Lecture4: de Brogile matter waves, Group velocity and stationary phase.Wave for a free particle.

L4.1 de Brogile wavelength in different frames

L4.2 Galilean transformation of ordinary waves

L4.3 The frequency of a matter wave

L4.4 Group velocity and stationary phase approximation

L4.5 Motion of a wave-packet

L4.6 The wave for a free particle

L4.1 de Brogile wavelength in different frames

MITOCW | watch?v=8x94EgM2Mpg
BARTON
ZWIEBACH:
De Broglie, as we discussed last time, we spoke about waves. Matter waves. Because people
thought, anyway light is waves so the surprising thing that would be that matters are waves.
So a free particle with momentum p can be associated to a wave-- to a plane wave, in fact–
plane wave-- with wavelength lambda equals Planck’s constant over p. So this wave is what
eventually becomes a famous wave function. So de Broglie was writing the example or trying
to write the example of what eventually would become wave functions, and the equations for
this wave would become the Schrodinger equation. So really, this is a pillar of quantum
mechanics. You’re getting there when you talk about this wave.
So Schrodinger’s equation is a wave equation for these matter waves and this plane wave
eventually will become the wave function and there is a Schrodinger equation for it. So it’s a
wave of what? he was asking-- de Broglie had little idea what that wave was. When you have
waves, like electromagnetic waves, you have polarization, you have directional properties, the
electric field points in some direction, the wave is polarized. Is there a same property for the
wave function? The answer is yes. We’ll have to wait a little in 8.04 to see it, but it has to do
with spin. When the particles have spin, there are directional properties of the wave and
typically, you use several wave functions that correspond to directional components of this
wave.
So photons are spin 1 particle electrons or spin 1/2 particles, so there will be directional
properties to it. But to begin with, let’s consider cases where this directional properties don’t
matter so much, and for the case of electrons, if the electrons have small velocities or they are
inside small magnetic fields where some of these properties of the spin is important, we can
ignore that and work with a wave function that will be a complex number. So it will be a wave
function-- we’ll denote it by the letter psi, capitol psi-- that depends on position and time, and
that’s the wave function. And to begin with, simplicity will be one of them, and it’s a complex
number. And it’s just one wave function.
And the obvious questions about this wave function are, is it measurable and what it’s
meaning is? So is it measurable? And what is its meaning? But to understand some of that-- in
fact, to get to realize that these waves are no ordinary waves, we’re going to think a little about
what it means to have a wave whose wavelength is inversely proportional to the momentum of
a particle. That’s certainly a strange statement and probably these are strange waves as you
will see. And by understanding that these are strange waves, we are ready to admit later on
that the interpretation could be somewhat surprising as well. And the nature of this number is,
again, a little strange as you will see.
So all of that will come by just looking a little more in detail at this formula of de Broglie and
asking a very simple question-- you have this particle moving with some momentum and I say,
OK, it has this much wavelength. How about the person? If one of you is moving relative to
me, like you usually do with Einstein, these observers that are boosted, but let’s just do nonrelativistic physics, what is called the Galilean transformation, in which there will be another
observer moving with constant velocity with respect to you and you and that other observer
compare the results on the momentum and the wavelength and see if you find a reasonable
agreement or things make sense.
So we’re going to try to think of p is h over lambda. And 2 pi’s are very useful sometimes. So
you put an h over 2 pi here and a 2 pi over lambda and you rewrite this in terms of quantities
that are a little more common-- one is h-bar and the other is called the wave number k. So
these are these two constants and this one is called the wave number. The 2 pi’s are all over
the place. If you have a wave with some frequency nu, there’s also a frequency omega, which
is 2 pi nu.
So we’re going to look at this wave, and it has some momentum and some wave number,
therefore it has some wavelength, and let’s see-- if we compare things between two different
frames, what do we find? So we’ll put the frame S and a frame S prime moving with some
velocity plus v in the x direction. So the setup is relatively common. We’ll have one frame here
that’s the S frame, and it’s the x-axis of the S frame. And the S prime frame coincided with the
S frame at time equals 0-- now it’s moving, so it’s now over here, it’s S prime. It has moved a
distance of vt-- it’s moving with velocity v and there’s t. And S prime has-- and x is x prime.
On this, we’re going to write a few things. We’re going to say we have a particle of mass m. It
has velocity v underbar, otherwise I’m going to get all my velocities confused. So this velocity v
is the velocity of the frame, v underbar is the velocity of the particle, and v underbar prime,
because the velocity depends on the frame of reference.
Similarly, it will have a momentum-- and all the things we’re doing are nonrelativistic, so
momentum p or p prime. Here is the particle. And that’s the position x prime with the particle.
And that’s a position x of the particle. So that’s our system. This particle is moving with some
velocity over here, and we’re going to compare these observations.
So it’s simple to write equations to relate the coordinates. So x prime, for example, is the value
of the corner at x of the particle minus the separation. So x minus vt. And I should say it here,
we’re assuming that t prime is equal to t, which is good nonrelativistically. It’s fairly accurate.
But that’s the exact Galilean answer-- when you talk about Galilean transformations and
Galilean physics, it’s very useful. Even in condensed matter physics, people write these days
lots of papers about Galilean physics, so when you have particles moving with low velocities,
it’s accurate enough, so might as well consider it.
And these are the two ways you transform coordinates, coordinates and time. So from this, we
can take a time derivative talking about the particle-- so we have dx prime and dt prime or t,
it’s your choice-- I guess I should put dt prime here, dx/dt minus v, which means that the
velocity v prime underbar is equal to v underbar minus little v. And that’s what you expect. The
difference of velocities is given by the subtraction of the velocity that the frame is moving. So if
this particular has some high velocity with respect to the lab frame with respect to this frame, it
will have a smaller velocity. So this sine seems right. And therefore, multiplying by m, you get
that p prime is equal to p minus mv.
So if you have that, we would have that lambda prime, the de Broglie wavelength measured by
either running person, is equal to h over p prime is equal to h over p minus mv, and it’s quite
different, quite substantially different from h over p, which is equal to the de Broglie
wavelength seen in the lab. So these two de Broglie wavelengths will differ very substantially.
If this would be a familiar type of wave-- like a sound wave that propagates in the medium, any
kind of wave that propagates in a medium, like a water wave or any wave of that type-- this
would simply not happen. In the case of those waves, you get a Doppler shift-- omega is
changed-- but the wavelength really doesn’t change. The wavelength is almost like something
you look at when you take a picture and whether you take a picture of the wave as you run or
you take a picture of the wave as you are sitting still, you’ll measure the same wavelength.
Let me convince you of that. It’s an opportunity to just do a little more formal transformations,
because these are going to be Galilean transformations, simple transformations. So our first
observation is that the de Broglie wavelength don’t agree, which pretty much, I think, intuitively
is saying that if you could just sort of see those waves and measure the distance between
peaks, they should agree, but they don’t, so there’s something very strange happening here.

L4.2 Galilean transformation of ordinary waves

MITOCW | watch?v=YdtHAIh-kas
BARTON
ZWIEBACH:
Do normal wave analysis to demonstrate that indeed these things should not quite happen. So
for that, so ordinary waves and Galilean transformations. So when you have a wave, as you’ve
probably have seen many times before, the key object in the wave is something called the
phaze of the wave. Phaze, the phaze. And it’s controlled by this quantity kx minus omega t. k
being the wave number, omega being the angular frequency and we spoke about. And the
wave may be sine of that phaze or cosine of that phaze or a linear combination of sines and
cosines, or E to this wave, any of those things could be your wave.
And whenever you have such a wave, what we say is that the phaze of this wave is a Galilean
invariant. Invariant. What it means is that two people looking at this wave, and they look at the
point on this wave, both people will agree on the value of the phaze, because basically, the
reality of the wave is based on the phaze, and if you have, for example, cosine of this phaze,
the place where this cosine is 0 is some of the phaze, and if the cosine is 0, the wave is 0, and
everybody should agree that the wave is 0 at that point. So if you have a place where the
wave has a maximum or a place where the wave is 0, this is an ordinary wave, everybody
would agree that at that place you have a maximum and in that place you have a 0.
So observers should agree on the value of this phaze. It’s going to be an invariant. And we
can rewrite this phaze in a perhaps more familiar way by factoring the k, and then you have x
minus omega over kt, and this is 2 pi over lambda, x minus-- this quantity is called the velocity
of the wave, and we’ll write it this way. And I’ll write in one last way-- 2 pi x over lambda minus
2 pi V over lambda t. And this quantity is omega and this quantity is k.
So this is our phaze. And we’ve said that it’s a Galilean invariant, so I will say that S should
see-- the observer S prime should see the same phaze-- phaze-- as S. So phi prime, the
phaze that S prime sees, must be equal to phi when referring to the same point. When
referring to the same point at the same time.
Let’s write this. So phi prime should be equal to phi. And phi, we’ve written there. 2 pi over
lambda x minus Vt. And this is so far so good, but we want to write it in terms of quantities that
S prime measures. So this x should be replaced by 2 pi over lambda x prime plus Vt minus Vt
like this. And I could even do more if I wish. I could put t prime here, because the t and t
primes are the same.
So phi prime, by the condition that these phazes agree, it’s given by this, which is by the
relation between the coordinates and times of the two frames, just this quantity. So we can
rewrite this as 2 pi over lambda x prime minus 2 pi over lambda V 1 minus little v over capital V
t prime. I think I got the algebra right. 2 pi over lambda, the sine-- yes, I grouped those two
terms and rewrote in that way.
So that is the phaze. And therefore, we look at this phaze and see, oh, whenever we have a
wave, we can read the wave number by looking at the factor multiplying x, and we can read
the frequency by looking at the factor multiplying t. So you can do the same thing in this case
and read, therefore, that omega prime, this whole quantity is this, omega prime. And this is k
prime, because they can respond to the frame as prime. So omega prime is equal to this 2 pi
V over lambda, which is omega, times 1 minus V over V. And k prime is equal to k or, what I
wanted to show, that lambda prime for a normal wave is equal to lambda for ordinary wave
moving in the medium.
So at this moment, one wonders, so what happened? What have we learned? Is that this wave
function is not like a sound wave. It’s not like a water wave. We’re doing everything nonrelativistic. But still, we’re seeing that you’re not expected to have agreement. That is, if
somebody looks at one wave function and you look at the same wave function, these two
people will not agree on the value of the wave function necessarily.
So the things that we conclude-- so the conclusions are that waves are surprising. So size are
not directly measurable-- measurable-- because if you had a quantity for which you could
measure, like a sound wave or a water wave, and you could measure aspects to it, they
should agree between different observables. So this is going to be something that is not
directly measurable-- not all of psi can be measured. Some of psi can be measured, and
you’re already heard the hints of that. Because we said any number that you multiply, you
cannot measure, and in the phase that you multiply, you cannot measure. So complex
numbers can’t be measured, you measure real numbers. So at the end of the day, these are
not directly measurable, per se.
The second thing is that they’re not Galilean invariant, and that sets the stage to that problem
6. You see, the fact that this phaze that controls these waves is Galilean invariant led you to
the quality of the wavelengths, but these wavelengths don’t do that. The de Broglie
wavelengths don’t transform as they would do for a Galilean invariant wave. Therefore, this
thing is not Galilean invariant, and what does that mean? That if you have two people and you
ask, what is the value of the wave function here at 103, the two observers might give you a
different complex number for the wave function. They will just not agree.
Not all is lost, because you will find how their measurements can be compared. That will be
the task of the problem. How-- if you have a wave function, how does your friend, that is
moving with some velocity, measure the wave function? What does this other person
measure?
So the end result, if you have a point here at some time t, the wave function psi of x and t is
not going to be the same as the wave function measured by the prime observer at x prime t
prime, so this point is the point x and t or x prime and t prime. These are two different labels
for the same point. You’re talking about the wave function at the same point at the same time.
You still don’t agree. These two people will not agree. If they agreed, this wave function would
have a simpler transformation law with a wavelength that this can serve.
So by simply discussing the Galilean properties of this wave, we’re led to know that the de
Broglie waves are not like normal matter waves that propagate in a medium or simple.

L4.3 The frequency of a matter wave

MITOCW | watch?v=3_qvO8bKGus
PROFESSOR: We’ve talked a lot about de Broglie saying that the wavelength is given by h over p. But we
have not said much yet about the frequency of the waves. So what is the frequency of those
matter waves?
So what is the frequency–
frequency–
of the matter waves.
So de Broglie did answer that same question. And the answer was obtained by analogy. We
have p equal h bar k.
And he said, well, just like the wavelength is determined by the momentum, we’ll have e equal
h bar omega. So the frequency-- so this equation is the one that now completes the story.
Omega is equal to e over h bar.
Fixes omega in terms of the energy. And we’re going to say a few things. In fact, this will be an
interesting digression into an important subject about waves that illustrates why this answer
makes a lot of sense.
And that’s, really, all you can do at this moment. This is a postulate of quantum mechanics.
That you do this thing, and with this, you get quantum mechanics. So the best thing we can do
is to explain why it makes sense in a number of ways, and then hope that the theory that you
built makes full sense.
So I want to remind you about velocities of waves. So if you have a wave now that it has k and
omega-- you have this thing. k minus omega, wave with a phase.
kx minus omega t. Then there is something called the phase velocity.
And it’s given by omega over k.
It’s the velocity in which the nodes and maxima of this plane wave move. So let’s see if this
makes some sense. Omega over k is the same thing as e over p.
We’re nonrelativistic, so let’s continue.
1/2 mv squared over mv. And this seems a little strange. 1/2–
v. So if I have a particle, you see, this is matter waves of energy e and momentum p. And e is
1/2 mv squared, the velocity of the particle. p is equal to mv. And now, somehow this wave
seems to be moving with half the speed of the particle. That looks pretty bad. What’s going
on?
Well, this is the usual story with waves. If the wave itself doesn’t-- a wave, a plane wave
carries no real information. It’s not the signal.
So many times when you try to represent the particle-- a little bit of information traveling–
representing it with a plane wave is actually quite wrong. You have to represent it with a wave
packet.
And therefore, this phase wave velocity being one half of the velocity of the particles seems to
just confirm the idea that, first, these waves are a little strange. And second, phase velocity is
not very meaningful physically.
The velocity that this more meaningful is v group velocity.
And it’s d omega dk evaluated at the value k that you’re using. k is a proxy for momentum. So
d omega dk may depend on k and omega. So if it’s d omega, dk is a function. Which value
should you use? Well, the value at the k that you’re propagating.
And this would be the same as d omega dk.
Is because of the constant separating-- the same as de vp.
But what is the kinetic energy in terms of the momentum? We wrote it last time. p squared
over 2m. That’s the kinetic energy expressed in terms of momentum. So this is d dp of p
squared over 2m.
Write p equal mv and you’ll recover the kinetic energy. And this is just-- because of the 2-- p
over m, which is the velocity of the particle. And this is the reason people believe de Broglie.
De Broglie made sense because the group velocity of this [? package ?] would be correct. And
that’s a very beautiful result. Actually, it’s true relativistically, as well.
If you put the energy and the momentum in relativity, this answer comes out exactly the same,
perfectly well. So to a large degree, since it also works for energy and momentum in relativity,
there was a motivation from relativity that I want to quote, although not elaborate on it too
much. So–
the motivation–
is that in special relativity–
relativity–
the components of the energy divided by c and the momentum form a 4-vector.
Just like position and time forms a 4-vector and transform nicely about-- with Lorenz
transformation-- e and p form a 4-vector. Nevertheless, when you consider phases–
like this, and you have x and t that form a 4-vector, the good behavior of phases also imply
that k and omega form a 4-vector. In fact, omega-- in relativity, omega over c and the k vector-
form a 4-vector. You see, in all the equations we’ve written-- and de Broglie-- de Broglie in
three dimensions or more, really, is p vector equal h bar k vector. And k is usually used for the
magnitude of this k vector.
So this is also 4-vector in special relativity. And therefore, vectors are things that transform
nicely. So it makes sense to say that one 4-vector is equal to another 4-vector. Because if it’s
true in one reference frame, it will be true in every reference frame.
So it’s almost irresistible to make them equal. And de Broglie, in some sense, said this is equal
to h bar. That’s de Broglie.
The interesting thing is that this is true relativistically. But actually, nonrelativistically, you can
make sense of this and set it equal to be the same things. And the phase velocities, group
velocities all makes sense.
Certainly, we’ve now said for [? minor ?] particles that e is equal to j bar omega. But another
statement would be that, yes, indeed, Einstein said that. That for photons, e was equal to h
bar omega or h nu.
And therefore, yes, whatever happens for photons happens for this matter waves. And so
also, so this is another argument. Group velocities is one. Special relativity is another reason.
And of, course photons.
Einstein–
said that e is equal h nu, which is equal to h bar omega.

L4.4 Group velocity and stationary phase approximation

MITOCW | watch?v=-UgQEHHXTRM
PROFESSOR: Velocity.
So we assume that we have an omega of k. That’s the assumption. There are waves in which,
if you give me k, the wavelength, I can tell you what is omega. And it may be as simple as
omega equal to kc, but it may be more complicated.
In fact, the different waves have different relations. In mechanics, omega would be
proportional to k squared. As you’ve seen, the energy is proportional to p squared. So omega
will be proportional to k squared.
So in general, you have an omega of k. And the group velocity is the velocity of a wave
packet-- packet–
constructed–
by superposition of waves.
Of waves.
All right. So let’s do this here. Let’s write a wave packet. Psi of x and t is going to be done by
superposing waves. And superposition means integrals, summing over waves of different
values of k. Each wave, I will construct it in a simple way with exponentials.
ikx minus army of kt.
And this whole thing, I will call the phase of the wave. Phi of k.
So that’s one wave. It could be sines or cosines, but exponentials are nicer. And we’ll do with
exponentials, in this case. But you superimpose them. And each one may be superimposed
with a different amplitude.
So what does it mean? It means that there is a function, phi of k, here. And for different k’s,
this function may have different values. Indeed, the whole assumption of this construction is
based on the statement that phi of k peaks. So phi of k–
as a function of k is 0 almost everywhere, except a little bump around some frequency that
we’ll call k0.
Narrow peak.
That is our wave. And depending on how this phi of k looks, then we’ll get a different wave.
We’re going to try to identify how this packet moves in time. Now–
There is a quick way to see how it moves. And there is a way to prove how it moves. So let me
do, first, the quick way to see how it moves. It’s based on something called the principle of
stationary phase. I doubt it was said to you in [? A03 ?] in that way.
But it’s the most powerful wave to see this. And in many ways, the quickest and nicest way to
see. Takes a little bit of mathematical intuition, but it’s simple. And intuition is something that. I
think, you have.
If you’re integrating–
a function–
multiplied-- a function, f of x, multiplied by maybe sine of x. Well, you have f of x, then sine of
x. Sine is 1/2 the times positive, 1/2 the times negative. If you multiply these two functions,
you’re going to get the function that is 1/2 time positive and 1/2 the time negative.
And in fact, the integral will contribute almost nothing if this function is slowly varying. Because
if it’s slowly varying, the up peak and the down peak hasn’t changed much the function. And
they will cancel each other.
So the principle of stationary phase says that if you’re integrating a function times a wave, you
get almost no contribution, except in those places where the wave suddenly becomes of long
wavelength and the phase is stationary. Only when the wave doesn’t change much for a while,
and then it changes again.
In those regions, the function will give you some integral. So that’s the principle of stationary
phase. And I’ll say it here, I’ll write it here. Principal–
of stationary phase. We’re going to use that throughout the course.
Phase. And I’ll say the following. Since–
phi anyway only peaks around k0–
This is the principle of stationary phase applied to this integral. Since-- since only for k roughly
equal to k0.
The integral–
has a chance–
To be non-zero.
So here is what I’m saying. Look. The only place where this integral contributes-- it might as
well integrate from k0 minus a little delta to k0 plus a little delta. Because this whole thing
vanishes outside.
And if we’re going to integrate here, over this thing, it better be that this wave is not oscillating
like crazy. Because it’s going to cancel it out. So it better stop oscillating there in order to get
that contribution, or send in another way.
Only when the phase stops, you get a large contribution. So on the phase stops varying fast
with respect to k. So you need-- need-- that the phase becomes stationary–
with respect to k, which is the variable of integration–
at k0.
So around k0, better be that the phase doesn’t change quickly. And the slower it changes, the
better for your integral. You may get something. So if you want to figure out where you get the
most contribution, you get it for k around k0, of course.
But only if this thing it’s roughly stationary. So being roughly stationary will give the following
result.
The result is that the main contribution comes when the phase, phi of k, which is kx minus
omega of kt, satisfies the condition that it just doesn’t vary. You have 0 derivative at k0.
So the relative with respect to k is x minus d omega of k dk at k0 t must be 0. Stationary
phase. Function phase has a stationary point. Look what you get. It says there that you only
get a contribution if this is the case.
So the value of x, where you get a big bump in the integral, and the time, t, are related by this
relation. The hump in this packet will behave obeying this relation. So x is equal to d omega dk
at k0 t.
And it identifies the packet as moving with this velocity. x equal velocity times times. This is the
group velocity.
End of the answer by stationary phase. Very, extremely simple.

L4.5 Motion of a wave-packet

MITOCW | watch?v=i81OpQJIH8U
PROFESSOR: Let me demonstrate now with plain doing the integral that, really, the shape of this wave is
moving with that velocity. So in order to do that, I basically have to do the integral.
And of course, if it’s a general integral, I cannot do it. So I have to figure out enough about the
integral. So here it is. We have psi of x and t. It’s integral dk phi of k e to the ikx minus omega
of kt.
OK. It’s useful for us to look at this wave at time equals 0 so that we later compare it with the
result of the integral. So phi psi at time equals 0 is just dk phi of k e to the ikx.
Only thing you know is that phi has peaked around k0. You don’t know more than that. But
that’s psi of x and time equals 0. Let’s look at it later. So we have this thing here. And I cannot
do the integral unless I do some approximations.
And I will approximate omega. Omega of k, since we’re anyway going to integrate around k0,
let’s do a Taylor series. It’s omega of k0 plus k minus k0, the derivative of omega with respect
to k at k0 plus order k minus k0 squared.
So let’s–
do this here. So if I’ve expanded omega as a function of k, which is the only reasonable thing
to do. k’s near k0 are the only ones that contribute. So omega of k may be an arbitrary
function, but it has a Taylor expansion.
And certainly, you’ve noted that you get back derivative that somehow is part of the answer, so
that’s certainly a bonus. So now we have to plug this into the integral. And this requires a little
bit of vision because it suddenly seems it’s going to get very messy.
But if you look at it for a few seconds, you can see what’s going on. So psi of x and t, so far, dk
phi of k e to the ikx.
So far so good. I’ll split the exponential so as to have this thing separate. Let’s do this. e to the
minus i. I should put omega of k times t. So I’ll begin. Omega of k0 times t. That’s the first
factor.
e to the minus i, the second factor. k–
k d omega dk.
k0 times t. And the third factor is this one with the k0. e to the minus-- it should be e to the
plus. i k not d omega dk.
k0 t. Plus order–
higher up. So e to the negligible–
negligible until you need to figure out distortion of wave patterns. We’re going to see the wave
pattern move. If you want to see the distortion, you have to keep that [INAUDIBLE]. We’ll do
that in a week from now.
This is the integral. And then, you probably need to think a second. And you say, look. There’s
lots of things making it look like a difficult integral, but it’s not as difficult as it looks. First, I
would say, this factor–
doesn’t depend on k. It’s omega evaluated at k0. So this factor is just confusing. It’s not–
doesn’t belong in the integral. This factor, too.
k0 is not a function of k. d omega dk evaluated at k0 is not a function of k. So this is not really
in the integral. This is negligible. This is in the integral because it has a k. And this is in the
integral.
So let me put here, e to the minus i omega of k0 t e to the minus-- to the plus i k0 d omega dk-
at k0 t. Looks messy. Not bad. dk. And now I can put phi of k.
e to the i k x minus these two exponentials, d omega dk at k0 times t. And I ignore this. So far
so good.
For this kind of wave, we already get a very nice result because look at this thing.
This quantity can be written in terms of the wave function at time equals 0. It’s of the same
form at 5k integrated with ik and some number that you call x, which has been changed to this.
So to bring in this and to make it a little clearer-- and many times it’s useful. If you have a
complex number, it’s a little hard to see the bump. Because maybe the bump is in the real part
and not in the imaginary part, or in the imaginary part and not in the real part.
So take the absolute value, psi of x and t, absolute value. And now you say, ah, that’s why.
This is a pure phase. The absolute value of a pure phase is that.
So it’s just the absolute value of this one quantity, which is the absolute value of psi at x minus
d omega dk k0 t comma 0.
So look what you’ve proven.
The wave function-- the norm of the wave function-- or the wave. The new norm of the wave
at any time t looks like the wave looked at time equals 0 but just displaced a distance.
If there was a peak at x equals 0, at time equals 0. If at time equals 0, psi had a peak when x
is equal to zero, it will have a peak-- This function, which is the wave function at time equals 0,
will have a peak when this thing is 0, the argument.
And that corresponds to x equals to d omega dk times t, showing again that the wave has
moved to the right by d omega dk times t. So I’ve given two presentations, basically, of this
very important result about wave packets that we need to understand.

L4.6 The wave for a free particle

MITOCW | watch?v=T6TQHNXy5Wg
PROFESSOR: So what are we trying to do? We’re going to try to write a matter wave.
We have a particle with energy e and momentum p. e is equal to h bar omega. So you can get
the omega of the wave. And p is equal to h bar k. You can get the k of the wave.
So de Broglie has told you that’s the way to do it. That’s the p and the k. But what is the wave?
Really need the phase to-- how does the wave look like? So the thing is that I’m going to do an
argument based on superposition and very basic ideas of probability to get-- to find the shape
of the wave.
And look at this possibility. Suppose we have plane waves-- plane waves in the x plus
direction. A particle that is moving in the plus x direction. No need to be more general yet.
So what could the wave be? Well, the wave could be sine of kx minus omega t. Maybe that’s
the de Broglie wave.
Or maybe the de Broglie wave is cosine of kx minus omega t.
But maybe it’s neither one of them. Maybe it is an e to ikx minus i omega t. These things move
to the right. The minus sign is there. So with an always an e to the minus i omega t. Or maybe
it’s the other way around. It’s e to the minus ikx plus i omega t. So always an e to the i omega
t.
And then you have to change the sine of the first term in order to get a wave that is moving
that way.
And now you say, how am I ever going to know which one is it? Maybe it’s all of them, a
couple of them, none of them.
That’s we’re going to try to understand.
So the argument is going to be based on superposition and just the rough idea that somehow
this has to do with the existence of particles having a wave.
And it’s very strange. In some sense, it’s very surprising. To me, it was very surprising, this
argument, when I first saw it. Because it almost seems that there’s no way you’re going to be
able to decide. These are all waves, so what difference can it make?
But you can decide.
So my first argument is going to be, it’s all going to be based on superposition. Use
superposition-- --position. Plus a vague notion of probability-- --bility.
So I’m going to try to produce with these waves a state of a particle that has equal probability
to be moving to the right or to the left.
I’m going to try to build a wave that has equal probability of doing this thing. So in case 1, I
would have to put a sine of kx minus omega t. That’s your wave that is moving to the right.
I have to change one sine here. Plus sine of kx.
Say, plus omega t.
And that would be a wave that moves to the right. Just clearly, this is the wave that moves to
the left. And roughly speaking, by having equal coefficients here, I get the sense that this
would be the only way I could produce a wave that has equal probability to move to the left
and a particle that moves to the right.
On the other hand, if I expand this you get twice sine of kx cosine omega t.
The fact is that this is not acceptable. Why it’s not acceptable? Because this wave function
vanishes for all x at t omega t equal to pi over 2, 3pi over 2, 5pi over 2. At all those times, the
wave is identically 0. The particle has disappeared. No probability of a particle. That’s pretty
bad. That can’t be right.
And suddenly, you’ve proven something very surprising. This sort of wave just can’t be a
matter particle. Again, in the way we’re trying to think of probabilities.
Same argument for 2 for same reason. 2-- So this is no good. No good.
The wave function cannot vanish everywhere at any time. If it vanished everywhere, you have
no particle. You have nothing. With 2, you can do the same thing. You have a cosine plus
another cosine.
Cosine omega t minus-- kx minus omega t plus cosine of kx plus omega t. That would be 2
cosine kx cosine omega t. It has the same problems.
Let’s do case number 3. Case number 3 is based on the philosophy that the wave that we
have-- e to the ikx minus i omega t always has an e to the minus i omega t as a phase. So to
get a wave that moves in the opposite direction, we have to do minus ikx minus i omega t.
Because I cannot change that phase. Always this [INAUDIBLE].
Now, in this case, we can factor the time dependence. You have e to the ikx e to the minus ikx
e to the minus i omega t.
And be left with 2 cosine kx e to the minus i omega t.
But that’s not bad. This way function never vanishes all over space. Because this is now a
phase, and this phase is always non-zero. The e to the minus i omega t is never 0. The
exponential of something is never 0, unless that something is real and negative.
And a phase is never 0. So this function never vanishes for all x-- vanishes for all x. So it can
vanish at some point for all time. But those would be points where you don’t find the particle.
The function is nonzero everywhere else.
So this is good. Suddenly, this wave, for some reason, is much better behaved than these
things for superposition. Let’s do the other wave, the wave number 4.
And wave number 4 is also not problematic.
So case 4, you would do an e to the minus ikx e to the i omega t plus an e to the ikx e to the i
omega t. Always the same exponential. This is simply 2 cosine of kx e to the i omega t. And it’s
also good.
At least didn’t get in trouble. We cannot prove it is good at this point. We can only prove that
you are not getting in trouble. We are not capable of producing a contradiction, so far.
So actually, 3 and 4 are good. And the obvious question that would come now is whether you
can use both of them or either one at the same time. So the next claim is that both cannot be
true at the same time.
You cannot use both of them at the same time. So suppose 3 and 4 are good. Both 3 and 4–
and 4 are both good-- both right, even. Then remember that superimposing a state to itself
doesn’t change the state.
So you can superimpose 3 and 4-- e to the ikx minus i omega t. That’s 3. You can add to it 4,
which is e to the minus ikx–
minus omega t. I factor a sine. And that’s 4.
And that should still represent this same particle moving to the right. But this thing is twice
cosine of kx minus omega t. So it would mean that this represents a particle moving to the
right.
And we already know that if this represents a particle moving to the right, you get in trouble.
So now, we have to make a decision. We have to choose one of them. And it’s a matter of
convention to choose one of them, but happily, everybody has chosen the same one.
So we are led, finally, to our matter wave. We’re going to make a choice.
And here is the choice. Psi of x and t equal to the ikx minus i omega t. The energy part will
always have a minus sign.
Is the mother wave or wave function for a particle with p equal hk and e equal h bar omega
according to de Broglie. You want to do 3 dimensions, no problem. You put e to the i k vector,
x vector, minus i omega t. On p, in this case, is h bar k vector. So it’s a plane wave in 3
dimensions.
So that’s the beginning of quantum mechanics. You have finally found the wave corresponding
to a matter particle. And it will be a deductive process to figure out what equation it satisfies,
which will lead us to the Schrodinger equation.

  • 10
    点赞
  • 11
    收藏
    觉得还不错? 一键收藏
  • 0
    评论

“相关推荐”对你有帮助么?

  • 非常没帮助
  • 没帮助
  • 一般
  • 有帮助
  • 非常有帮助
提交
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