Lecture 15: Simple harmonic oscillator III. Scattering states and step potential.

L15.1 Number operator and commutators (15:49)

L15.2 Excited states of the harmonic oscillator (18:19)

L15.3 Creation and annihilation operators acting on energy eigenstates (21:03)

L15.4 Scattering states and the step potential (10:34)

L15.1 Number operator and commutators (15:49)

MITOCW | watch?v=kefsxztSX74
PROFESSOR: Last time we discussed the differential equation. I’ll be posting notes very soon. Probably this
afternoon, at some time. And last time, we solved the differential equation, we found the energy eigenstates, and
then turned into an algebraic analysis in which we factorized the Hamiltonian.
Which meant, essentially, that you could write the Hamiltonian-- up to an overall constant that
doesn’t complicate matters-- as the product of an a dagger a. And that was very useful to
show, for example, that any energy eigenstate would have to have energy greater than h
omega over 2. We call this a dagger a the number operator n, which is a Hermitian operator.
Recall that the dagger of a product of operators is the reverse order product of the daggered
operators. So the dagger of a dagger a is itself.
And then a was related to x and p, and so was a dagger. Recall that x and p are Hermitian.
And there are overall constants here that were wrote last time, but now they’re not that urgent.
And a and a dagger, the commutator is equal to one. That was very useful.
Finally, we also show that while the energy of any state would have to be greater than h
omega over 2, if you had a state that is killed by a hat, it would have the lowest allowed
energy-- which is h omega over 2. And, therefore, that is the ground state.
And we looked at this differential equation, and we found this Gaussian wave function. And it’s
a first order differential equation. And, therefore, it has just one solution. And, therefore, there
is just one ground state, and it’s a bound state. And, of course, you wouldn’t expect more than
one ground state, because there’s no degeneracies in the bound state spectrum of a onedimensional potential.
So we found one ground state was phi 0, and it’s killed by a-- which means that it’s killed by nhat, because a is to the right in n-hat. So the a finds phi 0 and just kills it.
Now, the other thing to note is that the Hamiltonian is really, pretty much, the same thing as
the number operator multiplied by something with units of energy. The number operator has
no units, because a and a dagger have no units. And that’s very useful. So it’s like a
dimensionless version of the energy.
And, certainly, if you have an eigenstate of h it must be an eigenstate of n. And the eigenvalue
of n-- if we call it capital n. Therefore, you can imagine this equation acting on an eigenstate–
which happens to be an eigenstate of n or of h. On the left-hand side you would read the
energy, and on the right-hand side you would read the eigenvalue of the n operator.
So that gives you a very nice simple expression. You see that the energy is the number plus a
1/2 multiplied by h-bar [INAUDIBLE].
So that’s pretty much the content of what we reached last time. And now we have to complete
the solution. And the plan for today is to complete the solution, familiarize ourselves with these
operators, learn how to work with a harmonic oscillator with them. And then we’ll leave the
harmonic oscillator for the time being-- let you do some exercises with it-- but turn to scattering
states. So the second part of today’s lecture we’ll be talking about scattering states.
OK. So when we look at this thing and you have a number operator-- which encodes the
Hamiltonian-- it’s a good idea to try to understand how it interacts with the other operators that
you have here. And a good question, whenever you have operators, is the commutator. So
you can ask, what is the commutator of n with a?
And this commutator is going to show up. But it’s basically that kind of thing. If you have a and
a dagger, you ask, what is the commutator? If you have n, you ask, what is the commutator
with the other thing?
So n with a would be the commutator of a dagger with a-- a like that. And sometimes I will not
write the hats to write things more quickly. Now, in this commutator, you can move the a out,
and you have a dagger a a. And a dagger a is minus 1. Because aa dagger is 1. So this is
minus a.
So that’s pretty nice. It’s simple. How about n with a dagger? Well, this would be a dagger a
with a dagger. A dagger with a dagger commute. So this a dagger can go out, and you get aa
dagger. And that’s 1, so you get a dagger.
So it’s a nice kind of computation relation. You would have commutation reserve x with p given
a constant. Now n with a gives a number times a. N commutated with a dagger gives a
number times a dagger. And those numbers are pretty significant, so I’ll write this again. N with
a is minus a. And n with a dagger is plus a dagger.
This is part of the reason-- as we will see soon-- that the name of a-- which we call destruction
operator, because it destroys the vacuum-- it’s sometimes called lowering operator, because it
comes with a negative sign here. And we’ll see a better reason for that name. A dagger is
sometimes called the creation operator or the raising operator, because it increases some
number, as you will see. And here it’s reflected by these plots.
But we need a little more than that. We need a little more commutators than this. So for
example, if I would have the commutator of a with a dagger to the k-- you can imagine this.
You have to become very used and very comfortable with these commutation relations. And
sometimes the only way to do that is to just do examples. So I’m doing this with a k here.
Maybe-- when you review this lecture-- you should do it with k equals 2 or with k equals 3, and
do it a few times. Until you’re comfortable with these things, and you know what identities
you’ve been using. If this was a little quick, then go more slowly and make absolutely sure you
know how to do those commutators.
In here, I’m going to say what happens. You have an a and you have to move it across a
string of a-hats. Now, moving an a cross an a-hat-- because of the commutator-- gives you a
factor of 1, but it destroys the a and the a-hat. As you move the a across the a-hats-- because
this is a with all the a-hats, here, minus the a-hats times the a there.
So if you could just move the a all the way across-- then you cancel with this. What you get is
what happens when you’re moving it to across. And you’re moving across a string of those. So
each time you try to move on a across an a-hat, you get this factor of 1 and you kill the a and
you kill one a dagger.
So this answer will not have an a, and it will have one less a dagger. So a dagger to the k,
minus 1. And then I would argue-- and you should do it more slowly-- that you have to go
across k of those. And each time you get a factor of 1, and you lose the a and the a dagger.
So at the end you get a k.
You should realize that this is not all that different from the kind of commutators you had. Like,
with x to the n. This was very similar-- it might be a good time to review how that was done-- in
which that pretty much gives you an x to the n minus 1, times a factor of n, because p is a
derivative. You could almost think of a as the derivative with respect to a dagger. And then this
commutator would be 1.
So this is true. And there is also-- if you want-- an a dagger with a to the n or a to the k. This
would give you-- if you had just one of them you would get a minus sign, because a dagger
with a is that. But the same thing holds, you’re going to get one less a-hat.
So a-hat to the k minus 1. A factor of k-- because k times you’re going to move an a dagger
across an a. And a minus because you’re getting a dagger commutator with a, as opposed to
a commutator with a dagger-- which is 1.
So these are two very nice and useful equations that you should be comfortable with. Now,
this implies that you can do more with an n operator. So n with a-hat to the k, this time will be
minus k a-hat to the k. It doesn’t change the number of a-hats, because you’re now making
commutators with a dagger a. So each time you have this commuted with one a-hat, the a
dagger and the a give you 1, but you have another a back. So the power is the same.
The sign comes from this sign.
AUDIENCE: Shouldn’t the n there have a hat?
PROFESSOR: Yes, it should have a hat. I’m sorry. Yes. And, similarly, n a-hat dagger to the k. This is k a-hat
dagger to the k. So what happened before, that n-hat operator leaves the a same but puts a
number-- leaves the a dagger the same and puts a number. Here, you see it happening again.
N with a collection-- with a string of a-hats-- gives you the same string, but the number. And
with a collection of a daggers gives you the same collection of a daggers with a number. And
the number happens to be the number of a’s or the number of a daggers. So that’s the reason
it’s called the number operator, because the eigenvalues are the number of creation operators
or the number of destruction operators.
I was a little glib by calling it the eigenvalue. But it almost looks like an eigenvalue equation,
which have an operator, another operator, and a number times the second operator. It is not
exactly an eigenvalue equation, though, because with eigenvalues you would just have this
acting on the second one.
But the fact that this case appear here are the reason these are number operators. So it was a
little quick for many of you. Some of you may have seen this before. It was a little slow, but the
important thing is after a couple of days from now, or by Friday, you find all this very
straightforward.

L15.2 Excited states of the harmonic oscillator (18:19)

MITOCW | watch?v=xmjvqbYvY9o
PROFESSOR: Suppose you define now, one state called phi 1 as a dagger acting on phi 0. You could not
define any interesting state with a acting on phi 0 because a kills phi 0, so you try phi 0 like
this. Now you could ask, OK, what energy does it have? Is it an energy eigenstate? Well it is
an energy eigenstate if it’s a number eigenstate. And we can see if it’s a number eigenstate by
acting with the number operator. So N phi 1 is equal to N a dagger phi 0. OK.
Here comes trick. Maybe it’s too much to even call it a trick, number one. This thing you look
at it and you say, I want to sort of simplify this, learn something about it. If this is supposed to
be an eigenstate of N hat, I have to make it happen somehow.
Now n hat kills phi 0. So if I would have a term a dagger times N hat near phi 0, it would be 0.
So I claim, and this is a step that I want you to be able to do also quickly, that I can replace this
by the commutator of these two operators. The product is replaced by the commutators. Why?
Aren’t products simpler than commutators? No. We have formulas for commutators. And
products are, in general, more complicated.
And why is this correct? And you say, well, it is correct because this has two terms. The term I
want minus a dagger N hat. But the term a dagger N hat is 0 because N hat kills phi 0. So I
can do that because this is N dagger a hat, which is what I had, minus a hat dagger N on phi
0. And this term is 0. So you would have put a 2 here or a 3 here, or any number even. But the
right one to put is the commutators. So that’s this.
And now this commutator is already known. That’s why we computed it. It’s just a dagger. So
this is a dagger phi 0, and that’s what we call phi 1. So N hat on phi 1 is phi 1. N hat has
eigenvalue 1 on phi 1. So N is equal to 1. That’s the eigenvalue. It is an eigenstate. It is an
energy eigenstate. In fact how much energy, E, is h bar omega times N, which is 1, plus 1/2,
which is 3/2 h bar omega?
And look what this is. This is the reason this is called a creation operator. Because by acting
on the ground state, what people sometimes call the vacuum, the lowest energy state, the
vacuum is called the lowest energy state, by acting on the vacuum you get a state. I mean,
you’ve created a state, therefore.
How is this concretely done? Remember you had phi 0 of x, what it is, and a dagger over there
is x minus ip over m omega. So this is x minus-- or minus h bar over m omega d dx. So you
can act on it. It may be a little messy. But that’s it. It’s a very closed form expression.
Now, phi 0 was defined, the ground state such that it’s a normalized state. This means the
integral of phi 0 multiplied with phi 0 over x is 1. That’s how we had the ground state.
You could ask, if I’ve defined phi 1 this way, is simply normalized? So I’ll try it. And now you
could say, oh, this is going to be a nightmare. Normalizing phi 0 is difficult. Now I have to act
with a dagger, which means act with x, take derivatives. It’s going to grow twice as big. Then
I’m going to have to square it and integrate it. It looks very bad.
The good thing is those with these a’s and a daggers, you have to compute anything, pretty
much. See how we do it. I want to know how much is phi 1 with phi 1. Is it 1? And it’s
normalized or not? Then I say, look, phi 1 is a dagger phi 0, a dagger phi 0. So far so good.
But I just know things about phi 0. So let’s clear up one phi 0. At least I can move the a dagger
as an a. So this is phi 0 a a dagger phi 0.
Can I finish the computation in this line? Yes, I think we can. Phi 0. a with a dagger, same
story as before. a would kill phi 0. So you can replace that by a commutator. Commutator of a
with a dagger phi 0.
But the commutator of a with a dagger is 1, so this is phi 0 phi 0 and it’s equal to 1. Yes, it is
properly normalized. So that’s the nice thing about these a’s and a daggers. Just start moving
them around. You have to get practice. Where should you move it? Where should you put it?
When you replay something by a commutator, when you don’t. It’s a matter of practice.
There’s no other way. You have to do a lot of these commutators to get a feeling of how they
work and what you’re supposed to do.
Let’s do another state. Let’s try to do phi 2. I’ll put a prime because I’m not sure this is going to
work out exactly right. And this time, I’ll put an a dagger a dagger on the vacuum. Two a
daggers, two creation operators on the vacuum.
And now I want to see if this is an energy eigenstate. Well, this is a dagger squared on the
vacuum. So let’s ask, is N hat-- is phi 2 prime an eigenstate of N hat? Well I would have N hat
on a dagger squared on phi 0.
Again, by now you know, I should replace this by a commutator because N hat kills the phi 0,
so N hat with a dagger squared phi 0. And that commutator has been done. It’s two times a
hat dagger squared, two times a dagger squared on phi 0, which is 2 phi 2. That’s what we call
the state phi 2 prime. I’m sorry.
So again, it is an energy eigenstate. Is it normalized? Well, let’s try it. Phi 2 prime phi 2 prime
is equal to a dagger a dagger. Let me not put the hats. I’m getting tired of them. a dagger a
dagger phi 0.
Now I move all of them. This a dagger becomes an a, the next a dagger becomes an a here.
So this is phi 0 a a a dagger a dagger phi 0. Wow, this looks a little more complicated.
Because we don’t want to calculate that thing, really. We definitely don’t want to start writing x
and p’s.
But, you know, you decide. Take it one at a time. This a is here and wants to act on this thing.
And then this other a will, but let’s just concentrate on the first a that wants to act.
a would kill phi 0, so we can replace this whole thing by a commutator. So this is phi 0. The
first a is still there, but the second, we’ll replace it by the commutator, this commutator.
I’ve replaced this product, the product of a times this thing, by the commutator of those two
operators. And then I say, oh look, you’ve done that. a with a dagger to the k is k a dagger k
minus 1. So I’ll write it here. This will be a factor of 2 phi 0 a. And this is supposed to be now a
dagger to one power less, so it’s just a dagger phi 0.
So this is supposed to be 2a dagger. So that’s what I did. And again, this a wants to act on phi
0 and it’s just blocked by a dagger, but you can replace it by a commutator. a with a dagger
phi 0. And this is therefore a 1, so this whole result is a 2.
So this phi 2 prime, yes, it is the next excited state. Two creation operators on the ground
state. Energy and eigenvalues too. You had N equal zero eigenvalue for the ground state 1 for
phi 1, 2 for phi 2 prime. But it’s not properly normalized. Well, if the normalization gives you 2,
then you should define phi 2 as 1 over the square root of 2 a dagger a dagger on phi 0. And
that’s proper.
So it’s time to go general. The n-th excited state, we claim is given by an a dagger a dagger, n
of them, acting on phi 0 with a coefficient 1 over square root of-- we might think it’s n, but it’s
actually, you can’t tell at this far-- this one is n factorial. That’s what you need.
That is the state. And what is the number of this state? What is the number eigenvalue on phi
n? Well, it is 1 over square root of n factorial. The number acting on the a daggers, the n of
them, phi 0. You can replace by the commutator, which then is 2 times already. So it’s N
commutator with a dagger to the little n phi 0 times 1 over square root of n.
And how much is this commutator? Over there. This is N times a dagger to the n phi 0. So
between these three factors, you’re still getting n phi to the n. So the number for this state is
little n. It is an energy eigenstate. The N eigenvalue is little n. And the energy is h bar omega.
The eigenvalue of N hat, which is little n plus 1/2. So it is the energy eigenstate of number little
n. This is the definition.
And the last thing you may want to check is the normalization. Let me almost check it here.
No, I will check it. Let’s say I think this is a full derivation. Phi n with phi n would be two factors
of those, so I would have 1 over n factorial a dagger a dagger, n of them on phi 0, a dagger a
dagger, n of them again on phi 0. So then that’s equal to 1 over n factorial phi 0 a a, lots of a’s,
n of them, n a daggers, phi 0, like that. That’s what it is.
We had to move all the a daggers that were acting on the left input of the integral, or the inner
product, all the way to the right. And that’s it.
So now comes this step. And I think you can see why it’s working. Think of moving the first a
all the way here. Well, you can replace the first a with a commutator. But that a with lots of a
daggers, with n a daggers, would give you a factor of n, with n a daggers will give you a factor
of n times one a dagger less.
So to move the first a, there are n a daggers and you get one factor of n from this a. But for
the next a, there’s now n minus 1 a daggers, so this time you get a factor of n minus 1 when
you move it. From the next one, there’s going to be n minus 2 a daggers, so n minus 2. All of
them all the way up to one, cancels this n factorial, and that’s equal to 1.

L15.3 Creation and annihilation operators acting on energy eigenstates (21:03)

MITOCW | watch?v=BRFekCz4XQY
PROFESSOR: Important thing to do is to just try to understand one more thing. The creation and annihilation
operators-- what do they do to those states? You see, a creation operator will I add one more
a dagger, so somehow must change phi n into phi n plus 1.
A destruction operator with an a will kill one of these factors, and therefore it will give you a
state with lower number of phi n minus 1. And we would like to know the precise relations. So
look at this. Let’s do with an A on phi n. And we know it should be roughly phi n minus 1. This
is one destruction operator, but we can do it.
Look-- this is 1 over square root of n. A times a dagger to the n phi 0. A with a dagger to the n
phi 0, we can replace by a commutator again. Commutator of a, a dagger to the n phi 0. This
is 1 over square root of n factorial, and here we get a factor of n times a dagger to the n minus
1 phi 0.
You know, it’s all a matter of those commutators we on the left blackboard. But this state-- by
definition, we have n square root of n factorial. That’s state, by definition, is phi n minus 1
times square root of n minus 1 factorial. See, by looking at this definition and saying, suppose I
have n minus 1, n minus 1, this is phi n minus 1. So n minus 1 a daggers on phi 0 is n minus 1
factorial square root multiplied phi n minus 1.
And now we can simplify this-- square root of n factorial and square root of n minus 1 factorial
gives you just a factor of square root of n that with this n here, this square root of n, phi n
minus 1. Sop there we go-- here is the first relation. A is really a lowering operator. It gives you
an eigenstate 1 less energy, but it gives it with a factor of square root of n, that if you care
about normalizations, you better keep it.
That factor is there because the overall normalization of this equation was designed to make
the states normalized. Similarly, we can do the other operation, which is what is a dagger
acting on phi n. This would be 1 over square root of n factorial, but this time a dagger to the n
plus 1 on phi n, phi 0.
Because you had already a dagger to the n, and you put one more a dagger. But this thing is
equal to what? This is equal to square root of n plus 1 factorial times phi n plus 1. From the
definition-- I hope you’re not getting dizzy. Lots of factors here. But now you see that the n part
of the factorial cancels, and you get that a hat dagger phi n is equal to square root of n plus 1
phi n plus 1.
OK let’s do an application. Suppose somebody asks you to calculate example. The
expectation value of the operator x on phi n, the expectation value of p on phi n. How much
are they?
OK. This, of course, in conventional language, at first sight looks prohibitive. I would have to
get those phi n [? in ?] some Hermit polynomial hn, for which I don’t know the closed form
expression. It’s a very large polynomial, jumps 2 by 2, there is exponentials, I will have to do
an integral.
That’s something that we don’t want to do. So how can we do it without doing integrals? Well,
this one’s-- actually, you can do without doing anything. You don’t have to do integrals, you
don’t have to calculate. The answers are kind of obvious, if you think about it the right way.
That’s not the obvious part, to think about the right way.
But here it is. Look, what is this integral? This is the integral of x times phi and of x, those are
real, quantity squared. And the phi n’s are either even or odd, but the fight n squared are
even. And x is odd. So this integral should be 0, and we shouldn’t even bother. That’s it.
Momentum. Expectation value of the momentum. All these are stationary states. Cannot have
momentum. If it had momentum, here is the harmonic oscillator, here is the wave function. If it
has momentum, half an hour later it’s here. It’s impossible. This thing cannot have momentum.
This must be 0 as well.
OK. Now this one is something you actually proved in the first test-- the expectation value of
the momentum operator on a bound state with a real wave function was 0. And you did it by
integration-- but in fact you proved it in two ways, in momentum space, in coordinate space, is
[? back ?] the same thing.
OK. So these ones were too easy. So let’s try to see if we can find something more difficult to
do. Well, actually, before doing that I will do them anyway with this notation. So what would I
have here? I would have phi n x phi n. And I say, oh, I don’t know how to do things with x.
That’s a terrible thing. I would have to do integrals.
But then you say, no. X-- I can write in terms of a and a daggers. And a and a daggers you
know how to manipulate. So this is a formula we wrote last time, and it’s that x is equal to
square root of h over 2m omega, a plus a dagger.
So x is proportional to a plus a dagger. So here is a square root of h, 2m omega, phi n, a plus
a dagger, on phi n. Now, this is 0, and why is that? Because this term is a acting on phi n.
Well, we have it there-- is square root of n, phi n minus 1. And a dagger acting on phi n is
square root of n plus 1, phi n plus 1.
But the overlap of phi n minus 1 with phi n is 0, because all these states with different energies
are orthogonal. It’s probably a property I should have written somewhere here. Which is-- not
only they’re well-normalized, but phi n phi m is delta nm.
If the numbers are different, it’s zero. And you see this is something intuitively clear. If you
wish, I’ll just say here-- these are 0, and this is 0 because the numbers are different. If you
have, for example, a phi 2 and phi 3, or let’s do the phi 3 and a phi 2, then you have roughly a
dagger, a dagger, a dagger, phi 0, a dagger, a dagger, phi 0. And then is equal to phi 0, three
a’s, and two a daggers. Correct?
And now you say, OK, this a is ready to kill what is on the right hand side. On the right side to
it. But it can’t because there are a daggers. But that a is going to kill at least one of the a
daggers. So an a kills an a dagger. The second a will kill the only a dagger that is left. And now
you have an a that is ready to go here, no obstacle whatsoever, and kills the phi 0, so this is
zero.
So each time there are some different number of eight daggers on the left input and the right
input, you get 0. If you have more a daggers on the right, then move them to the left, and now
you will have more a’s than a daggers and the same problem will happen. The only way to get
something to work is they are the same.
But this of course is guaranteed by our older theorems that the-- eigenstates, if Hermitian
operators with different eigenvalues are orthogonal. So this is nice to check things, but it’s not
something that you need to check.
All right. So now let’s say you want to calculate the the uncertainty of x in phi n. Well, the
uncertainty of x squared is the expectation value of x squared and phi n minus the expectation
value of x on phi n. On this already we know is 0, but now we have a computation worth our
tools.
Let’s calculate the expectation value of x squared in phi n. And if you had to do it with Hermit
polynomials, it’s essentially a whole days work. Maybe a little less if you started using
recursion relations and invent all kinds of things to do it. It’s a nightmare, this calculation.
But look how we do it here. We say, all right, this is phi n x hat squared phi n. But x hat
squared would be h bar over 2m omega, phi n times a plus a dagger time a plus a dagger phi
n. Now I must decide what to do, and one possibility is to try to be clever and do all kinds of
things.
Now, you could do several things here, and none is a lot better than the other. And all of them
take little time. You have to develop a strategy here, but this is sufficiently doable that we can
do it directly.
So what does it mean doing directly? Just multiply those operators. So you have phi n times a
a plus a dagger a dagger plus a a dagger plus a dagger a. All that on phi n. I just multiplied,
and now I try to think again. And I say oh, the first term is to annihilation operators acting on
phi n.
The first is go give you phi n minus 1. Second is going to give me a phi n minus 2 by the time it
acts. And a phi n minus 2 is orthogonal to a phi n. So this term cannot contribute. You know,
this term has two more a’s than this one. So as we just sort of illustrated, but it just doesn’t
match. These two terms acting on phi n would give you a phi n minus 2. And that’s orthogonal.
So this term cannot do anything. Nor can this, because both raise. So this will end up as phi n
plus two, for example, using that top property over there. Over there-- the box equation there.
If you have two a daggers acting on phi n, you will end up with a phi n plus 2. So this term also
doesn’t contribute. And that’s progress-- the calculation became half as difficult.
OK, that-- now we-- maybe it’s a little more interesting. But again, you should you should
refuse to do a [? long ?] computation. Whenever you’re looking at those things, you have the
temptation to calculate-- refuse that temptation. Look at things and let it become clear what’s
going on.
There are two terms here-- a dagger and a dagger a. That’s not even a commutator, it’s sort
of like an anti-commutator. That’s strange. But this, a dagger a, is familiar. That’s n. The
operator n. And we know the n eigenvalue, so this is going to be very easy. This is n hat.
The other one is not n hat, because it’s in the wrong order. N hat has a dagger a. But this
operator can be written as the commutator plus the thing in reverse order-- that equation we
had on top-- ab is equal to ab commutator plus ba. So this is equal to a a dagger plus a
dagger a. And this is 1. Plus another n hat.
So look-- when you have a and a dagger multiply, it’s either n hat or it’s 1 plus n hat. And
Therefore x squared expectation value has become h bar over 2mw phi m, and this whole
parenthesis is 1 plus 2 n hat phi n. And this is h bar over 2mw, phi n, and this is a number.
Because phi in is an n hat eigenstate. So it’s 1 plus two little n, phi n phi times 1 plus two [?
little ?] n.
And here is our final answer-- expectation value of x squared is equal to h bar over and m
omega, n plus 1/2 phi n. This is a fairly non-trivial computation. And that is, of course, because
the expectation value of x is equal to zero, is the uncertainty or x squared. It grows, the state is
bigger, as the quantum number n grows.
By a similar computation, you can calculate that you will do in the homework, the expectation
value of b squared and phi n, and then you will see how much is delta x, delta p, on the
[INAUDIBLE] on phi n. How much it is.

L15.4 Scattering states and the step potential (10:34)

MITOCW | watch?v=0ABYYJSvkVk
PROFESSOR: Scattering states are energy eigenstates that cannot be normalized. And when you say this
cannot be normalized, so what’s the use of them? They don’t represent particles.
Well, it’s like they e to the ipx over h bar, those infinite plane waves. Each one by itself cannot
be normalized, but you can conserve wave packets that are normalized. So the whole intuition
that you get with scattering states is based on the idea that we’re going to construct energy
eigenstates.
This time we cannot think of them as states of a particle. Bound states, yes. We can think of
them, they’re normalizable. But this energy eigenstates and bounded scattering states are not
states of one particle. So we definitely have to go back and produce wave packets.
But the intuition from those energy eigenstates is very valuable. So scattering states. And we
call them sometimes scattering states because they look like the process of scattering. This
will be non-normalizable energy eigenstates. And you’ve played a little with some of them. And
we’ll now study one case in detail. We’ll try a couple of cases between today and next lecture.
So the step potential. And the step potential is a potential that is 0 up to x equals 0. Here’s the
x-axis, and then suddenly there’s a step at v0. And here is the potential. But then the wave,
this is here, goes up. It’s a step.
And any energy eigenstates here has to be bigger. The energy has to be bigger than the
lowest point of the potential. You know that, you kind of have an energy that is like that less
because this would have decayed exponentially for infinite distance. It just, all over it would
have to decay exponentially. It’s impossible.
So all the energy states, eigenstates here, must have positive energy. So we have actually
qualitatively two possibilities. The energy may be less than v0, might be greater than v0. It
would look like you have to solve the problem two times. Happily, we’ll solve one, then let the
other happen by analytic continuation.
So here is the energy. I’ll take the energy greater than v0. But whatever is the energy, even if
it’s less than v0, the solution over here is going to be an exponential or a cosine and a sine, a
non-decaying function, and therefore can’t be normalized because it’s non-decaying forever
and ever. So it cannot be normalized.
So how do we write the solution for the energy eigenstate? It’s a psi of x. Well, I should write
two formulas: a formula for what’s happening on the left side, and the formula for what’s
happening on the right side.
Now I have a choice actually here. There’s two ways of visualizing this. I can visualize it as a
wave that is coming from the left, moving here. Or a wave that is coming from the right. So
let’s visualize this solution as a wave that’s coming from the left. It will be a little easier.
So I will write it. A e to the ikx. OK. Why is it coming from the left? Because if you put the
energy-- that I will not put it, it’s in stationary state, presumably this is a state with some fixed
energy. You will have a factor e to the minus iEt over h bar. And when you see kx minus Et,
you know that that’s a wave that is moving to the right. So this A e to the ikx is moving to the
right.
And then what will happen? Now it’s a matter of finding a solution of Schrodinger’s equation.
So you can try to find the solution of Schrodinger equation, but you have to write some
answers for what’s happening on the right. I will write an answer here, that we’ll put C e to the i
k bar x. And another k. Well, we’ll see now what those k’s are.
I say the following. Here, the energy is bigger than the potential so it has to be a wave. But
here they energy is still bigger than the potential so it also must be a wave. But a wave with
different kinetic energy, different momentum, therefore different de Broglie wavelength and
different k. But we know from Schrodinger’s equation what that should be. This wave is also
moving to the right, because probably if I have a wave moving to the right here, it produces
some transmitted wave to the right. But then, you could try solving the Schrodinger equation
with this. It won’t be enough because physically you would expect the wave bouncing back as
well from here.
So I will put a B e to that minus ikx. That’s a wave moving towards the left with an unknown
coefficient. And, now let’s get those constants. I’ll finish in two minutes.
What is k? Well if you have energy E, you know that the energy is h squared k squared over 2
m. You can look at the Schrodinger equation with 0 potential over there. And therefore, k
squared is also 2mE over h squared. It’s a combination you’ve been seeing quite a bit. The
intuition for k bar should be that k bar squared is 2m times the kinetic energy, so it should be e
minus v0 over h squared.
So these are k and k bar. And the wave function must be continuous at x equals 0. That gives
you A plus B equal to C. At 0, all the exponentials vanish. And the derivative must be
continuous at x equals 0. And the derivative being continuous because there’s no delta
function anywhere here. So you have ikA, that’s the derivative of the first term, minus ikB, the
same k in that region of course, is equal to i k bar C. So from this you get A minus B is equal to
k bar over kC.
Two equations and two unknowns. And that’s OK, even though there are three coefficients,
because the way to think of this is that you’re sending in some wave and you’re going to get
some reflection and some transmission. So in some sense, A is the input. You could want to
call it 1 or whatever. So what we’re looking for is what is B over A? And what is C over A?
And these two equations, it’s a one line computation. I’ll write the answer. B over A is k minus k
bar over k plus k bar. Do it for fun. And C over A is 2k over k plus k bar.
B gives you a sense of how much is reflected. C, how much is transmitted. But this is the
beginning. Because this is not a particle coming in. So we’ll have to build the packet and send
it in and see how this relations tell you what’s going to happen. So this is a nice story that we
will develop next time.

  • 7
    点赞
  • 8
    收藏
    觉得还不错? 一键收藏
  • 0
    评论
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