MIT8.04 Lecture3: Photoelectric effect, Compton scattering, and de Broglie wavelength.

L3.1 The photoelectric effect

L3.2 Units of h and Compton wavelength of particles

L3.3 Compton Scattering

L3.4 de Broglie’s proposal

L3.1 The photoelectric effect

MITOCW | watch?v=byEaU9ILHmw
PROFESSOR: Last time, we spoke about photons in the context of an interferometer. The Mach-Zehnder
interferometer. And we saw the very unusual properties of photons and interference, and how
relatively simple interference a effect can be used to produce a very surprising measurement.
Today we’re going to backtrack and go from the beginning, and think about photons as
physicists did 100 years ago, and how, by thinking about photons, they pretty much came up
with quantum mechanics. So we want to trace this back. And the best place to start, probably,
is with a photoelectric effect.
The photoelectric effect is an experiment done by Hertz in 1887, in which he irradiated plates.
That means shine light, high energy beams of light, on metal plate, and he found that
electrons were released. Those were called photo electrons. And therefore, you would get a
photoelectric current from those electrons. So this is the effect we want to discuss now, is the
photoelectric effect. And it’s Hertz, 1887.
So first a description. So polished metal plates irradiated may emit electrons. And these are
called photo electrons sometimes. Photo electrons is just an electron that was released due to
a photon. And therefore, we get a photoelectric current. OK so so far, so good. But what was
special about this experiment?
The first step that was special was that there was a critical frequency. If you would take a
sample and you would irradiate it with light, and you would begin with light with very low
frequency, nothing would happen. And all of a sudden after a certain frequency, boom. You
would get a current. So there is a threshold frequency, nu0, such that only for Nu greater than
nu0 there is a current. So no current for lower frequencies.
Now as it turned out, nu0 depends on the metal you’re irradiating. And even more, it’s a
complicated thing to calculate. It depends on the surface of the metal, so that’s why Hertz
apparently had to polish the metal. And this frequency, if the metal is irregular, may depend on
where you shine. So you’d better prepare the metal very nicely. And it may even depend on
the crystalline nature of the metal, because it’s a many body effect. You see, anticipating the
resolution, there is this piece of metal and there are a few free electrons running around. And
they run around among the crystalline structure of the metal. And removing them, it’s going to
take some energy, and that energy depends on the metal and the arrangement and all kinds
of things. So this nu0 depends on the metal and the configuration of atoms at the surface.
Third property was kind of interesting. The magnitude of the current was proportional to the
intensity of the light. Magnitude current is proportional to the light intensity. And perhaps the
last one and fourth, a rather important one, a very crucial property, is that you could observe
the energy of the photo electrons, and it seemed to be independent of the light intensity. So
energy of the photo electrons is independent of the intensity of light The number of photo
electrons would depend on the intensity of light, but not the energy of the photo electrons.
Now there is more to that, but I think it was not quite exactly noticed by Hertz. So Hertz
probably didn’t notice all these things. But the last one, that maybe we can put in brackets
here, is that the energy of the photo electrons E gamma-- oh, no-- E of the electrons increases
linearly with the frequency of the light.
So this photoelectric effect was not easy to understand if you thought of light as a wave. And
Einstein came up with an answer that he almost said what was going on, but didn’t quite use
the word. He said that light comes in bundles of energy. And in a beam, you have bundles of
energy quanta. Didn’t quite say light is a particle. He himself was a little non-committal, I think,
about this concept. But Einstein, in 1905, gives the natural explanation and says that light is
composed of quanta. He would have to wait until 1920s until the name photons came up,
given by a chemist, Lewis. So he called them quanta.
These are later photons, and I will use the name photons from the beginning. With energy, E
equal h nu, where nu is the frequency and h was Planck’s constant. Planck had already
introduced the constant in trying to fit the black body spectrum. The black body spectrum, the
intensity of light is a function of frequency in black body radiation, had a particular curve.
Planck tried to fit it and he realized he needed one constant and he called it h. That’s Planck’s
constant. And the same constant that Planck introduced, reappeared in Einstein’s proposition.
This is Planck’s constant.
So the picture that Einstein and others had was that you would have kind of a potential here
and plot energy over here, and maybe this is some distance. And you have a metal and there
is the electrons captured here. And here is zero energy. So they have negative energy, they’re
captured. And you need some amount of energy, w, which is called the work function, that
depends on the type of metal you have. And if you could supply that energy, w, to any one of
these electrons that are bound in this metal, they would come out and not be attracted any
more and would be able to fly free.
So it is like an escape velocity, you’re bound by the gravity of earth. You need something
velocity, some energy to shoot you out. Same thing here. So this w, or work function, is
defined as the energy needed to release an electron. And that work function is that thing that
depends on the metal you have and the structure and how well you’ve polished the surface.
So if this is true, then Einstein, if he was right with this property, there would be the following
statement that you could make. The energy of the electron, which is, roughly speaking, one
half mv squared, would be equal to the energy that the photon [INAUDIBLE], minus the work
function.
So you supply a photon. Some of the energy goes into the work function, but the rest of the
energy goes into giving free energy to these electrons. So you have the energy of the photon
minus w, which is what you need to just take it out with 0 velocity. And then the rest of the
energy of the photon would be transmitted as kinetic energy of the electron. So if this is true,
this would be h nu minus omega. And this was considered a prediction, because that
statement that the energy of the electron increases linearly with the frequency, was not quite
obvious to people. Experiments were not fine enough. Measuring the energy of the emitted
particles was not all that easy either.
So this was Einstein’s prediction. And the experimental confirmation took 10 years to come. It
was verified by Millikan in 1915. So Millikan, in 1915, measures the energy of the photo
electrons, verifies Einstein’s conjecture, and actually produces, by measuring so carefully the
energy of the photo electrons, produces a measurement of h, which is the best to that point.
And h is measured to better than 1%, so a very accurate measurement of h.
And perhaps you would say, OK, so this is all wonderful, now everybody believes in photons.
But that’s quite far from the truth. They didn’t believe in photons too much, because Maxwell
had been too successful. And Einstein himself knew that once you started believing in particles
like photons, you had this subject with this case of loss of determinism and waves that we
have sometimes as particles and things he didn’t like much. So people were quite reluctant to
believe in these things. It’s quite amazing. So it took a while still.
So let’s do a simple exercise to introduce some numbers here, and show you how to do some
very simple computation. So let me do an example. You shine UV light with lambda 290
nanometers on a metal with work function 4.05 EV. What is the energy, E of the photo
electrons, and what is their speed?
Now it is a goal of mine, and of the instructors in this course, that a calculation like that, you
should be able to do without turning on your iPhone and checking what the h bar is, and
getting a few constants, and what is an EV or all these things. Well nowadays, you can just
check Wolfram Alpha and they will give you the answer for this, in beautiful, beautiful
calculations. Just copy the question like that, pretty much, I think it will answer it for you. But
you should be able to do back of the envelope calculations, in which you estimate things
quickly. And with one significant digit, you don’t even need a calculator to do this.
So let’s see how one does this thing. So the first thing to do is to figure out what is the energy
of this photon. That’s the first problem. So the energy of a photon is h nu. But nu it’s not
lambda. So nu time lambda is c, so this is hc over lambda, where c is the speed of light. OK,
hc lambda, we could do it if we knew h.
I must say, I never remember what h is in normal units. Joule seconds, six point something, I
don’t quite remember it. So what do I do? I use h bar, which is h over 2 pi. So h is 2 pi h bar c
over lambda. And here is where you-- here is the first thing that maybe you want to remember
by heart. h bar c is a pretty nice number, it’s about 200mev times a fermi, If you want it more
precise, it’s 197.33, if you want to get five digits, but it’s pretty close to 200 mev fermi. And
what is a fermi? It’s 10 to the minus 15 meters.
OK, so with this number, I claim you can do pretty much all you want to do. So here you have
2pi times 200 mev times 10 to the minus 15 meters divided-- I’ll put 197 here-- divided by
lambda, which is 290 nanometers, which is 10 to the minus 9 meters. So 10 to the minus 9
and 10 to the minus 15 is 10 to the minus 6 up. And this is a million ev, which is 10 to 6 ev. So
all these meters cancel and there’s just an ev left. So this is 2 pi times 197 over 290 ev. And
you certainly could estimate this like 2 over 3 times 2 pi, which is 6. And that’s about four. And
if you want to do it more carefully, it comes out to 4.28 ev. And the nice thing is that the
answer comes in ev’s. And the work functions, everybody gives them in ev’s, so it’s a
convenient thing.
So at this moment, you have this electron energy of the photon being this. So energy of the
electron is energy of the photon minus the work function, which is 428 minus 405 ev, and it’s
0.23 ev. That’s a kinetic energy and that should be a non-relativistic electron because the rest
mass of an electron is about half a million ev. It is 511,000 ev. So this is fairly non-relativistic,
but how slow is it? Is it moving a centimeter per second? No, it’s moving pretty fast.
You can write this as one half mv squared. And then what do you do? You put one half m of
the electron c squared v squared over c squared. And this is one half of 511,000 ev times v
over c squared. So do the arithmetic, it’s 20.46 over that, square root and multiply it by the
speed of light. You can do this roughly in your head. And the velocity comes out to 284
kilometers per second, so it’s pretty fast.

L3.2 Units of h and Compton wavelength of particles

MITOCW | watch?v=S9RjSQro2e0
BARTON
ZWIEBACH:
Now that we’ve introduce h, h is a very important quantity in quantum mechanics. So let’s talk
a little more about h, its units, and we already put one number that I really wish you will
remember. Now let’s talk about the units of h and some other things you can do with h. So
units of h.
So if you have a quantity that appears for the first time and as it appears here, E equal h nu,
this is a good place to understand the units of h because the units of h would be units of
energy divided by units of frequency. And I put this square brackets to denote units.
Now what are the units of energy? We’re going to work with units that are characterized by M,
L and T-- mass, length, and time. So energy, you think kinetic energy and you say, MV
squared, so that’s a mass and velocity squared is L squared over T squared. So that’s units of
energy. Frequency is cycles per unit time. Cycles have a number of units, so it’s 1 over time
here. So then you say that it’s ML squared over T. So that’s the first answer and that’s a nice
answer, although it’s never quite that useful in this way, so we try to rearrange it.
And I will rearrange it the following way to think-- you see, it’s nice to think of what physical
quantity that we are familiar, hats units of h-bar. We know these units of h-bar are energy over
frequency, but that’s not a single physical quantity, so let’s look at it and separate this as L
times MLT. That’s the same thing. And then I see an interesting thing-- this is the units of
position, or length. Length or radius. Distance. And this has the units of momentum p.
Momentum. So this product has the units of angular momentum.
And perhaps that’s the most important quantity that has the units of h-bar. It’s something that
you should remember. Of h-bar has units of angular momentum, that’s why when people talk
about a particle of spin 1/2, they say the angular momentum is 1/2 of h-bar, and that has the
right units. So spin 1/2 particle-- 1/2 particle-- means that the magnitude of the intrinsic
angular momentum is 1/2 of h-bar. h or h-bar have the same units, they just differ by a 2 pi
that-- unfortunately, we have to be careful about that 2 pi, it affects numbers and some
formulas are nicer without the bar, some formulas are less nice.
So OK. So another thing that you could say is that this h allows you to construct all kinds of
new quantities. And that’s a nice thing to do. Whenever you have a new constant of nature
that comes up, and we have the speed of light, Planck’s constant, Newton’s constant-- seem
to be the three fundamental units of nature-- you can do some things. And you can look at this
quantity-- h is proportional to rp and get an inspiration. So you can think h has units of r times
p. And you can say, look-- if I have any particle with mass M, I can now associate a length to it.
I can invent a length associated to that particle.
And how do I do it? Well, this has units of length, so all I have to do is take h and divide by p.
Well, that will be one way to get the length where p is the momentum, and it will be called the
de Broglie wavelength.
But there is another way. Suppose this particle is just not moving and you have the momentum
and you say, wow, momentum is not moving, so what’s going on here? So think of it at rest
and then you say, well, you still can construct a length. You can put h and divide by the mass
times the velocity of light, why not? That’s a velocity, it is a constant of nature. So that way,
you associate a length to any particle of a given mass. You don’t have to tell me what is the
momentum. You can just know the mass and it has a length associated to it.
So it’s called the Compton-- Compton-- wavelength of a particle. And I want to make sure you
don’t confuse, it’s not the same as de Broglie wavelength that we will see later. It’s not the
same as the de Broglie wavelength. This is the Compton wavelength of the particle. And you
can say, all right, good, you give me a particle of some mass, I can tell you what a length
associated to it-- why would it be important? It will be important in two different ways-- through
an experiment and through a thought experiment, which I want to do right now.
You see, I could ask the following question-- I have this particle, has a mass M. I use the
speed of light, so with that mass M, I could associate out to this particle a rest energy. MC
squared. That’s the rest energy. And then I could ask, what is the wavelength of a photon that
has the same energy as the rest energy of this particle? So you translate the question into a
question of a length. Once you have some energy, there is a natural length, which is the
wavelength of a photon with that energy.
So let’s ask this question independently of what we did. So what is the wavelength–
wavelength-- of a photon whose energy is the rest mass-- rest mass-- of a particle? So the
rest mass is MC squared, and that’s the energy of this photon. And we know that energy of a
photon h nu or hC over lambda, and there, we can calculate the lambda. Lambda is hC over
MC squared, and no surprise, it gives us h over MC and that thing is the Compton wavelength.
So it’s sometimes called l-Compton of the particle of mass M.
So this is a way that you can think of this particle. You think of a particle, you have a Compton
wavelength, and that Compton wavelength is the wavelength of light that has that rest energy.
And that actually has experimental implications in high energy particle physics. Because if you
have an electron and it has a Compton wavelength, and you shine a photon that has that size,
that photon is carrying as much energy as the rest energy of the electron. And in particle
theory and quantum field theory, particles can be created and destroyed, so this photon
maybe can do some things and create more particles out of this electron, particle equation
could start. happening.
So it will be difficult to isolate a particle to a size smaller than its Compton wavelength,
because the photons could do such damage to the particle by creating new particles or doing
other things to it.
So for an electron, let’s calculate the Compton wavelength. So l-Compton of an electron would
be h over MeC, and you would do h-- you would do 2 pi h-bar C over and MeC squared. And
you’ve got 2 pi 197.33 MeV fermi, and here you would have 0.511 MeV. So this gives you
2,426 fermi, or about 2.426 picometers. Picometers is kind of a natural length. Picometer is 10
to the minus 12 meters. The Bohr radius is about 50 picometers, so that’s how big this thing is.
Is it still much bigger than the size of the nucleus? The nucleus is a few fermis. A single proton
is about a fermi big. And nucleus grow slowly, so you can have a big nucleus with 200 particles
maybe of a radius of 7 or 8 fermi. So this is still a lot bigger and it’s a very interesting quantity
that will show up very soon.

L3.3 Compton Scattering

MITOCW | watch?v=WR88_Vzfcx4
PROFESSOR: So we’re building this story. We had the photoelectric effect. But at this moment, Einstein, in
the same year that he was talking about general relativity, he came back to the photon. And
there there’s actually a quote of Einstein’s saying, his greatest discoveries, for sure, were
special relativity and general relativity. The photo-- he got the Nobel Prize for the photoelectric
effect, and he certainly helped invent the quantum theory and many important things in this
subject, but in retrospect, his greatest successes were that. But he may have not quite seen it
exactly that way. He wrote, that some stage of his whole life had been a difficult struggle
against the quantum, pulsed by some small happy interludes of some other discoveries. But
the quantum theory certainly made him very-- well, he was very suspicious about the truth, the
deep truth of the quantum theory.
So 1916, he is busy with general relativity, but then he’s more ready to admit that the photon is
a particle, because he adds that the photon now has momentum as well. So it’s a-- these
photons that were not called photons yet are quanta for energy. But now he adds it’s also for
momentum. So this already characterizes particles. You see, there is the relativistic relation,
well known by then, that e squared minus p squared c squared is equal to m squared c to the
fourth.
You might say, well, this is a little surprising. And If you don’t remember too much special
relativity, this may not quite be your favorite formula. Your favorite formulas might be that the
energy is mc squared divided by 1 minus v squared over c squared. And that the momentum
is the ordinary momentum again multiplied by this denominator like this. But these two
equations, with a little bit of algebra, yield this equation, which summarizes something about a
particle, that basically if you know the energy and the momentum of a particle, you know its
mass.
And it comes out from this. This is the relativistic version of similar equations in which you have
energy one half mv squared, momentum, mv, and then energy equal p squared over 2m, a
very important relation that you can check. For out of these two comes this one. And this is
nonrelativistic. So for photons, we will have particles of zero mass. Photons have zero mass,
m of the photon equals zero, and therefore e is equal to pc for a photon.
So we can look at what the photon momentum is, for example photon momentum. We can
treat it as some particle and the photon momentum would be e of the photon divided by c, or h
Nu of the photon divided by c. And it’s h over lambda of the photon of gamma. So this is a very
interesting relation between the momentum of the photon and the wavelength of photons.
So the idea that the photon is really a particle is starting to gather evidence, but people were
not convinced about it until Compton did his work. So the same Compton that we used this
length over there, he works on this problem and does the following. So is Compton scattering,
the name of the work. Compton Scattering.
So what is Compton scattering? It is x-rays shining on atoms again, but this time, these are
very energetic photons, energetic x-rays. X-rays can have anything from a hundred EV to
100kEV, 100,000 electron volts. And what are the energies, of binding energies of electrons in
atoms? 10 EV, 13 EV for hydrogen. So you’re talking about 100,000 EV coming in, so it’s
easily going to shake electrons and release them very easily.
So you’re going to have almost-- even though you’re shining on electrons that are bound to
atoms, it’s almost like shining light on free electrons if it’s x-rays. So a few things happening.
So this is photons scattering on electrons. Scattering on electrons that are virtually free. And
the first thing that happens is that there is a violation of what was called the classical Thomson
scattering, that you may have started in 802.
So the reason Compton scattering did the job and physicists finally admitted the photon was a
particle is that it made it look like particle collision of a photon with an electron, it could
calculate and measure and treat the photon as a particle, just like another particle like the
electron, and out came the right result. So the classical Thomson scattering was a photon as a
wave.
And what does that do? Well, you have a free electron, and here comes an electromagnetic
wave, E and B. And if it’s low frequency wave, low energy electron, this electric-- the magnetic
field does very little, because this election doesn’t move too fast and the velocity is being
small, the Lorentz force is very small. But the electric field shakes the electron. And as the
electron is being shaken, it’s accelerating, and therefore it radiates itself. And it radiates in a
pattern, so you get photons out.
And the pattern is the following. I’ll write the formula with this cross-section. We’ll maybe not
explain too much about what this cross-section means, it could be a nice thing for recitation.
This is the formula for the Thomson cross-section as a function of the angle between the
incident direction of the photon and the photon that emerges. So you detect photons out and
this is the cross-section. What does it mean, cross-section? Well this has units. I will say, very
briefly, units of area. Area per solid angle, but solid angle has no units.
So if you imagine a little solid angle here and you multiply by this cross-section, it gives you
some area that represents the solid angle you’re looking at to see how many photons you get.
And the solid angle that you have multiplied here to give you an area, the area should be
thought of as the area that captures from the incoming beam the energy that is being sent into
this solid angle. So it represents an area, and an area represents an energy, because if you
have a beam coming in from a magnetic wave through a little area, some energy goes in.
So that area that you get is that area that extracts from the incident beam the energy that you
need to go in this solid angle direction. So basically, this is a plot of intensity of the radiation as
a function of angle. But the most important thing, not only-- this is not quite accurate when the
photon is of high energy. The thing that is pretty wrong about this is that the outgoing photon
or wave has the same frequency as the original wave.
So that’s a property of this scattering. The electron is being moved at the frequency of the
electric field, and therefore the frequency of the radiation is the same. And this is all classical.
But out comes, when you have a high energy, this thing is not accurate, and you have a
different result. So what did Compton find? Well, the first thing is a couple of observations. .
Treat the photon as a particle.
OK, so it has some energy and some momentum. The electron has some energy and
momentum. You should analyze the collision using energy and momentum conservation. So
before the collision, you have an incoming photon that has some energy and some
momentum, and you have an electron, maybe here. And then after a while, the electron flies
away in some direction, E minus. And the photon also, a photon prime of different frequency
flies away. It’s like a collision. You can do this calculation and maybe it could even be done in
recitation. It’s a relativistic calculation.
You were asked in first homework to show that the photon can not be absorbed by the
electron, and that uses the relativistic relations if you want to show you just can’t absorb it. It’s
not consistent with energy and momentum conservation. So it’s something you can try to
figure out. The other thing that should become obvious is that the photon is going to lose
some energy, because as it hits the electron, it gives the electron a kick. The electron now has
kinetic energy. Think of this in the lab. The electron was static, the photon was coming. After a
while, the electron has moved, it’s moving now with some velocity. The photon must have lost
some energy. So photon loses energy. And therefore, the final lambda must be bigger than
the initial lambda. Remember the shorter the wavelength, the more energetic the photon is.
So what is the difference? That’s the result of a calculation. It’s a nice calculation, all of you
should do it. It’s probably in some book, in many books. And it’s a nice exercise also for
recitation. Lambda final minus lambda initial. Or I’ll write it differently. Lambda final is equal to
lambda initial plus something that depends on the angle theta, in fact, has a one minus cosine
theta dependence. But here has to be something with units of length
And the only party you have here is the electron. And this electron has some length, which is
the quantum wavelength, which is very natural for a Compton scattering problem, of course.
And it’s here, h over mec. So the l Compton of the electron. And that’s the correct formula for
the loss of energy, or change in frequency. So the most you can get is if you don’t interact
when theta is equal to zero, the photon keeps going, doesn’t even kick the electron. And then
you get zero, the initial lambda is equal to the final lambda. But this can be as large as two, for
totally backwords photon emitted. So theta equals pi, cosine pi is minus 1, you get 2. At 90
degrees, you get the Compton shift. So it’s a very nice thing, you even know already what’s
happening here.
So let’s describe the experiment itself, of how it was done. So he used, Compton, the
experiment have a source of molybdenum x-rays that have lambda equals 0.0709
nanometers. So smaller than nanometers, it’s 70 picometers. And that corresponds to e
photon-- that’s pretty small, so it must be high energy-- and it’s 17.49 kEV. That’s very big
energy. And there was a carbon foil here. And you send the photons in this direction. And they
were observed at several degrees, but in particular, I’ll show you a plot of how it looked for
theta equal 90 degrees, so detector.
So source comes in, carbon is there, and what do you get? You get the following plot.
Intensity-- so you plot the intensity of the photons that you detect, as a function of the
wavelength of the photons that you get, because there’s supposed to be a wavelength, a shift
of wavelength. So it’s actually quite revealing, because you get something like this. A bump
and a bigger bump here. Something like that. Pretty surprising, I think, to first approximation.
And here is about-- the first bump happens to have about the same wavelength as the
incoming radiation, 0.0709. I should write it a little more to the left. 0.0709 nanometers over
here. And then there is another peak at lambda f, about 0.0731 nanometers.
And the question is, what is the interpretation? Why are there two peaks and what’s going on?
Anybody has any idea? Let me ask a simpler question. Which is the lambda that corresponds
to the prediction of the fact that the wavelength must change, the smaller one or the bigger
one? The bigger one. You certainly should loose energy, so the lambda f, the thing we were
expecting to see, presumably that thing, because we were expecting to see that at 90
degrees, the photons have this thing.
So we seem to observe this one. And let’s look at it in a little more detail. You have this 0.0731
nanometers and you have the original light was at 0.0709 nanometers. So the difference is
0.0022 nanometers, which is 10 to the minus 9, but it’s exactly, or pretty close, to this thing,
because a picometer is 10 to the minus 3 nanometers. So this is 0.0024 nanometers. So this
is pretty nice. Look, at 90 degrees, cosine theta is zero. So the difference between initial and
final wavelengths should be equal to the Compton wavelength with about 0.0024 nanometers.
And that’s about it pretty close. So this peak is all right. Should’ve been there. The other peak,
why is it there?

L3.4 de Broglie’s proposal

MITOCW | watch?v=dnuZx9fZHsU
PROFESSOR: This is Louis-- L-O-U-I-S d-e Broglie. And this is not hyphenated nor together. They are
separate. And the d is not capitalized apparently too. And it’s 1924, the photon as a particle is
clear, and the photon is also a wave.
And Louis de Broglie basically had a great insight in which he said that if this is supposed to be
a universal or a real basic physical property that photons are waves and particles, we knew
them as waves and now we know they’re particles. But if they are dualed with respect to each
other, both descriptions are in different regimes and in a sense, a particle at the end of the day
has wave attributes and particle attributes. Wave attributes because it interferes and is
described by waves. And particle attributes is because it has a definite amount of energy, it
comes in packets, they cannot be broken into other things-- this could be a more general
property.
And in a sense, you could say that de Broglie did a fundamental step almost as important as
Schrodinger when he claimed that all matter particles behave as waves as well. Not just the
photon, that’s one example, but everybody does. So associated to every modern particle,
there is a wave.
But that is quite interesting because in quantum mechanics, you have the photon and it’s a
particle, but it’s associated to a wave and if you are a little quick, you say, oh sure, the
electromagnetic wave, but no, in quantum mechanics, it’s the probability amplitude to be some
work. Those are the numbers we tracked in the mass and the interferometer, the probability to
be sampled. We didn’t track the waves or a single photon, the wave was a wave of probability
amplitude, something they didn’t know at all about yet at that time.
So de Broglie’s says just like the photons have properties of particles and properties of waves,
every particle has properties of waves as well and every wave has a property of particles. But
what is left unsaid here is yes, you have a wave, but a wave of what? And we’ve already told
you a little bit, the answer has to do with probability waves. So it’s very strange that the
fundamental equation for a wave that represents a particle is not an electric field or a sound
wave or this, it’s for all of them is a probability wave. Very, very surprising. But that’s what de
Broglie’s ideas led to.
So if you had a photon, you would say it’s a particle, and when you think of it as a particle, you
would say it’s a bundle of some energy and some momentum. And if you think of it as a wave,
you would say it has a frequency. And that’s a particle wave duality or in some sense, a
particle wave description of this object-- you have a particle and a wave at the same time.
When we have this, we have a particle wave duality. And de Broglie said that this is universal
for all particles. Universal. And it appeared the name of matter waves. These are the matter
waves that we’re going to try to discuss. These are the waves of something that are probability
amplitudes we’re going to try to discuss. So you could say wave of what? What. And that
comes later, but the answer is probability amplitudes, those complex numbers whose squares
are probabilities.
So just like we had for a photon, de Broglie’s idea was that we would associate to a particle a
wave that depends on the momentum. So remember, the Compton wavelength was a
universal-- for any particle, the Compton wavelength is just one number, but just for photons,
the wavelength depends on the momentum, so in general, it should be dependent on the
momentum. So we say that for a particle of momentum p, we associate a wave-- a plane
wave, in fact-- a plane wave, so we’re getting a little more technical, with of lambda equals h
over p, which is the de Broglie wavelength-- de Broglie wavelength.
So it’s a pretty daring statement. It was his PhD thesis and there was no experimental
evidence for it. It was a very natural conjecture-- we’ll discuss it a lot more next lecture-- but
there are very little evidence for it.
So experiments can a few years later, and people saw that you could interfere or diffract
electrons. They would behave, colliding into lattices like waves, and those are rather famous
experiments of Davisson and Germer. So particles, just like you do as an interference effect–
a two slit interference effect in which you have a screen, a slit and a screen, and you shine
photons and then you get an interference effect over here because of the wave nature of
photons, or in quantum mechanics, you would say, because there are probability amplitudes,
that are complex numbers that have to be interfered between the possibilities of the two paths,
because every photon goes through both paths at the same time, these experiments of
interference, or two slit interference, were done for electrons.
And then, eventually, they’ve been done for bigger and bigger particles, so that it’s not just
something that you do with elementary particles now. There’s experiments done about three
years ago-- I will put on the web site or on the notes some of these things so that you can see
them, but now you can throw in molecules here, molecules that have a weight of now 10,000
atomic mass units, like 10,000 protons, like hundreds of-- 430 atom molecules, and you can
get an interference pattern, so it’s pretty ridiculous. It’s almost like, you one day so you throw a
baseball and you’re going to see an interference pattern, but, you know, we’ve got to things
with about 10,000 hydrogen atoms and de Broglie wavelengths of 1 picometer, which are
pretty unbelievable.
So the experiments are done with those particles and in fact with electrons. People do those
experiments and they’re in very beautiful movies in which you see those electrons hitting on
the screen and then-- I’ll give you some links so you can find them as well-- and you see one
electron falls here and it gets detected and two electrons, three electrons, four electrons, five
electrons, six electrons-- by the time you get 10,000 electrons, you see lots of electrons here,
very well here, lots of electrons here, and the whole interference pattern is created by sending
one electron at that time in an experiment that takes several hours and it’s reduced to a movie
of about one minute.
So particles, big particles interfere, not just photons interfere. So those particles have some
waves, some matter waves discovered by de Broglie, and next lecture, we’re going to track the
story from de Broglie to the Schrodinger equation where the nature of the wave suddenly
becomes clear.

评论
添加红包

请填写红包祝福语或标题

红包个数最小为10个

红包金额最低5元

当前余额3.43前往充值 >
需支付:10.00
成就一亿技术人!
领取后你会自动成为博主和红包主的粉丝 规则
hope_wisdom
发出的红包
实付
使用余额支付
点击重新获取
扫码支付
钱包余额 0

抵扣说明:

1.余额是钱包充值的虚拟货币,按照1:1的比例进行支付金额的抵扣。
2.余额无法直接购买下载,可以购买VIP、付费专栏及课程。

余额充值