Lecture 21: Legendre equation. Radial equation. Hydrogen atom 2-body problem.

L21.1 Associated Legendre functions and spherical harmonics (18:51)

L21.2 Orthonormality of spherical harmonics (17:57)

L21.3 Effective potential and boundary conditions at r=0 (14:28)

L21.4 Hydrogen atom two-body problem (25:04)

L21.1 Associated Legendre functions and spherical harmonics (18:51)

MITOCW | watch?v=Lt2Y6fLJ09Q
PROFESSOR: We’re talking about angular momentum. We’ve motivated angular momentum as a set of
operators that provided observables, things we can measure. Therefore, they are important.
But they’re particularly important for systems in which you have central potentials. Potentials
that depend just on the magnitude of the radial variable. A v of r that depends just on the
magnitude of the vector r relevant to cases where you have two bodies interacting through a
potential that just depends on the distance between the particles.
So what did we develop? Well, we discussed the definition of the angular momentum operator.
You saw they were permission. We found that they satisfy a series of commutators in which lx
with ly gave ih bar lz, and cyclical versions of that equation, which ensure that actually you can
measure simultaneously the three components of angular momentum. You can measure, in
fact, just one. Happily we found there was another object we could measure, which was the
square of the total angular momentum.
Now, you should understand this symbol. It’s not a vector. It is just a single operator. l squared
is, by definition, lx times lx, plus ly times ly, plus lz times lz. This is this operator.
And we showed that any component of angular momentum, be it lx, ly, or lz, commutes with l
squared. Given that they commute, it’s a general theorem that two permission operators that
commute, you can find simultaneous eigenstates of those two operators. And therefore, we
set up for the search of those wave functions that are simultaneous eigenstates of one of the
three components of angular momentum. Everybody chooses lz and l squared. lz being
proportional to angular momentum has an h bar m.
We figured out by looking at this differential equation that if we wanted single valued wave
functions-- wave functions would be the same at phi, and at phi plus 2 pi, which is the same
point. You must choose m to be an integer. For the l squared operator we also explained that
the eigenvalue of this operator should be positive. That is achieved when l, whatever it is, is
greater than 0, greater or equal than 0. And the discussion that led to the quantization of l was
a little longer, took a bit more work.
Happily we have this operator, and operator we can diagonalize, or we can find eigenstates for
it. Because the Laplacian, as was written in the previous lecture, Laplacian entering in the
Schrodinger equation has a radial part and an angular part, where you have dd thetas, and
sine thetas, and the second defies square. All these things were taken care of by l squared.
And that’s very useful.
Well, the differential equation for l squared-- this can be though as a differential equation–
ended up being of this form, which is of an equation for the so-called Associate Legendre
functions. For the case of m equals 0 it simplifies very much so that it becomes an equation for
what were eventually called Legenre polynomials.
We looked at that differential equation with m equals 0. We called it pl 0. So we don’t write the
zeros. Everybody writes pl for those polynomials.
And looking at the differential equation one finds that they have divergences at theta equals 0,
and a theta equal pi, north and south pole of this spherical coordinate system. There aren’t
divergences unless these differential equations has a polynomial solution that this is serious
the recursion relations terminate.
And that gave for us the quantization of l. And that’s where we stopped. These are the
Legendre polynomials. Solve this equation for m equals 0.
Are there any questions? Anything about the definitions or? Yes?
AUDIENCE: Why do we care about simultaneous eigenstates?
PROFESSOR: Well, the question is why do we care about simultaneous eigenstates. The answer is that if you
have a system you want to figure out what are the properties of the states. And you could
begin by saying the only thing I can know about this state is its energy. OK, well, I know the
energy at least.
But maybe thinking harder you can figure out, oh, you can also know the momentum. That’s
progress. If you can also know the angular momentum you learn more about the physics of
this state. So in general, you will be led in any physical problem to look for the maximal set of
commuting operators. The most number of operators that you could possibly measure. You
know you have success at the very least, if you can uniquely characterize that states of the
system by observables.
Let’s assume you have a particle in a circle. Remember that the free particle in a circle has
degenerate energy eigenstates. So you have two energy eigenstates for every allowed
energy, except for 0 energy, but two energy eigenstates. And you would be baffled.
You’d say, why do I have two? There must be some difference between these two states. If
there are two states, there must be some property that distinguishes them. If there is no
property that distinguishes them, they should be the same state. So you’re left to search for
another thing.
And in that case the answer was simple. It was the momentum. You have a particle with some
momentum in one direction, or in the reverse direction. So in general, it’s a most important
question to try to enlarge the set of commuting observables. Leading finally to what is initially
called a complete set of commuting observables.
So what do we have to do today? We want to complete this analysis. We’ll work back to this
equation. And then work back to the Schrodinger equation to finally obtain the relevant
differential equation we have to solve if you have a spherical symmetric potential. So the
equation will be there in a little while.
Then we’ll look at the hydrogen atom. We’ll begin the hydrogen atom and this task why?
Having a proton and an electron we can reduce this system to as if we had one particle in a
central potential. So that will be also very important physically.
So let’s move ahead. And here there is a simple observation that one can make. Is that the
differential equation for p l m depends on m squared.
We expect to need values of m that are positive and negative. You have wave functions here,
of this form. The complex conjugate ones should be thought as having m negative. So we
expect positive and negative m’s to be allowed.
So how did people figure this out? They, in fact, figured out that if you have these polynomials
you can create automatically the solutions for this equation. There’s a rule, a simple rule that
leads to solutions. You put p l m of x is equal to 1 minus x squared, to the absolute value of m
over 2.
So there are square roots here, possibly. An absolute value of m means that this is always in
the numerator, whether m is positive or negative. And d, dx acting exactly absolute value of m
times on p l x. The fact is that this definition solves the differential equation star.
This takes a little work to check. I will not check it, nor the notes will check it. It’s probably
something you can find the calculation in some books. But it’s not all that important. The
important thing to note here is the following.
That this provides solutions. Since this polynomial is like x to the l plus x to the l minus 2 plus
coefficients like this. You can think that most m equal l derivatives-- if you take more than l
derivatives you get 0. And there’s no great honor in finding zero solution of this equation.
These are no solutions.
So this produces solutions for an, absolute value of m, less than l. So produces solutions for
absolute value of m less or equal to l. And therefore m in between l and minus l.
But that’s not all that happens. There’s a little more that takes mathematicians some skill to do.
It’s to show that there are no more solutions. You might seem that you were very clever and
you found some solutions, but it’s a theorem that there are no more solutions. No additional
regular solutions. I mean solutions that don’t diverge.
So this is very important. It shows that there is one more constraint on your quantum numbers.
This formula you may forget, but you should never forget this one. This one says that if you
choose some l which corresponds to choosing the magnitude of the angular momentum, l is
the eigenvalue that tells you about the magnitude of the angular momentum.
You will have several possibilities for m. There will be several states that have the same l, but
m different.
So for example you’ll have l equal 0, in which case m must be equal to 0. But if you choose
state with l equals 1, or eigenfunctions with l equal 1, there is the possibility of having m equals
minus 1, 0, or 1. So are three waves functions in that case. Psi 1, minus 1, psi 1, 0, and psi 1,
1.
So in general when we choose a general l, if you choose an arbitrary l, then m goes from
minus l, minus l plus 1 all the way up to l. These are all the values which are 2l plus 1 values.
2l and the 0 value in between. So it’s 2l plus 1 values.
The quantization in some sense is done now. And let me recap about these functions now. We
mentioned up there that the y l m’s are the objects. The spherical harmonicas are going to be
those wave functions. And they have a normalization, n l m, an exponention, and all that.
So let me write, just for the record, what a y l m looks like with all the constants. Well, the
normalization constant is complicated. And it’s kind of a thing you can never remember by
heart. It would be pointless. OK. All of that.
Then a minus 1 to the m seems useful. e to the i m phi, p l m of cosine theta. And this is all
valid for 0 less 0 m positive m. When you have negative m you must do a little variation for m
less than 0 y l m of theta and phi is minus 1 to the m y l minus m of theta and phi complex
conjugated.
Well, if m is negative, minus m is positive. So you know what that is. So you could plug this
whole mess here. I don’t advise it. It’s just for the record.
These polynomials are complicated, but they are normalized nicely. And we just need to
understand what it means to be normalized nicely. That is important for us. The specific forms
of these polynomials we can find them.
The only one I really remember is that y 0 0 is a constant. It’s 1 over 4 pi. That’s simple
enough. No dependents. l equals 0, m equals 0.
Here is another one. y1 plus minus 1 is minus plus square root of 3 over 8 pi e to the plus
minus i phi sine theta. And the last one, so we’re giving all the spherical harmonics with l
equals 1. So with l equals 1 remember we mentioned that you would have three values of m.
Here they are. Plus or minus 1 and 0.

L21.2 Orthonormality of spherical harmonics (17:57)

MITOCW | watch?v=gKSRrTik1SA
PROFESSOR: How do we state the issue of normalization? See, the spherical harmonics are functions of
theta and phi. So it makes sense that you would integrate over theta and phi-- solid angle. The
solid angle is the natural integration. And it’s a helpful integration, because if you have solid
angle integrals and then radial integrals, you will have integrated over all volume.
So for this spherical harmonic, solid angle is the right variable. And you may remember, if you
have solid angle, you have to integrate over theta and phi. Solid angle, you think of it as a
radius of one. Here is sine theta.
So what is solid angle? It’s really the area on a sphere of radius one. The definition of solid
angle is area over radius squared, the part of the solid angle that you have.
If you’re working with a sphere of radius one, it’s the area element is the solid angle. So the
area element in here would be, or the integral over solid angle-- this is solid angle, d omega–
you would integrate from theta equals 0 to theta equals pi of sine theta, d theta. And then you
would integrate from 0 to 2 pi of d phi.
Now, you’ve seen that this equation that began as a differential equation for functions of theta
ending up being the differential equation for a function of cosine theta. Cosine theta was the
right variable. Well, here it is as well, and you should always recognize that. This is minus d of
cosine theta.
And here you would be degrading from cosine theta equals 1 to minus 1. But the minus and
the order of integration can be reversed, so you have the integral from minus 1 to 1 of d cos
theta and then the integral from 0 to 2 pi of alpha. So this is the solid angle integral. You
integrate d of cosine theta from minus 1 to 1 and 0 to 2 pi of d phi.
And we will many times use this notation, the omega, to represent that integral so that we
don’t have to write it. But when we have to write it, we technically prefer to write it this way, so
that the integrals should be doable in terms of cosine theta. So what does this all mean for our
spherical harmonics? Well, our spherical harmonics turned out to be eigenfunctions or
Hermitian operators. And if they have different l’s and m’s, they are having different
eigenvalues.
So eigenfunctions of Hermitian operators with different eigenvalues have to be orthogonal. So
we’ll write the main property, which is the integral of 2l, say l prime, m prime, of theta and phi.
And he put the star here. I’ll put the star just at the Y, and he’s complex conjugated the whole
thing.
Remember that in our problem, one wave function was complex conjugated. The other wave
function is not complex conjugated. They’re different ones because l and m and l prime and m
prime could be different. So orthogonality is warranted. Two different ones with different l’s
and different m’s should give you different values of this integral.
So at this moment, you should get a delta l prime l delta m prime m. And if l and m are the
same as l prime and m prime, you have the same spherical harmonic. And all this tremendous
formula over there, with 2l plus 1 and all these figures, you’re guaranteed that in that case you
get 1 here. So this formula is correct as written. That is the orthonormality of this solution.
Probably, this stage might be a little vague for you. We saw this a long time ago. We may want
to review why eigenfunctions of Hermitian operators with different eigenvalues are orthogonal
and see if you could prove. And you do it, or is it kind of a little fuzzy already? We saw it over a
month ago.
So time to go back to the Schrodinger equation. So for that, we remember what we have. We
have minus h squared over 2m Laplacian of psi plus V of r psi equals E psi. And the Laplacian
has this form, so that we can write it the following way-- minus h squared over 2m 1 over r d
second dr squared r psi-- I won’t close the brackets here-- minus this term.
So I’ll write it minus one over h squared r squared l squared psi plus V of r psi is equal to E psi.
Now, you could be a little concerned doing operators and say, well, am I sure this l squared is
to the right of the r squared? r and l-- l has momentum. Momentum [INAUDIBLE] with r.
Maybe there’s a problem there.
But rest assured, there is no problem whatsoever. You realize that l squared was all these
things with dd phetas and dd phis. There was no r in there. It commutes with it. There is no
ambiguity.
We can prove directly that l squared commutes with r, and it takes a little more work. But
you’ve seen what l squared is. It’s the dd thetas and dd phis. It just doesn’t have anything to
do with it. So now for the great simplification.
You don’t want any of your variables in this equation. You want to elicit to a radial equation. So
we try a factorized solution. Psi is going to be-- of all the correlates-- is going to be a product a
purely radial wave function of some energy E times a Ylm of theta and phi. And we can
declare success if we can get from this differential equation now a radial differential equation,
just for r.
Forget thetas and phi. All that must have been taken care of by the angular momentum
operators. And we have hoped for that. In fact, if you look at it, you realize that we’ve
succeeded. Why?
The right-hand side will have a factor of y and m untouched. V of r times psi will have a factor
of Ylm untouched. This term, having just r derivatives will have some things acting on this
capital R and Ylm untouched. The only problem is this one. But l squared on Ylm is a number
times Ylm. It’s one of our eigenstates.
Therefore, the Y’s and m’s drop out completely from this equation. And what do we get? Well,
you get minus h squared over 2m 1 over r, d second, dr squared, r capital RE minus l squared
on psi lm-- or Ylm now-- is h squared times l times l plus 1. So the h squared cancels. You get
l times l plus 1 r squared, and then we get the RE of r times the psi lm that has already-- I
started to cancel it from the whole equation.
So I use here that L squared from the top blackboard over there, has that eigenvalue, and the
psi lm has dropped out. Then I have the V of r RE equals E time RE of r. So this is great. We
have a simplified equation, all the angular dependencies gone.
I now have to solve this equation for the radial wave function and then multiply it by a spherical
harmonic. And I got a solution that represents a state of the system with angular momentum l
and with z component of angular momentum m. The only thing I have to do, however, is to
clean up this equation a little bit. And the way to clean it up is to admit that, probably, it’s better
as an equation for this product.
So let’s clean it up by multiplying everything by r. dr squared of little r RE. If I multiplied by r
here, I will have plus h squared over 2m l times l plus 1 over r squared rRE plus V of r times r
times RE equals E times r times RE of r. So we’ll call u of r r times RE of r, and look what we’ve
got. We’ve got something that has been adjusted, but things worked out to look just right.
Minus h squared over 2m d second, dr squared u of r plus-- let me open a parentheses V of r
plus h squared over 2m r squared, l times l plus one u of r is equal to E times u of r.
Here it is. It’s just a nice form of a one-dimensional Schrodinger equation. The radial equation
for the wave function dependents, a long r, has become a radial one-dimensional particle in
that potential, in which you should remember two things. That this u is not quite the full radial
dependent. The radial dependent is RE, which is u over r.
But this equation is just very nice. And what you see is another important thing. If you look at
the given particle in a potential, you have many options. You can look first for the states that
have 0 angular momentum-- l equals zero-- and you must solve this equation.
Then you must look at l equals 1. There can be states with l equals one. And then you must
solve it again. And then you must solve for l equals 2 and for l equals 3 and for all values of l.
So actually, yes, the three-dimensional problem is more complicated than the one-dimensional
problem, but only because, in fact, solving a problem means learning how to solve it for all
values of l.
Now, you will imagine that if you learn how to solve for one value of l, solving for another is not
that different. And that’s roughly true, but there’s still differences. l equals 0 is the easiest
thing.
So if the particle is in three dimensions but has no angular momentum-- and remember, l
equals 0 means no angular momentum-- it’s this case. l equals 0 means m equals 0. l squared
is 0. lz is 0. This is 0 angular momentum.

L21.3 Effective potential and boundary conditions at r=0 (14:28)

MITOCW | watch?v=_XDm2cxC-UU
PROFESSOR: There is an effective potential, as it’s described here. Yes, there is a potential, the central
force, but if there is angular momentum, it’s like a centrifugal barrier. With angular momentum
it becomes very difficult to reach the origin. Because if you want to reach the origin, you have
to spin faster and faster. So it becomes very hard to reach the origin.
And this is like a repulsive potential here. So you have an effective potential. You started with
just a potential, the central potential, but that will become an effective potential. This whole
thing in brackets over there is sometimes called the effective radial potential. And it’s this V of r
plus h squared over 2mr squared l times l plus 1.
So if you have a Coulomb potential, that would try to make the electrons go all the way to the
proton. This is attractive. Well, even that is not problematic for l equals 0. But for l different
from 0, you have to add here a potential that diverges. And it’s positive, say, for some l
positive and falls off very fast because forms of like r squared, this, the Coulomb potential falls
off at 1/r. So when you combine the two, so this is just the centrifugal barrier, but you combine
the two, this diverges faster as well, so the total potential is something like that in between.
And then you can have bound states. And all the theorems we’ve learned about bound states
and eigenstates of one-dimensional potential, now you discover that they’re very useful in
three dimensions. Many things just carry through.
So this is effective potential. And it’s going to recap and remember that the solution that you’ve
written is Re. But Re is u of r over r Y lm of theta and phi.
And one more thing that I want you to realize, does the function u depend on l? Yes. The
differential equation depends on l. Does the function u depend on m? Yes or no?
STUDENT: [INAUDIBLE]
PROFESSOR: No. There’s no m in the differential equation, so that should be good enough. So what’s
happened is the m dependence is kind of very simple, always very simple. It is the e to the im
phi and nothing else. The u doesn’t know about it.
On the other hand, the u depends on l because it shows in the differential equation. So I could
write here of e and l because it depends on l, and it is the solutions of the Schrodinger
equation in one potential. So there will be quantization of energy, or there might be stationary
equation in one potential. So there will be quantization of energy, or there might be stationary
states that depend on the energy. So this is the function that knows about l, knows about the
energy. And we’ve been totally successful with the angular dependence. Yes.
STUDENT: That should be theta and phi, dy?
PROFESSOR: Yes, theta and phi. Thank you. Good.
Normalization, the last thing that has to work out nicely. Let’s try to see what does the
normalization say about this function.
Well, we should find that the integral of psi squared d cubed x is equal to 1. But that integral,
as you now know, it’s the integral of r squared dr from 0 to infinity. And how do you integrate
over volume? You integrate over r. That has the right units. And you integrate over solid angle,
d omega.
Of psi squared, so let’s do the arithmetic here. We have a Y lm star of theta and phi, a Y lm of
theta and phi. We have a u squared and then r squared. Poor graph. But anyway, you can
read it.
Now, what happens is just good stuff. r squared cancels. And this solid angle angle integral is
a perfect integral for our normalization. So this gives you 1. And therefore the end result for all
this integral is just the integral from 0 to infinity of dr u of r squared. And that must be equal to
1.
So not only is little u a nice variable that satisfies a one-dimensional Schrodinger equation, but
you can remember that your more complicated wave function will be normalized if u is
normalized in the one-dimensional sense. If u squared integrated over x is equal to 1, yes,
you’re normalized.
So the set-up to convert the three-dimensional differential equation into a one-dimensional
differential equation has been very successful. We’ve reduced it to a one-dimensional
problem. We have to solve those. Each time somebody gives you a spherical potential, you
look at that equation, the radial equation, and try to solve for u’s. If you solve for u’s, then you
can append the angular dependents that correspond to angular momentum eigenstates. We
call this angular moment eigenstates. They are the most you can ask from angular
momentum.
And then you have many solutions. So I want to conclude that part of the analysis by
mentioning something about solutions with the appropriate boundary conditions. We have x, in
the one-dimensional problem we call it x. And you’d run from minus infinity to infinity.
The one difference here is that you have r, and r runs from 0 to infinity. So you may wonder if
you have some issues with r going to 0. What should the wave function do when r goes to 0?
OK. Then the way we think about it is not completely general but is good enough. We think of
the differential equation as we have here and imagine a potential that when r goes to 0, the
centrifugal barrier dominates. So our potential of the form 1 over r to the fourth would be even
more singular than the barrier. And I don’t know what happens in that case. Most likely it’s no
good, not interesting. You cannot find solutions.
But if the potential, like the Coulomb potential, is weaker than the centrifugal barrier as r goes
to 0, the centrifugal barrier dominates when r goes to 0. And this differential equation, as r
goes to 0, has a potential infinite term which corresponds to the centrifugal barrier. So we think
of this as r goes to 0, the differential equation roughly becomes minus h squared over 2m d
second u dr squared plus h squared l times l plus 1 over r squared of u, roughly 0. At least the
leading behavior of these things should work out correctly.
So the h squareds-- here is a 2m, as well, I’m sorry. The h squared over 2m’s cancel, and you
get d second u dr squared plus-- no, is equal to l times l plus 1 over r squared u. And we’re
only interested for this as r goes to 0.
And it’s not an exact statement. It’s a discovery. We’re trying to discover what’s happening with
the wave function near r equals 0.
Well, this has two kinds of solutions. You can try a polynomial. So this could go like r to the l
plus 1. If you take two derivatives, that works out. You get the l plus 1 and l. And it solves the
equation. Or r to the minus l also solves the equation.
As you can imagine, this is going to be problematic in general. It’s too divergent. It will not be
possible to normalize it, in general, for arbitrary values of l.
So this is a very brief analysis. I’m not going into all the detail that probably this deserves at
this moment. But this is ruled out, and this is ruled in.
The only thing-- so it’s true that this one is ruled out, and it has problems for normalization. It is
too divergent as l. But for l equals 0, it’s not divergent. But for l equals 0, there’s another
reason why this is not good. It turns out that for l equals 0 this doesn’t quite solve the
Schrodinger equation, the exact Schrodinger equation.
So the bottom line of this analysis is that we will have u of r behave like r to the l plus 1 as r
goes to 0. And in particular, when l is equal to 0, u of r will behave like r. So it will vanish as r
goes to 0.
So the only question is how fast it vanishes. It vanishes as r goes to 0 for l equals 0. It
vanishes even faster for higher l. So it always vanishes, the wave function at r equals 0. And
that’s why we can usually think of it as having an infinite barrier.
The wave function could not exist for r less than 0. That physically doesn’t exist. And the
boundary conditions are such that that cannot happen.
All right. We’ve finished the discussion of the radial equation. Now, our main interest with the
radial equation, of course, is, at this moment, the hydrogen atom. So we’re going to turn to
that.

L21.4 Hydrogen atom two-body problem (25:04)

MITOCW | watch?v=7q32Wnm4dEw
BARTON
ZWIEBACH:
Hydrogen atom is the beginning of our analysis. It still won’t solve differential equations, but we
will now two particles, a proton, whose coordinates are going to be coordinate of the proton,
subbing Xp for the proton, and momentum of the proton. And there’s an electron. And there’s
the coordinates, the three coordinates of the electron, and the three components of the
momenta of the electron.
And these are your canonical variables. This means that the components of this object satisfy
the standard commutation relations. That is-- I have to write it like the following. They
considered the coordinates of the proton, the i-th component. And the momentum of the
proton, the j-th component, that’s equal to ih bar delta ij.
You see, we used to code for its x, y, and z, and momenta Px, Py, Pz. You could’ve called it
X1, X2, X3, momenta P1, P2, P3. And in that way, you can use a parameter delta over here.
So, these are the commutation relations of X’s and P’s, but X and P for a proton. So, the
proton has its X, has its P, and it behaves like an X and P that we’ve studied.
The electron has its X, its P, and also behaves the same way as coordinates P of electron j
equal ih bar delta ij. You see, this is because x operator, y operator, z operator-- you can call
them xi operators with 1, 2, 3. The Px, Py, Pz operators are better called P sub i, with i running
from 1, 2, and 3, and x being the first, y the second, z being the last.
Then the fact that the x just fails to commute with Px and y fails to commute with Py, and z fails
to commute with Pz is xi pj equal ih bar delta ij. And this is what I’m saying here with this
notation because there are too many subscripts. There’s a P for proton, but there’s an i for the
first, second and third component. And there’s the momentum of proton, which has 1, 2, 3
components. So, those are our dynamical variables.
And then when you try to solve this, you have the issue of wave functions. What happens to
the wave functions? Should I invent a wave function for the electron and a wave function for
the proton? No. That’s not good. You shouldn’t a wave function for everybody, which is a wave
function that depends on the coordinates X of the electron. And if you had just an electron, you
would have a wave function that depends on X of the electron and a wave function that
depends on the coordinates of the proton.
And how would you normalize it? Well, this thing times d cubed Xe times d cubed Xp is the
probability, dP to find the electron around the little cube about the point, Xe, and the proton
around the little cube around Xp. And then if you want to see what is it probability to find the
electron at some point regardless of the proton, you integrate over the whole proton.
And then you should just have a wave function. So, if you integrate this psi squared-- oh, I’m
sorry, psi is squared. If you this psi squared over the proton, you’re left with some probability
density for the electron. So, this is the good thing about having more and more particles. You
still have just one wave function and one Schrodinger equation. You have more particles, and
it’s just one wave function.
That’s why people eventually talk about the wave function of the universe. I say, well, the
whole universe has one wave function for every, for every molecule, for every elementary
particle. There’s just one wave function. That’s nice about the Schrodinger equation.
And, OK, so we have this, and what is the Hamiltonian? The Hamiltonian is the kinetic energy
associated to the proton plus the kinetic energy associated to the electron plus the potential
energy, which just the depends on the distance between the electron and the proton.
And then and write the differential equation, the Schrodinger equation. And the P of the proton
would be thought as ddx of the proton. And the P of the electron would be like ddx of the
electron. And that differential equation would make sense.
So what is our problem showing that we have-- can think in some way as a potential or a
Schrodinger equation that the most-- just some distance that separates the particles, and
some equation for the rest. So, we have to change variables.
So what is the change of variables? I’m going to motivate some of that. Well, we have two
pairs of canonical variables. We have the proton canonical variables, X and P, of the proton,
and the electron canonical variables, X and P, of the electron. So how can we have different
kind of variables that will lead to a more interesting thing?
We can imagine that maybe we should define a new X that is the difference for which the
magnitude of that X will be what we call r. And that’s true, but it’s not so easy to do that from
the beginning, so let’s try to do what most people would do from experience and say, I know
there’s something simple about a system of doing the wrapping values, center of mass. If the
center of mass moves with constant velocity, that’s what it does always.
So, if classical notions of how you treat two-body systems can be used in a quantum setting,
we should be able to define a new quantum coordinate associated to the center of mass, and
a new quantum momentum associated to the center of mass.
And the center of mass momentum is always the sum of the momenta, so we will define at P,
a capital P-- I hope that my P’s are going to be always clear. They’re not supposed to have
before this one, the little bar below. This is the total, P, and it’s going to be defined as the
momentum of the proton plus the momentum of the electrons.
And now, I want to define an X that goes with this P. And by that, I mean that this X must have
a commutation relation with this P that tells, yes, you are an X because X with P should be ih
bar. So, whatever I define here, that should happen. Now, you could imagine that we usually
do something like this-- we multiply each particle according to its mass. Pe.
So we’re weighting the momenta associated to-- oh, I’m sorry. The positions, Xp, Xe,
associated to the mass. And this, of course, would not have the right units. It doesn’t have
units of coordinate, so this is the center of mass. So, this will be center of mass quantum
variables.
And in order to have units, you must divide by a mass, so it’s not too unreasonable to put the
sum of masses here. That’s how you define it anyway in classical mechanics. And now, we
must ask is X with P the i-th component of the j-th component. Is it really equal to ih bar delta
ij?
And the answer, I would say, is yes because, here, let’s do this one. You will have mp Xpi plus
me Xei over mp plus m e. You see, once you do one, all the rest you should be able to do just
without thinking.
P, on the other hand, is P of the momentums of j plus the momentum of the electron, so j.
What is it equal to? And then you say, look, from Xp with momentums of P, yes, they give me
ih bar delta ij, but there will be ih bar. The first factor there will be an extra mp over mp plus m
e. So this contribution comes from the first term that only talks to the momentum of the proton.
Doesn’t talk to the momentum of the electron.
And the second term for the coordinate of the electron only talks to the momentum of the
electron. And it does give you an ih bar delta ij, but this time will be with an m e over mp plus
m e. And this factor is equal to 1. So, yes, we can box this. This is a pair of quantum
mechanical canonical variables. They have the right commutation. They have the right units,
the right commutator, the right everything.
So next time when you see this, you will say, X with P-- you say, Xp with P is 1. Xe with P is 1.
mp plus-- 1. Yes, it works. You forget ih bar delta ij. This is more clear that it will work. Try it.
Next, we need the second pair of canonical variables, so-- well, actually I should use this one
for this. So for the second pair, it’s reasonable to use the X that we anticipated, so we’ll have a
relative coordinate, X. And I don’t know how I call it-- X relative, or-- no, just little x. Little x will
be defined as X of the electron. I think that’s coordinates. X of the electron minus X of the
proton. This is natural. We want a little x like that.
Already at this point, you could be paralyzed with fear. Something could have gone wrong at
this moment. Imagine if this x doesn’t commute with these guys. Then it’s all a disaster
because this should be electron and proton commute with each other. I should have written
there, not only those are the commutation relations, everything else is 0. Any X of electron with
a momentum of a proton is 0. That’s why they’re two independent pairs of canonical variables
that we can treat. It better be that this is an independent pair of canonical variables, so it better
be that this x and whatever P I’m going to define here commute with these guys.
And as far as this x, happily it works out because X’s commute with any X’s here, so this little x
definitely commutes with the little x. But the fact that this commutes with P could have killed it,
but it doesn’t because of the minus sign. One of the commutators of the little x with the capital
P would give for the electron an ih bar, and for the proton, a minus ih bar. And they cancel. So,
all is good so far.
So, this is so far so good-- commutes with capital X and capital P. So now, we have here
something, and we could put that number, alpha Pe minus beta P of the proton. And we all
know what alpha and beta are. But now you can more or less be confident. I need that this
momentum with this x give me 1, ih bar. So, that would put the condition on alpha and beta.
In fact, the condition that xi with P bar j give you ih bar delta ij, you can imagine what it is. It’s
that alpha plus beta is equal to 1. And the condition that this momentum commute with the
center of mass position, which it can fail to do so-- what can it give? Well, W alpha of the P of
the electron goes with m e, so it’s something like alpha m e minus beta mp is equal to 0.
And please-- oh, this commutator should be 0. Please make sure you know how to do this.
You can get those conditions. I’m going maybe a little fast for you to just follow it up. And so, at
this moment, we can solve for alpha and beta-- two equations. So alpha is equal to mp over
me plus mp. And beta is equal to m e over m e plus mp.
And we have a pair of canonical variables as well here, therefore, the relative coordinate and
the relative momentum, in some sense. It’s useful to define two symbols. So, we use mass m
e mp over mp plus mp. And the total mass, which is m e plus mp. In the case of a heavy
proton, the reduced mass-- the proton is heavy compared with the electron. You can ignore
the electron here, cancel them, e. And the reduced mass is called this. So, in terms of the
reduced mass, alpha is equal to mu over m e, and beta is mu over mp. Those are unit-free
constants.
So summarizing, the second pair of canonical variables are Xe minus-- Xe minus Xp. And P
equals mu Pe over m e minus P sub p over mp.
All right, so-- so, the last step that we’re going to follow, and after that, we’ll get the
Hamiltonian and stop there. And we’ll discuss solutions next time. But finally, we have enough
equations to solve for the momenta we have in the Hamiltonian in terms of the center of mass
momenta and the relative momenta.
So, from equation star and double star here, you can solve for these momenta. So, P of the
proton will be equal to mp over M, big momenta minus little momenta. And P of the electron is
m e over capital M, big momenta plus little momenta.
Well, one more equation that we have to write, and that’s the key to the simplicity of all what
we’ve done. At this moment, we have a very good physical insight into what we’ve done, but
we want to see if the math collaborates. And the good physical insight has been that center of
mass motion should be independent of the relative motion, and of we have the right X.
So, the thing that we have to compute is the first two terms in the Hamiltonian, P of the proton
squared over 2m of the proton plus p of the electron squared over 2 mass of the electron. So,
I have 1 over 2m of the proton, and I have m of the proton over capital M, P minus little p, and
plus 1 over 2m of the electron-- this is squared-- m of the electron over capital M, big P plus
little p.
And the plus and minus are extremely reassuring because the cross terms that couple the two
are going to vanish. You can see cross terms will have a minus sign. mp will be canceled. m e
will cancel. They will cancel. So at the end of this little calculation that takes couple of lines,
you get 1 over 2 capital M total center of mass of this squared plus 1 over 2 mu little
momentum squared.
This is a kinetic operator. And then you say, great success, the kinetic energy now has
separated into a center of mass contribution and a relative contribution. This will allow us to
now with a little step, separate the total Schrodinger equation into center of mass motion and
relative motion, and the relative motion will have [INAUDIBLE] potential. So we’ll do the
punchline next time.

  • 19
    点赞
  • 26
    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