Lecture 14: Simple harmonic oscillator II. Creation and annihilation operators.

L14.1 Recursion relation for the solution (12:25)

L14.2 Quantization of the energy (23:23)

L14.3 Algebraic solution of the harmonic oscillator (16:50)

L14.4 Ground state wavefunction (15:58)

L14.1 Recursion relation for the solution (12:25)

MITOCW | watch?v=RxWfrE3o-9k
PROFESSOR: This was a differential equation for the energy eigenstates phi. Supposed to be normalizable
functions. We looked at this equation and decided we would first clean out the constant. We
did that by replacing x by a unit-free coordinate u.
For that we needed a constant that carries units of length, and that constant is given by this
combination of the constants of the problem. H-bar, m, and omega-- the frequency of the
oscillator.
We also defined the unit-free energy-- calligraphic e. In terms of which the real energies are
given by multiples of h omega over 2. So the problem has now become-- and this whole
differential equation turns into this simple differential equation. Simple looking-- let’s say
properly-- differential equation for phi-- as a function of u-- which is the new rescaled
coordinate, and where the energy shows up here.
And for some reason, this equation doesn’t have normalizable solutions. Unless those
energies are peculiar values that allow a normalizable solution to exist.
We looked at this equation for u going to infinity, and realized that e to the plus minus u
squared over 2 are the possible dependencies. So we said-- without loss of generality-- that
phi could be written as some function of u to be determined times this exponential.
And we hope for a function that may be a polynomial. So that the dependence at the infinity is
governed by this factor.
So with this for the function phi, we substitute back into the differential equation. Now the
unknown is h. So you can take the derivatives and find this differential equation for h-- a
second order differential equation. And that’s the equation.
It may look a little more complicated than the equation we started with, but it’s much simpler,
actually. There would be no polynomial solution of this equation, but there may be a
polynomial solution of the second equation.
So we have to solve this equation now. And the way to do it is to attempt a serious expansion.
So we would try to write h of u equals the sum over j. Equals 0 to infinity. Ak, u to the k. P
equals 0 to infinity.
Now, one way to proceed with this is to plot this expansion into the differential equation. You
Now, one way to proceed with this is to plot this expansion into the differential equation. You
will get three sums. You will have to shift indices. It’s kind of a little complicated. Actually,
there’s a simpler way to do this in which you think in the following way.
You have this series and you imagine there’s a term aj, u to the j, plus aj plus 1, u to the j plus
1, plus aj plus 2. U to the j plus 2. And you say, let me look at the terms with u to the j in the
differential equation. So just look at the terms that have a u to the power j.
So from this second-- h vu squared-- what do we get? Well, to get a term that has a u to the j,
you must start-- if you take two derivatives and to end up with u to the j-- you must have
started with this. U to the j plus 2.
So this gives you j plus 2, j plus 1-- taking the two derivatives-- aj plus 2, u to the j. From the
series, the term with u to the j from the second hvu squared is this one.
How about for minus 2u dh, du? Well, if I start with h and differentiate and then multiply by u,
I’m going to get u to the j starting from u to the j. Because when I differentiate I’ll get u to the j
minus 1, but the u will bring it back. So this time I get minus 2 a j-- or minus 2. One derivative j.
That’s aj, u to the j.
So it’s from this [? step. ?] Minus 2. You differentiate and you get that.
From the last term, e minus 1 times h, it’s clearly e minus 1 times aj u to the j. So these are my
three terms that we get from the differential equation.
So at the end of the day, what have we gotten? We’ve gotten j plus 2, j plus 1, aj plus 2, minus
2jaj, plus e minus 1 aj. All multiplied by u to the j. And that’s what we get for u to the j.
So if you wish, for the whole differential equation-- all of the differential equation-- you get the
sum from j equals 0 to infinity of these things. And that should be equal to zero.
So this is the whole left-hand side of the differential equation. We calculated what is the term u
to the j. And there will be terms from u to the zeroth to u to the infinity. So that’s the whole
thing.
And we need this differential equation to be solved. So this must be zero. And whenever you
have a function of u like a polynomial-- well, we don’t know if it’s a polynomial-- and it stops.
But if you have a function of u like this, each coefficient must be 0.
Therefore, we have that j plus 2, times j plus 1, times aj plus 2, is equal to 2 j plus 1 minus e,
aj. I set this whole combination inside brackets to 0. So this term is equal to this term and that
term on the other side. You get a plus 2j. A plus 1 and minus e.
So basically this is a recursion relation. Aj plus 2 is equal to 2j plus 1 minus e, over j plus 2, j
plus 1 times aj.
And this is perfectly nice. This is what should have happened for this kind of differential
equation-- a second-order linear differential equation. We get a recursion that jumps one step.
That’s very nice. And this should hold for j equals 0, 1, 2-- all numbers.
So when you start solving this, there’s two ways to solve it. You can decide, OK, let me
assume that you know a0-- you give it. Give a0. Well, from this equation-- from a0-- you can
calculate a2. And then from a2 you can calculate a4. And successively.
So you get a2, a4-- all of those. And this corresponds to an even solution of the differential
equation for h. Even coefficients. Even solution for h. Or-- given this recursion-- you could also
give a1-- give it-- and then calculate a3, a5-- and those would be an odd solution.
So you need two conditions to solve this. And those conditions are a0 and a1, which is the
same as specifying the value of the function h at 0-- because the value of the function h at 0 is
a0. And the value of the derivative of the function at 0, which is a1.
[INAUDIBLE] h of mu is a0 plus a1u plus a2u squared. So the derivative at 0 is [? h1. ?] And
that’s what you must have for solving a differential equation-- a second-order differential
equation for h. You need to know the value of the function at zero and the value of the function
at the derivative of the function-- at zero, as well.
And then you can start integrating it. So this first gives you a solution a0 plus a2u squared,
plus a4u to the fourth. And the second is an a1u plus a3u cubed, plus these ones.
So all looks pretty much OK.

L14.2 Quantization of the energy (23:23)

MITOCW | watch?v=Y6Ma-zn4Olk
PROFESSOR: We have to ask what happens here? This series for h of u doesn’t seem to stop. You go a 0, a
2, a 4. Well, it could go on forever. And what would happen if it goes on forever?
So if it goes on forever, let’s calculate what this aj plus 2 over aj as j goes to infinity. Let’s see
how the coefficients vary as you go higher and higher up in the polynomial. That should be an
interesting thing.
So I pass the aj that is on the right side and divide it, and now on the right-hand side there’s
just this product of factors. And as j goes to infinity, it’s much larger than 1 or e, whatever it is,
and the 2 and the 1 in the denominator. So this goes like 2j over j squared. And this goes
roughly like 2 over j.
So as you go higher and higher up, by the time j is a billion, the next term is 2 divided by a
billion. And they are decaying, which is good, but they’re not decaying fast enough. That’s a
problem. So let’s try to figure out if we know of a function that decays in a similar way.
So you could do it some other way. I’ll do it this way. e to the u squared-- let’s look at this
function-- is this sum from n equals 0 to infinity 1 over n u squared to the n. So it’s u to the 2n–
1 over n factorial, sorry.
So now, since we have j’s and they jump by twos, these exponents also here jump by two. So
that’s about right. So let’s think of 2n being j, and therefore this becomes a sum where j equals
0, 2, 4, and all that of 1 over-- so even j’s.
2n is equal to j-- so j over 2 factorial over 1, and then you have u to the j. So if I think of this as
some coefficient c sub j times u to the j, we’ve learned that c sub j is equal to 1 over j divided
by two factorial. In which case, if that is true, let’s try to see what this cj plus 2 over cj-- the
ratio of two consecutive coefficients in this series.
Well, cj plus 2 would be j plus 2 over 2 factorial, like this. That’s the numerator, because of that
formula. And the denominator would have just j over 2 factorial.
Now, these factorials make sense. You don’t have to worry that they are factorials of halves,
because j is even. And therefore, the numerators are even-- divided by 2. These are integers.
These are ordinary factorials.
There are factorials of fractional numbers. You’ve seen them probably in statistical physics
and other fields, but we don’t have those here. This is another thing. So this cancels.
If you have a number and the number plus 1, which is here, you get j plus 2 over 2, which is 2
over j plus 2. And that’s when j is largest, just 2 over j, which is exactly what we have here. So
this supposedly nice, innocent function, polynomial here-- if it doesn’t truncate, if this recursive
relation keeps producing more and more and more terms forever-- will diverge. And it will
diverge like so. If the series does not truncate, h of u will diverge like e to the u squared.
Needless to say, that’s a disaster. Because, first, it’s kind of interesting to see that here, yes,
you have a safety factor, e to the minus u squared over 2. But if h of u diverges like e to the u
squared, you’re still in trouble. e to the u squared minus u squared over 2 is e to the plus u
squared over 2. And it actually coincides with what we learned before, that any solution goes
like either plus or minus u squared over 2.
So if h of u doesn’t truncate and doesn’t become a polynomial, it will diverge like e to the u
squared, and this solution will diverge like e to the plus u squared over 2, which was a
possibility. And it will not be normalizable. So that’s basically the gist of the argument.
This differential equation-- whenever you work with arbitrary energies, there’s no reason why
the series will stop. Because e there will have to be equal to 2j plus 1, which is an integer. So
unless je is an integer, it will not stop, and then you’ll have a divergent-- well, not divergent;
unbounded-- far of u that is impossible to normalize.
So the requirement that the solution be normalizable quantizes the energy. It’s a very nice
effect of a differential equation. It’s very nice that you can see it without doing numerical
experiments, that what’s going on here is an absolute requirement that this series terminates.
So here, phi of u would go like e to the u squared over 2, what we mentioned there, and it’s
not a solution.
So if the series must terminate, the numerator on that box equation must be 0 for some value
of j, and therefore there must exist a j such that 2j plus 1 is equal to the energy. So basically,
what this means is that these unit-free energies must be an odd integer. So in this case, this
can be true for j equals 0, 1, 2, 3. In each case, it will terminate the series.
With j equals 0, 1, 2, or 3 there, you get some values of e that the series will terminate. And
when this series terminates, aj plus 2 is equal to 0. Because look at your box equation. aj, you
got her number, and then suddenly you get this 2j plus 1 minus e.
And if that’s 0, the next one is zero. So, yes, you get something interesting even for j equals 0.
Because in that case, you can have a0, but you will have no a2, just the constant. So I will
write it.
So if aj plus 2 is equal to zero, h of u will be aj u to the j plus aj minus 2u to the j minus 2. And
it goes down. The last coefficient that exists is aj, and then you go down by two’s.
So let’s use the typical notation. We call j equals n, and then the energy is 2n plus 1. The h is
an u to the n plus an minus 2. You do the n minus 2, and it goes on.
If n is even, it’s an even solution. If n is odd, it’s an odd solution. And the energy e, remember,
was h omega over 2 times e-- so 2n plus 1. So we’ll move the 2 in, and e will be equal to h
omega n plus 1/2.
And n in all these solutions goes from 0, 1, 2, 3. We can call this the energy en. So here you
see another well-known, famous fact that energy levels are all evenly spaced, h omega over 2,
one by one by one-- except that there’s even an offset for n equals 0, which is supposed to be
the lowest energy state of the oscillator. You still have a 1/2 h bar omega.
This is just saying that if you have the potential, the ground state is already a little bit up. You
would expect that-- you know there’s no solutions with energy below the lowest point of the
potential. But the first solution has to be a little bit up. So it’s here and then they’re all evenly
spaced.
And this begins with E0; for n equals 0, e1. And there’s a little bit of notational issues. We used
to call the ground state energy sometimes e1, e2, e3, going up, but this time it is very natural
to call it E0 because it corresponds to n equals 0. Sorry. Those things happen.
No, it’s not an approximation. It’s really, in a sense, the following statement. Let me remind
everybody of that statement. When you have even or odd solutions, you can produce a
solution that you may say it’s a superposition, but it will not be an energy eigenstate anymore.
Because the even solution that stops, say, at u to the 6 has some energy, and the odd solution
has a different energy. So these are different energy eigenstates. So the energy eigenstates,
we prove for one-dimensional potentials, are not chosen to be even or odd for bound states.
They are either even or odd.
You see, a superposition-- how do we say like that? Here we have it. If this coefficient is even,
the energy sum value-- if this coefficient is odd, the energy will be different. And two energy
eigenstates with different energies, the sum is not an energy eigenstate.
You can construct the general solution by superimposing, but that would be general solutions
of the full time-dependent Schrodinger equation, not of the energy eigenstates. The equation
we’re aiming to solve there is a solution for energy eigenstates. And although this concept I
can see now from the questions where you’re getting, it’s a subtle statement.
Our statement was, from quantum mechanics, that when we would solve a symmetric
potential, the bound states would turn out to be either even or odd. It’s not an approximation.
It’s not a choice. It’s something forced on you.
Each time you find the bound state, it’s either even or it’s odd, and this turned out to be this
case. You would have said the general solution is a superposition, but that’s not true. Because
if you put a superposition, the energy will truncate one of them but will not truncate the other
series. So one will be bad. It will do nothing.
So if this point is not completely clear, please insist later, insist in recitation. Come back to me
office hours. This point should be eventually clear. Good.
So what are the names of these things? These are called Hermite polynomials. And so back to
the differential equation, let’s look at the differential equations when e is equal to 2n plus 1. Go
back to the differential equation, and we’ll write d second du squared Hn of u.
That will be called the Hermite polynomial, n minus 2udHn du plus e minus 1. But e is 2n plus
1 minus 1 is 2n Hn of u is equal to 0. This is the Hermite’s differential equation.
And the Hn’s are Hermite polynomials, which, conventionally, for purposes of doing your
algebra nicely, people figured out that Hn of u is convenient if-- it begins with u to the n and
then it continues down u to the n minus 2 and all these ones here. But here people like it when
it’s 2 to the n, u to the n-- a normalization. So we know the leading term must be u to the n.
If you truncate with j, you’ve got u to the j. You truncate with n, you get u to the n. Since this is
a linear differential equation, the coefficient in front is your choice. And people’s choice has
been that one and has been followed.
A few Hermite polynomials, just a list. H0 is just 1. H1 is 2u. H2 is 4u squared minus 2. H3 is
A few Hermite polynomials, just a list. H0 is just 1. H1 is 2u. H2 is 4u squared minus 2. H3 is
our last one, 8u cubed minus 12u, I think. I have a little typo here. Maybe it’s wrong. So you
want to generate more Hermite polynomials, here is a neat way that is used sometimes.
And these, too, are generating functional. It’s very nice actually. You will have in some
homework a little discussion.
Look, you put the variable z over there. What is z having to do with anything? u we know, but
z, why? Well, z is that formal variable for what is called the generating function. So it’s equal to
the sum from n equals 0 to infinity. And you expand it kind of like an exponential, zn over n
factorial.
But there will be functions of u all over there. If you expand this exponential, you have an
infinite series, and then you have to collect terms by powers of z. And if you have a z to the 8,
you might have gotten from this to the fourth, but you might have gotten it from this to the 3
and then two factors of this term squared or a cross-product.
So after all here, there will be some function of u, and that function is called the Hermite
polynomial. So if you expand this with Mathematica, say, and collect in terms of u, you will
generate the Hermite polynomials. With this formula, it’s kind of not that difficult to see that the
Hermite polynomial begins in this way.
And how do you check this is true? Well, you would have to show that such polynomials satisfy
that differential equation, and that’s easier than what it seems. It might seem difficult, but it’s
just a few lines.
Now, I want you to feel comfortable enough with this, so let me wrap it up, the solutions, and
remind you, well, you had always u but you cared about x. So u was x over a. So let’s look at
our wave functions.
Our wave functions phi n of x will be the Hermite polynomial n of u, which is of x over a, times
e to the minus u squared over 2, which is minus x squared over 2a squared. And you should
remember that a squared is h bar over m omega. So all kinds of funny factors-- in particular,
this exponential is e to the minus x squared m omega over h squared over 2. I think so-- m
omega over 2h bar.
Let me write it differently-- m omega over 2h bar x squared. That’s that exponential, and those
are the coefficients. And here there should be a normalization constant, which I will not write.
It’s a little messy. And those are the solutions. And the energies en were h bar omega over 2 n
plus 1/2, so E0 is equal to h bar omega over 2. E1 is 3/2 of h bar omega, and it just goes on
like that.

L14.3 Algebraic solution of the harmonic oscillator (16:50)

MITOCW | watch?v=8CCFPgd_P1w
PROFESSOR: SHO algebraically. And we go back to the Hamiltonian, p squared over 2m plus 1/2 m omega
squared x hat squared. And what we do is observe that this some sort of sum of squares plus
p squared over m-- p squared over m squared omega squared. So the sum of two things
squared. Now, the idea that we have now is to try to vectorize the Hamiltonian.
And what we call vectorizing is when you write your Hamiltonian as the product of two vectors,
V times W. Well actually, that’s not quite the vectorization. You want kind of the same vector,
and not even that. You sort of want this to be the Hermitian conjugate of that. And if there is a
number here, that’s OK. Adding numbers to a Hamiltonian doesn’t change the problem at all.
The energies are all shifted, and it’s just how you’re defining the zero of your potential, is doing
nothing but that.
So vectorizing the Hamiltonian is writing it in this way, as V dagger V. And you would say, why
V dagger V? Why not VV dagger or VV or V dagger V dagger? Well, you want the Hamiltonian
to be Hermitian. And this thing is Hermitian. You may recall that AB dagger. The Hermitian
conjugate of AB dagger is B dagger A dagger. So the Hermitian conjugate of this product is V
dagger times the dagger of V dagger. A dagger of a dagger is the same operator, when you
dagger it twice, you get the same.
So this is Hermitian. V dagger times V is a Hermitian operator, and that’s a very good thing.
And there will be great simplifications. If you ever succeed in writing a Hamiltonian this way,
you’ve gone 90% of the way to solving the whole problem. It has become infinitely easier, as
you will see in a second, if you could just write this vectorization.
So if you had x minus-- x squared minus this, you would say, oh, clearly that’s-- A squared
minus B squared is A minus B time A plus B, but there’s no such thing here. It’s almost like A
squared plus B squared. And how do you sort of factorize it? Well, actually, since we have
complex numbers, this could be A minus IB times A plus IB. That is correctly A squared plus B
squared, and complex numbers are supposed to be friends in quantum mechanics, so having
Is, there’s probably no complication there.
So let’s try that. I’ll write it. So here we have x squared plus p squared over m squared omega
squared. And I will try to write it as x minus i p hat over m omega times x plus I p hat over m
omega. Let’s put the question mark before we are so sure that this works. Well, some things
work. The only danger here is that these are operators and they don’t commute. And when we
do this, in one case, in the cross-terms, the A is to the left of B, but the other problem the B is
to the left of A. So we may run into some trouble. This may not be exactly true.
So what is this? This x with x, fine. x squared. This term, p with ps, correct. Plus p squared
over m squared omega squared. But then we get plus i over m omega, x with p minus p with x,
so that x, p commutator. So vectorization of operators in quantum mechanics can miss a few
concepts because things don’t commute. So the cross-terms give you that, and this x, p is I h
bar, so this whole term will give us the following statement.
What we’ve learned is that what we wanted, x squared plus p squared over m squared omega
squared is equal to-- so I’m equating this line to the top line-- is equal to x hat minus i p hat
over m omega times x hat plus i p hat over m omega. And then, from this whole term, i with i is
minus, so it’s h bar over m omega. So I’ll put it in-- it’s a minus h bar over m, [INAUDIBLE]. So
here is plus h bar over m omega times a unit vector, if you wish.
OK. So this is very good. In fact, we can call this V dagger and this V. Better call this V first and
then ask, what is the dagger of this operator? Now, you may remember that, how did we
define daggers? If you have phi with psi and the inner product-- with an integral of five star psi-

  • if you have an A psi here, that’s equal to A dagger phi psi.
    So an operator is acting on the second wave function, moves as A dagger into the first wave
    function. And you know that x moves without any problem. x is Hermitian. We’ve discussed
    that p is Hermitian as well, moves to the other side. So the Hermitian conjugate of this
    operator is x, the p remains means p, but the i becomes minus i. So this is correct. If this
    second operator is called V, the first operator should be called V dagger. That is a correct
    statement. One is the dagger of the other one.
    So the Hamiltonian is 1/2 m omega squared times this sum of squares, which is now equal to
    V dagger V plus h bar over m omega. So h hat is now 1/2 m omega squared V dagger V plus
    a sum, which is plus 1/2 h bar omega. So we did it. We vectorized the Hamiltonian V dagger V,
    and this is quite useful.
    So the Vs, however, have units. And you probably are aware that we like things without units,
    so that we can see the units better. This curve is perfectly nice. It’s a number added to the
    Hamiltonian. It’s h omega, it has units of energy, but this is still a little messy. So let’s try to
    clean up those Vs, and the way I’ll do it is by computing their commutator, to begin with.
    So let’s compute the commutator of V and V dagger and see how much is that commutator.
    It’s a simple commutator, because it involves vectors of x and V. So it’s the commutator of x
    plus ip over m omega, that’s V, with x minus ip over m omega. So the first x talks only to the
    second piece, so it’s minus i over m omega x, p. And for the second case, you have plus i over
    am omega p with x. This is i h bar, and this is minus i h bar. Each term will contribute the
    same, i times minus i is plus, so h bar over and, omega times the 2. That is V dagger V. VV
    dagger, I’m sorry. 2 h bar over m omega.
    So time to change names a little bit. Let’s do the following. Let’s put square root of m omega
    over 2 h bar V. Have a square root of m omega over 2 h bar V dagger, commute to give you
  1. That’s a nice commutator. It’s one number-- or an operator is the same thing. So I brought
    the square root into each one.
    And we’ll call the first term-- because of reasons we’ll see very soon-- the destruction operator,
    A square root of m omega over 2 h bar V. It’s called the destruction operator. And the dagger
    is going to be A dagger. Some people put hats on them. I sometimes do too, unless I’m too
    tired. 2h bar V dagger. And those A and A daggers are now unit-free-- and you can check
    That-- Because they have the same units. And A with A dagger is the nicest commutator, 1.
    Is A a Hermitian operator? Is it? No. A is not Hermitian. A dagger is different from A. A is
    basically this thing, A dagger is this thing. So not Hermitian. So we’re going to work with these
    operators. They’re non-Hermitian. I need to write the following equations. It’s very-- takes a
    little bit of writing, but they should be recorded, they will always make it to the formula sheet.
    And it’s the basic relation between A, A dagger, and x and p. A is this, A dagger, as you know,
    is x minus ip hat over m omega. Since I’m copying, I’d better copy them right. x, on other
    hand, is the square root of h bar over 2m omega A plus A dagger, and p is equal to i square
    root of m omega h bar over 2 A dagger minus A.
    So these four equations, A and A dagger in terms of x and p and vice versa, are important.
    They will show up all the time. Here are the things to notice. A and A dagger is visibly clear
    that on is the Hermitian conjugate of the other. Here, x is Hermitian. And indeed, A plus A
    dagger is Hermitian. When you do the Hermitian conjugate of A plus A dagger, the first A
    becomes an A dagger. The second A, with another Hermitian conjugation, becomes A. So this
    is Hermitian. But p is Hermitian, and here we have A dagger minus A. This is not Hermitian, it
    changes sign. Well, the i is there for that reason, and makes it Hermition. So there they are,
    they’re Hermitian, they’re good.

L14.4 Ground state wavefunction (15:58)

MITOCW | watch?v=vnyxYtj0mfE
PROFESSOR: So let’s write the Hamiltonian again in terms of v and v dagger. So for this equation, v dagger
v, from this equation, is equal to 2 h bar over m omega, a dagger a. And immediately above
equation v dagger v’s, this-- we substitute into the Hamiltonian. And Hamiltonian becomes the
nice object h bar omega, a dagger a plus 1/2 if you want.
All right. We did this hard work of factorization. We have to show what’s good for. Well, in fact,
we’re going to be able to solve the harmonic oscillator without ever talking about differential–
almost ever talking about differential equations. In fact, we will not talk about the second order
differential equation. Thanks to our great work here, we will have to talk about a first order
differential equation, and a much simpler one. And only one, not for and n 1, 2, 3, infitinty,
infinity number of polynomials.
It’s a great simplification. Other Hamiltonians admit factorization. In fact, there’s whole books
of factorizable Hamiltonians, because those are the nicest Hamiltonians to solve. Let’s see
why, though. We haven’t said why yet.
Here is the leading thing that we can do. Remember we recalled 5 psi the integral dx phi star
of x psi of x. This is just notation. So the expectation value, calculate the expectation value of
the Hamiltonian in some state psi. Could be the general state sum.
So what are you supposed to do? You’re supposed to do psi h psi. This is normalized state,
the expectation value is the integral of psi star h psi. That’s what this is. But now, let’s put in
this information. And the expectation value of this would be psi h-- or let me do it this way-- h
bar ome-- well, let’s go slow. Psi h omega a dagger a a psi plus h omega over 2 psi.
So I just calculated h on psi, and I wrote what it is-- h omega this, this term. So this is two
terms-- h omega psi a dagger a psi plus h omega over 2 psi psi. OK. So what did I gain with
the factorization? So far, it looks like nothing. But here we go-- this term is equal to 1, because
the wave function is normalized.
And here I can do one thing-- I can remember my definition of a Hermitian conjugate. I can
move an operator and put its Hermitian conjugate on the other side. So think of this operator,
a dagger-- a dagger is acting on this wave function. What is a dagger? It’s this. And p is h bar
over idex.
So this-- you know how to act. But if a dagger is here, I can put it on the first wave function, but
I must put the dagger of this operator, and the dagger of a dagger is a. So this is h omega, a
psi a psi plus h omega over 2.
Now here comes the next thing. If this is an inner product, any phi phi is greater or equal than
0, because you would have phi star phi, and that’s positive. So any of that [INAUDIBLE] is
greater or equal than 0. Note-- here you have some function, but here the same function. It is
this case. That thing is greater or equal than 0. That is the great benefit of the factorized
Hamiltonian-- if h has a v dagger v, you can flip the v dagger here and it becomes v psi v psi,
and it’s positive. And you’ve learned something very important, and you can get positive
energies.
In fact, from here, since this is positive, this must be greater or equal than h omega over 2.
Because this is greater than equal than 0. So the expectation value of the Hamiltonian-- if you
would be thinking now of energy eigenstates, the energy eigenvalue is the expectation value
of the Hamiltonian in an energy eigenstate must be greater than h bar omega over 2.
And in some blackboard that has been erased, we remember that the lowest energy state had
energy-- there it is. The lowest energy state has energy h omega over 2. So look at this, and
you say, OK, this shows that any eigenstate must have energy greater than h omega over 2.
But could there be one state for which the energies exactly h omega over 2.
Yes, if this inner product is 0. But for an inner product of two things to be 0, each function must
be 0. So from this, we conclude that if there is a ground state, it’s a state for which a phi-- or a
psi is equal to 0. So this is a very nice conclusion. So if the lower bound is realized, so that you
get a state with energy equal h bar over 2, then it must be true that a psi is equal to 0.
And a psi equal to 0 means x plus ip over m omega on psi is equal to 0, or x plus p is h bar
over i ddx, so this is h bar over m omega d/dx on psi of x is equal to 0. And that was the
promised fact. We have turned the second order differential equation into a first order
differential equation.
Think of that magic that has happened to do that. You had a second order differential equation
because the Hamiltonian has x squared b squared. By factorizing, you go two first order
differential operators. And by Hermeticity, you were led to the condition that the lowest energy
state had to be killed by a. That’s why a is called the annihilation operator. It should be killed.
And now you have to solve a first order differential equation, which is a game. An easy game
compared with a second order differential equation.
So let’s, of course, solve it. It doesn’t take any time. Let’s call this the ground state. If it exists.
And this gives you d psi 0 v x is equal to minus m omega over h bar x psi . 0. This can be
degraded easily or you can guess the answer. It’s an exponential. Anything that differentiates
that you should extend the same function as an exponential-- e to the minus m omega 2 h
squared x squared is the solution.
Psi 0 of x is equal to some number times that. This was-- the number is the Hermit
polynomials sub 0, and that exponential, this exponential, we wrote a few blackboards ago. It’s
a good exponential. It’s a perfect Gaussian. It’s our ground state. And 0, if you want to
normalize it, m 0 is equal to m omega over phi h bar to the 1/4.
And that is the ground state. And it has energy, h omega over 2. You could see what the
energy is by doing this very simple calculation. Look, get accustomed to these things. H hat psi
0. What is h? Is h omega a dagger a plus 1/2 acting on psi 0. The a acting on psi 0 already
kills it. Because that’s the defining equation. Well that’s 0. And you get 1/2 h bar omega,
confirming that you did get this thing to be correct.
So this is only the beginning of the story. We found the ground state, and now we have to find
the excited states. Let me say a couple of words to set up this discussion for next time. The
excited states appear in a very nice way as well. So first a tiny bit of language, of h bar. This
equal h omega, a dagger a is usually called the number operator. We’ll explain more on that
next time.
So n number operator is a dagger a. It’s a permission operator, and it’s pretty much the
Hamiltonian. It’s the number, it’s called. Why is it called the number is what we have to figure
out. It is a counting operator-- it just looks at the state and counts things. So what does this
give us? Well, we also know that the number operator kills phi 0, because a kills psi 0. A kills it.
So that’s what we have.
So we did say that a was a destruction operator, annihilation operator, because it annihilates
the ground state. So if a annihilates the ground state, a dagger cannot annihilate the ground
state. Why? Because a dagger with a computator is equal to 1. Look at this. This is a a dagger
minus a dagger a.
Act on the ground state. That’s it. Now this term kills it. But this term better not kill it, because it
has to give you back the ground state if this is true. And this is true. So a dagger doesn’t kill
the ground state. Since it doesn’t kill it, it’s called a creation operator.
So you have this state, but now there’s also this state a dagger acting on the vacuum. And
there’s a state a dagger a dagger acting on the vacuum. And all those. And what we will figure
out next time is that, yes, this is the ground state. And this is the first excited state. And this is
the second excited state. And goes on forever. So we’ll have a very compact formula for the
excited states of the harmonic oscillator. They’re just creation operators acting on the ground
state or the [INAUDIBLE].

  • 8
    点赞
  • 9
    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