations
(equations 5.2 and 5.3) yield results that match precisely the results of real experiments, provide
reliable proof of the wave nature of electromagnetic radiation.

5.5 The particle nature of electromagnetic radiation


In the previous section we conclusively demonstrated that electromagnetic radiation, in the form of
light, is a wave. In this section we will demonstrate equally conclusively that electromagnetic
radiation, in the form of light, consists of particles.

In 1905, Einstein published a paper 15 in which he proposed that electromagnetic radiation consisted
of particles called photons. Furthermore, he stated that the energy of photons is quantised. In other
words, the energy of a photon can only have a certain minimum value, given by the equation

E = hf 5.4

where h is the Planck constant 16 and has a value h = 6.63 1034 J s

This value hf, the energy of a single photon, represents the minimum amount of energy a light
wave can have; if the total energy is greater than this then it must be an integer multiple of this
value (and whatever the multiple is, that is the number of photons in the wave).

Why did Einstein propose these seemingly radical ideas? In order to answer this question we need to
look at another classic experiment; the demonstration of the photoelectric effect.

5.6 The photoelectric effect


The photoelectric effect was well-known by the 1900s, having been discovered and extensively
investigated in the letter part of the previous century. However, until the birth of quantum physics
no satisfactory explanation of the effect could be given.

If a beam of light of short enough wavelength is used to illuminate the clean surface of a metal, the
light will cause electrons to be ejected from the surface. This is the photoelectric effect. It is possible
to measure the kinetic energy of the electrons that are ejected (known as photoelectrons) from the
target and so deduce something about the nature of the energy transfer from the incident light to
the electron.

Electrons in a metal are held there by electric forces and so to escape from the material an electron
must gain a certain, minimum amount of energy. This amount of energy is characteristic of the

15
1905 was Einsteins annus mirabilis in which he published four papers (on Brownian motion, special relativity,
the photoelectric effect and mass-energy equivalence) that played a huge role in defining modern physics. It
was the paper on the photoelectric effect that won him the Nobel prize, although any of the others might have
done the same.
16
Max Planck introduced (somewhat reluctantly) the notion of quantisation in 1901. In order to correctly
model the well-known blackbody radiation curves he found that he had to quantise the energy of the
oscillating sources upon which his model was based. Although it solved the problem of the blackbody curves,
Planck (and others) thought for a long time that quantisation was essentially a mathematical trick that could
not reflect physical reality. Einstein then used the idea of quantisation to explain the photoelectric effect.

44
Waves: part 3

material and is called its work function . Typically, for common photoelectric materials the work
function is of the order of a few electron volts 17 (eV).

Figure 5.7 Schematic representation of the photoelectric effect, in which light of sufficiently short wavelength, incident
on the surface of a metal target, causes energetic electrons (photoelectrons) to be ejected.

If we now think about light as being described by classical wave theory, there are three simple (and
easily testable) predictions that we can make about the photoelectric effect.

1. There should be a time delay between the light first hitting the metal and the initial release of
photoelectrons

The classical description of light as a wave implies that the energy of the wave is uniformly
distributed across the wavefronts and is continually and smoothly transferred by the travelling
wave. If this is the case, it is quite simple to make an estimate of how long it would take for an
individual atom to absorb enough energy to eject an electron. For a relatively weak light source it
turns out that the expected delay can be anything from a few 10s of seconds up to a few minutes.

2. The resulting maximum kinetic energy Ek,max of the ejected electrons should be proportional to the
intensity of the incident light

The intensity of a light wave is proportional to the amplitude of the oscillating electric field. An
electron in the metal target should oscillate sinusoidally due to the oscillating electric force on it
from the waves electric field. If the amplitude of the electrons oscillation is big enough, the
electron can break free of the metal target and is ejected from the surface. So if we increase the
amplitude of the light wave we would expect more energy to be transferred to the electron, and for
the electron to escape with a higher kinetic energy. Note that many electrons that have been given
enough energy to overcome the work function of the metal will lose some of their kinetic energy
through collisions etc before they are ejected from the metal. Some electrons, however, will manage
to escape without losing any of their kinetic energy. These will have the maximum kinetic energy,
Ek,max, and we would expect this maximum value to be proportional to the intensity of the incident
light.

3. The photoelectric effect should occur for light of any frequency

We know that the energy of an electromagnetic wave is proportional to its frequency. So, the rate of
energy transfer from a wave of low frequency will be less than that for a high frequency wave and
we might expect it to take longer for an atom to absorb enough energy to eject a photoelectron. But
there is no reason why the photoelectron wont eventually be ejected. Furthermore, for light of low
frequency, we should (according to classical wave theory) just be able to increase the intensity of the

17
Remember, 1 eV = 1.6 1019 joules

45
Waves: part 3

light source and so transfer enough energy to eject a photoelectron. Therefore we would expect the
photoelectric effect to be possible for light of any frequency.

With these clear predictions in mind, let us look at two simple photoelectric experiments that can
put them to the test.

Experiment 1:

the frequency of the incident light is fixed


the intensity of the light is varied and Ek,max of the photoelectrons is measured.

The unexpected result of this first experiment is that all measurements show that for light of a given
frequency, Ek,max does not depend on the intensity of the light source. Whether the light source is
dazzlingly bright or so weak as to be hardly visible, the maximum kinetic energy of ejected
photoelectrons is always the same.

Experiment 2:

the frequency of the incident light is varied


Ek,max of the photoelectrons is measured as a function of frequency

The easiest way to show the results of this experiment is to plot a graph of the maximum kinetic
energy of the photoelectrons as a function of frequency (see Figure 5.8). Whilst the graph does show
that for a large part of the range, the maximum kinetic energy is (as we would expect) directly
proportional to the frequency, it is very clear that there is a threshold frequency (f t ) below which the
photoelectric effect does not happen. Below f t there are simply no photoelectrons ejected from the
target, regardless of how intense the light source is.

Figure 5.8 Graph of the maximum kinetic energy of a photoelectron ejected from a metal target as a function of the
frequency of the illuminating light. ft is the threshold frequency below which the photoelectric effect does not occur
(contrary to the predictions of classical wave theory).

46
Waves: part 3

One further point to note from both of these experiments is that there is no delay between
illuminating the sample and the appearance of the first photoelectron 18.

Clearly, these two experiments are enough to show that all three of the predictions made under the
assumption that light is a classical wave are incorrect. Trying to find an explanation for the results of
the photoelectric experiments was the reason for Einsteins proposals, given at the start of this
section. So lets see how these seemingly anomalous results are relatively easy to understand as
soon as we start thinking about light as being comprised of particles (photons) rather than as a wave.

The first thing to understand is that now, in the particle model, energy is transferred to an electron
as the result of a single photon being absorbed. Energy is not transferred continuously as successive
wavefronts pass through the target. The energy arrives in lumps or quanta, each quantum having a
value of hf joules (the energy of one photon). This explains why there is no delay before the first
photoelectrons are ejected. As soon as the first photon arrives (assuming that hf is greater than the
materials work function) it can transfer its energy, in a single absorption event, to an electron. It
also explains why the maximum kinetic energy of the photoelectrons is independent of the intensity
of the incident light. In the particle model, increasing the intensity of the light simply increases the
number of photons. It doesnt change the energy of individual photons because the frequency isnt
changed. Increasing the number of photons doesnt change the amount of energy transferred from a
single photon to an electron and so the kinetic energy of the photoelectrons is unchanged.

Now lets think about whats going on in the second photoelectric experiment. As we know, an
electron requires a minimum energy to escape from the metal target. If the frequency of the light
is such that hf is less than , then the electron cannot escape (as we have seen, increasing the
intensity doesnt help). Hence the existence of a threshold frequency; at frequencies below ft the
photon energy hf is less than . Above the threshold frequency, hf > and the electron has enough
energy to escape.

Above the threshold frequency any excess energy (hf ) that an electron acquires will appear as
the kinetic energy. Those electrons that escape without losing any of this kinetic energy will then
have the maximum possible kinetic energy (equal to hf ).

The results of the photoelectric experiments can be summed up in the photoelectric equation:

= hf
Ek ,max 5.5

So we have seen that an explanation of the photoelectric effect certainly requires quantum physics.
We have also seen that Einsteins explanation provides a compelling argument for the particle
nature of electromagnetic radiation.

18
At least, nothing greater than about 109 s

47
Waves: part 3

5.7 Light as a particle and a wave


We now have good experimental evidence (from Youngs slits and the photoelectric experiments)
that light exhibits wave-like and particle-like properties. When we consider (at the quantum level)
the point at which a beam of light transfers energy, we have to think in terms of a particle. The
energy is transferred in quanta, known as photons, each with energy hf. However, when the light
beam is simply travelling through space there is no way of discerning the passage of individual
photons. Rather, the light seems to travel as a wave and is subject to diffraction, interference,
refraction, reflection etc, just as we would expect from classical wave theory.

So, is there a way in which we can reconcile the idea of light as both particle and wave? Let us return
to the double slit interference experiment. Figure 5.9 shows the basic set-up and the wavefronts
from each slit, which overlap and interfere to produce the familiar pattern of alternating maxima
and minima in intensity.

Copyright2011 John Wiley & Sons, Inc. All rights reserved.


Figure 5.9 Double-slit interference experiment, with the addition of a movable photon detector positioned in the plane
of the screen. Also shown is a plot of the light intensity as a function of position on the screen, showing the alternating
intensity maxima and minima of the interference pattern.

Now, imagine a small photon detector D positioned at a point in the plane of the screen C. The
detector could be based on a photoelectric device, designed to give an audible click when it absorbs
a photon. We would find that the detector produces a series of randomly timed clicks, each click
indicating the transfer of energy from the light wave to the screen via a photon absorption. If the
detector were moved slowly up or down (as shown by the arrow in Figure 5.9), we would find that
the rate of clicks increases or decreases, with the maxima and minima in the click rate corresponding
exactly to the intensity maxima and minima of the interference fringes.

The implication of this thought experiment is that we cannot predict exactly when a photon will be
detected at a given point; photons are detected at individual points at random times. We do know,

48
Waves: part 3

however, that the relative probability of detecting a single photon at any point, in a specified time
interval, is proportional to the light intensity at that point 19.

This provides us with a probabilistic description of a light wave; when viewed at the quantum level
we should think of light not only as an electromagnetic wave but also a probability wave. When the
wave is travelling between the source and the screen it travels as a probability wave that, due to
diffraction and interference, produces a pattern of probability fringes on the screen 20.

The concept of probability waves provides one possible interpretation of the wave-particle duality
problem.

5.8 Matter waves and the de Broglie wavelength


Acceptance of the wave-particle duality of light gives rise to many interesting questions, just one of
which is if light, which behaves in many respects as a wave, can also have particle-like properties, is it
possible that entities that we normally think of in terms of particles (such as electrons, protons,
atoms, molecules etc) might also have wave-like properties?

This symmetrical argument is what led Louis de Broglie (in his 1924 PhD thesis) to propose that all
particles of matter with mass m and velocity v must have a real wave associated with them. Indeed,
he went further than this simple proposal and suggested that the wavelength of these so-called
matter waves would be related to the particles linear momentum p (where p = mv) and would be
given by the following equation:

h h
= = 5.6
p mv

It should be noted that this form of the de Broglie wavelength equation applies only to non-
relativistic particles 21 i.e. where the particle has velocity v << c

In order to understand the origins of this equation, it is necessary to understand how it came to be
accepted that photons (the mass-less quanta of light energy) do, in fact, have momentum.

It was Einstein (in 1916) who proposed that photons have linear momentum. He did so to explain
the results of scattering experiments 22 that appeared to show that when a photon is absorbed by an
atom, both energy and momentum are transferred (in fact, the results of such interactions resemble
in many respects classical collisions between two objects).

19
This means that the probability of detecting a photon is proportional to the square of the amplitude of the
wave. In the case of light, this is equal to the square of the amplitude of the electric vector of the wave
20
A fascinating variation of the double-slit experiment (first carried out by G. I. Taylor in 1909) is one in which
the light source used is so extremely weak that it only emits one photon at a time. Thinking in terms of
particles, one assumes that each photon passes through one slit or the other. Remarkably, although the
pattern builds up one point at a time (each point representing a single photon absorption), if the experiment is
allowed to run for long enough then a pattern of interference fringes can be seen to emerge on the screen. So
presumably the photons must be passing through both slits at the same time.
21
If a particle has a velocity that is a significant fraction of the speed of light then a relativistic form of the
equation must be used, that incorporates a Lorentz factor
22
So-called Compton scattering

49
Waves: part 3

Einstein had previously shown the mass-energy equivalence, summed up by the famous equation

E = mc 2 5.7

So, although we think of a photon as having zero mass, we can calculate an equivalent mass for a
photon of energy hf that is given by

E hf
=
m =
c2 c2

Substituting this into the equation for linear momentum (using v = c, because this is light)

hfc hf
= =
p mc =
c2 c

and, since c = f, we obtain

h
p= 5.8

So, the equation for the de Broglie wavelength of a matter wave comes from a simple
rearrangement of equation 5.8.

Experimental confirmation of the existence of matter waves came soon after de Broglies proposal 23
and the concept of matter waves, with wavelengths given by de Broglies equation, is now routinely
applied to matter from electrons to neutrons, to atoms, and even to complex molecules.
Furthermore, there are numerous examples of modern technologies that rely on an understanding
and utilisation of the wave-like properties of matter; for example, electron microscopy, electron and
neutron diffraction analysis, quantum confinement semiconductor devices etc

5.9 Black body emission


In Physics there is a special term for a (hypothetical) object that is a perfect absorber of
electromagnetic radiation. An ideal black body is an object that absorbs all e/m radiation falling on it,
regardless of the frequency of the radiation or its angle of incidence. The perfect black body is a
theoretical concept that cannot be realised in practice. However, the behaviour of many objects can
be approximated to ideal blackbody behaviour.

A common model of a blackbody object is a small hole in a large, insulated enclosure (see Figure
5.10). Radiation of any wavelength (i.e. any frequency), at any angle, will be absorbed by the hole 24.

As a perfect absorber, it might seem reasonable to assume that a black body would, indeed, appear
black to an observer (rather like a black hole of the astronomical variety 25). However, this is not
necessarily the case and the reason is due to another important property of black bodies.

23
The electron diffraction experiments of Davidson and Germer and of George Thompson
24
Even this is not quite a perfect absorber. Incident radiation with wavelengths longer than the diameter of
the hole will be partially reflected

50
Waves: part 3

In addition to being a perfect absorber, a black body in thermal equilibrium (in other words, at
constant temperature) is also an ideal emitter; it emits energy in the form of e/m radiation across
the full range of frequencies, with intensity that is independent of direction (its an isotropic emitter).
The radiation emitted is known as blackbody radiation and a plot of the relative intensity of the
blackbody radiation as a function of frequency is called the blackbody spectrum. A key feature of
the blackbody spectrum is that it is dependent only on the surface temperature of the object.

Returning to our model of a black body as a hole in an insulated


enclosure (Figure 5.10) we can think of the radiation that enters the
hole as being repeatedly absorbed and re-emitted by the walls of the
enclosure until, in accordance with the second law of
thermodynamics, the radiation and the walls of the enclosure reach
thermal equilibrium. The hole in the enclosure will allow some
radiation to escape 26 and this radiation will now approximate to
blackbody radiation 27, with a spectrum that is characteristic of the
Figure 5.10 An approximate ' temperature.
model of a black body, as a small
Blackbody spectra at a variety of temperatures, ranging from
hole in an insulated enclosure.
100 K to 10 000 K are shown in Figure 5.11. Note that because of the
very wide range of temperatures, the curves are plotted using logarithmic scales. It can be seen from
this diagram that at room temperature (300 K) the peak of the blackbody radiation occurs at a
wavelength of about 10 m (in the infrared, as we would expect).

Figure 5.11. Ideal blackbody spectra for surface temperatures from 100 K to 10 000 K. Note the use of a log scales.

25
A black hole can be considered as a good approximation to an ideal black body. It certainly acts as a perfect
absorber. Initially, the question remained as to how a black hole could also be considered to emit radiation.
This question was answered by Stephen Hawkings explanation of so-called Hawking radiation, which shows
that black holes do emit radiation and so can lose energy and, eventually, evaporate!
26
Since the hole is small, we assume that the escaping radiation has negligible effect on the thermal
equilibrium
27
The approximation breaks down for wavelengths that are greater than the dimensions of the hole.

51
Waves: part 3

It is also clear from these curves (and corresponds to everyday experience) that thermal radiation is
only really emitted with significant intensity in the visible region of the spectrum (i.e. is only easily
seen) at temperatures in excess of about 1000 K.

Blackbody spectra can be described mathematically by the Planck function. The Planck function is a
distribution curve that shows, for any given temperature T, the relative intensity of the blackbody
radiation at each wavelength (or frequency). The function itself is a complicated one and for our
purposes it is not necessary to show it in full; here we will simply refer to it using the following
notation:

B (T) where it is expressed as a function of wavelength, for given T

B v (T) where it is expressed as a function of frequency 28, for given T

What the Planck function actually calculates is the power radiated from any point within the black
body, into one solid angle unit, per metre squared of projected emitting surface, per spectral unit (m
for wavelength, or Hz for frequency). So the units of B (T) are W sr-1 m-3 and the units of B v (T) are
W sr-1 m-2 Hz-1. This value is generally referred to as the spectral radiance of the black body.

A slightly simpler form of the Planck function, both to visualise and to use, is called the spectral
radiant emittance. This tells us the power per square radiated from the surface of a blackbody, per
unit wavelength or frequency. It is given by 29

B (T) as a wavelength distribution, with units W m-3, or

B v (T) as a frequency distribution, with units W m-2 Hz-1

Whatever form it is in, the key result of the Planck function is that hotter objects have a distribution
peaking at shorter wavelengths (higher frequencies) and emit more power per square metre from
their surface (see Figure 5.11 and Figure 5.12)

Figure 5.12 shows black body curves for temperatures of 3000 K, 4000 K and 5000 K. Comparing the
peak wavelengths of the three distributions gives an illustration of an empirical relationship that was
discovered even before Planck had managed to derive an accurate mathematical description of the
blackbody curves. Wiens displacement law can be written as

maxT= b= constant 5.9

where b is Wiens displacement constant and is equal to 2.9 103 m K. For an object that can be
approximated to a black body, we can use Wiens law to obtain a value for the surface temperature
of an object, based only on the peak wavelength of the blackbody spectrum.

Also shown on the graph in Figure 5.12 (the black line) is the theoretical curve for a blackbody at
T = 5000 K, but calculated using the accepted theory that was subsequently replaced by Plancks

28
Note that in this case it is customary to use the symbol v (nu) for frequency, rather than f.
29
The pre-factor of comes as a result of adjusting the full expression to give the radiation emitted from a
surface i.e. into a hemisphere of solid angle 2, with an angular correction factor of 0.5 allowing for the fact
that the radiation within the blackbody is not all directed perpendicularly to the emitting surface

52
Waves: part 3

function. It illustrates what was known as the ultraviolet catastrophe. This was a way of expressing a
known flaw in the theory based purely on classical physics; it predicted that at the extreme of very
short wavelengths, the amount of radiated energy should increase in proportion to the square of the
frequency (and so approach infinity). It was as a result of trying to solve this problem that Planck had
to incorporate the revolutionary idea of the quantisation of energy into his function and in doing so
he kick-started the whole of modern (quantum) physics.

Figure 5.12. Blackbody curves for temperatures of 3000 K, 4000 K and 5000 K. Note that the axes are now plotted on
linear (not log) scales. Also shown is the (incorrect) prediction for the 5000 K curve made using purely classical theory. It
illustrates the so-called ultraviolet catastrophe.

Many real objects radiate like black bodies to a good approximation; for example, the element of an
electric heater, an incandescent light bulb, a candle flame, the human body, stars and planets. For all
such objects we have seen how Wiens displacement can be useful in relating surface temperature
to peak emission wavelength. We will now see how the blackbody curves of these objects can be
used to tell us about the power that they are radiating.

Given that the Planck function tells us the power per square metre per spectral unit (unit of
wavelength or unit of frequency), it should be clear that the area under the curve will be equal to
the total power emitted per square metre of the objects surface. In either case, whether using
B (T) or B v (T), the units will be watts per square metre 30. If we calculate this area between two
values of wavelength or frequency, we get the power emitted per m2 in that spectral range. If we
calculate the area under the whole curve, by integrating the Planck function, then we get the total

30
Area under wavelength distribution: units are W m-3 m, which gives W m-2. Area under a frequency
distribution: units are W m-2 Hz-1 Hz, which gives W m-2

53
Waves: part 3

power emitted per square metre by the object. This is referred to as the blackbody flux F and is
given by the Stefan-Boltzman law:

F = T 4 [W m-2] 5.10

Where is the Stefan constant and is equal to 5.67 10-8 W m-2 K-4

Another straightforward, but useful, result follows: if we have an object that approximates to a black
body with surface area A, we can calculate the total power radiated by the object by simply
multiplying the black body flux by the surface area:

total radiated power = F A = T 4 A [W]

5.10 Spectra
In section 5.2 we introduced the concept of the electromagnetic spectrum. In this context we used
the term spectrum just to mean the entire range of possible wavelengths (or frequencies 31) of e/m
radiation. In this section we will be looking in more detail at the subject of spectra and how they can
give us useful information about the object or substance from which they originate.

Individual objects have their own, characteristic spectra consisting of selected wavelengths. Each
spectrum contains information, not only about which wavelengths are present but also about the
relative intensity of those wavelengths. It is this knowledge that can give us an insight into the
nature and properties of the source of the spectrum. Furthermore, since it simply requires a
measurement of wavelength, it is information that can be obtained at very great (e.g. astronomical)
distances.

Figure 5.13 shows the three most common types of spectra, as seen in the visible part of the e/m
spectrum.

Figure 5.13. The three most common types of spectra (a) continuous (b) emission line and (c) absorption line

The first (Figure 5.13a) is the so-called continuous spectrum. This is a spectrum that contains all of
the wavelengths across a given spectral range, with no gaps or distinct lines. Continuous spectra

31
Note that spectra can be displayed in terms of wavelength or frequency (using the relation c = f to convert
from one to the other). Here, we will tend to refer to spectra in terms of wavelength.

54
Waves: part 3

most commonly originate from hot, dense solids, liquids or gases. The blackbody curves discussed in
the previous section are examples of continuous spectra and examples of typical sources would
include such things as incandescent lamps, stars, human bodies etc. As we saw in the previous
section, for any object that can be approximated to a black body we would expect e/m radiation to
be emitted over a continuous range of wavelengths, with a wavelength profile that is characteristic
of the temperature of the object.

The next two types of spectra are called emission line and absorption line spectra. Before
addressing the line aspect suggested by these names, we need to explain the general difference
between an emission and an absorption spectrum.

An emission spectrum results when a substance emits radiation at certain, discrete wavelengths (in
the case of a line spectrum) or over a range of wavelengths (in the case of a continuous spectrum).
An absorption spectrum is obtained when a substance is illuminated by the continuous spectrum of
a separate source. As the radiation passes through the substance in question, certain wavelengths
are selectively absorbed so that the transmitted radiation is no longer continuous but has gaps (dark
lines) corresponding to the absorbed wavelengths. Figure 5.14 shows a schematic representation of
how the different types of spectra may be produced.

Figure 5.14 A schematic illustration of how continuous, emission line and absorption line spectra are produced. Note
that the absorption line spectrum is a combination of the spectra of the illuminating source and the intervening
substance. The dark (missing) lines are, however, characteristic only of the substance through which the continuous
spectrum is passing.

The distinctive lines that are a feature of both emission and absorption line spectra (see Figure 5.13b
and c), indicate that these are examples of atomic spectra.

We know that electrons associated with individual atoms can only have certain allowed energy
levels (due to quantisation), and that they move between levels by the absorption or emission of
photons. Line spectra, therefore, typically originate from thin gases (which can be treated as a
collection of individual atoms or molecules) that emit or absorb photons whose wavelengths
correspond to the energy difference (E) between two atomic levels.

55
Waves: part 3

Recalling equation 5.4, the energy of a photon is given by hf. So, we can write an expression for the
wavelength of a spectral line in terms of the difference between the energy levels in a particular
atom:

hc hc
E = hf = = 5.11
E

A particular type of gas will always absorb or emit the same characteristic wavelengths and, since we
can calculate precisely the energy levels of different atoms, it is possible to use the emission lines to
identify a gas. Note that the simultaneous appearance of many wavelengths in a line spectrum is due
to the fact that, in any gas, there are many atoms emitting or absorbing different wavelengths at the
same time.

By way of example, Figure 5.15a shows the visible part of the emission line spectrum of hydrogen. It
is well known that the allowed energy levels for hydrogen are given (in electronvolts) by the
equation (in which n is an integer):

13.6
En = [eV] 5.12
n2

Figure 5.15(a) shows the Balmer series of hydrogen emission lines (b) shows the allowed energy levels for an electron in
a hydrogen atom

56
Waves: part 3

Using equation 5.12 we can produce a diagram such as that shown in Figure 5.15(b), which shows
the n = 1 to n = 6 energy levels for a hydrogen atom. As we have just seen, spectral lines arise from
photons that are emitted when electrons move between energy levels. The lines shown in Figure
5.15a, known as the Balmer series, originate from electron transitions to the n = 2 energy level.

For example, the blue line at 434 nm has a wavelength that (using equation 5.11) can be shown to
correspond to the energy difference between the n = 5 and n = 2 levels.

So it is clear that spectra, whether continuous or line, emission or absorption, can convey useful
information about the substance or object from which they originate. In order to interpret these
spectra, though, we need some means of measuring the characteristic wavelengths and their
relative intensities. This sort of measurement is the basis of an extensive range of analytical
techniques, known collectively as spectroscopy.

An essential part of analysing the wavelength profile of a spectrum is separating out the rays of
different wavelengths; in other words dispersing the beam of light so that different wavelengths are
displaced through different angles. This means that when they hit a screen, or some other detector,
the wavelength of a ray can be found simply by measuring its position along the screen (or detector).
This is the basis of a spectrometer. The dispersive element of a spectrometer can be a prism, in
which refraction through the material of the prism causes the differential angular separation of the
various wavelengths. More common, however, is to use a dispersive element known as a diffraction
grating which, as the name suggests, uses the phenomenon of diffraction to separate out
wavelengths.

In section 5.4 we saw how light of a particular wavelength, passing through a double slit apparatus,
would produce bright fringes at positions on the screen where

n
sin =
D

(where D was the separation between the slits and n = 0, 1, 2, 3... etc).

So, by measuring the position of the bright fringes on a screen, we can calculate the wavelength of
the light producing them.

A diffraction grating is similar to a double slit arrangement (the intensity peaks obey the same
formula) except there are many more slits 32 and they are separated by very small distances.
A typical diffraction grating may have as many as 1000+ parallel slits in the space of 1 mm. The
result of this is that the bright fringes are much narrower and brighter, allowing more accurate
measurement of .

32
The slits are usually formed by ruling very narrow lines onto a substrate that is either transparent or
reflective. The ruled lines then define a series of parallel windows that can diffract incident light by
transmission or reflection

57
Waves: part 3

Figure 5.16a shows a typical intensity profile for monochromatic light incident on a diffraction
grating. Figure 5.16b shows the actual appearance of the spectrum when projected onto a screen
(assuming that the light is red).

Copyright2011 John Wiley & Sons, Inc. All rights reserved.

Figure 5.16 (a) Typical intensity profile obtained when monochromatic (red) light is dispersed by a diffraction grating (b)
shows how the spectrum would appear if it was projected onto a screen or detector

The fact that gratings produce narrow, well-separated peaks also means that if a beam of light
containing more than one wavelength is dispersed using a diffraction grating, it is still possible to
identify the individual peaks resulting from the different wavelengths, and so measure the various
wavelength components present in the spectrum.

58

You might also like