Sunteți pe pagina 1din 18

847: FUNDAMENTALS OF PHYSICS II

Lecture 20 - Quantum Mechanics II [April 5, 2010]


Chapter 1: Review of Double Slit Experiment Using Electrons [00:00:00]

Professor Ramamurti Shankar: Well, this is just informal discussion till everybody's in here. So any
questions on the subject? What?
Student: Every question.
Professor Ramamurti Shankar: Everything. Okay, well, you know what, you guys should stop and
ask more things as you go along, because there is just no way you could get all of this. And it's a little
strange and only by talking about it, you will at least know what's going on. There's no way to make it
reasonable. It's not a reasonable world out there. I can only tell you what it is. I take that view; when I
teach quantum mechanics, just tell the rules and say, "This is what happens. This is how we calculate
things." And whether you like the formulas or not, it's not my concern. And the fact that it doesn't look
like daily life, also not my concern, because this is not daily life. Strange things happen. But you have to
keep me informed on how much you're following and what you are understanding, at any stage. Don't
wait for this to end, because it's not something where you can go on the last day and figure everything out.
And I will try to repeat at every stage what has gone up to that point, because the whole thing is only a few
lectures, maybe six or so. I can afford to go back every time to the beginning.
But I know that it makes sense to me, because I've seen it, and I don't know how it sounds to you. I have
no clue. You know that and so you have to speak up. You can ask any question you want, and I will try to
answer you, if it's within the realm of possibility.
Okay, so what have I said so far? So let me summarize. Even if you never came to last lecture, here is what
you should know about the last lecture, okay? Here's what I said. First thing I said is, everything is really
particles, all things, electrons, photons, protons, neutrons. They are all particles, so let there be no doubt
about that. By that, I mean if one of them hits your face, like an electron, you will feel it in only one tiny
region, one spot. Electron dumps all its charge, all its momentum, all its energy to one little part of your
face. So there's nothing wavelike about that. It's not like getting hit by a boxing glove, which can hit your
whole face. An electron hits one dot, or if it's an electron-detecting screen, only 1 pixel is hit by the
electron. And into that pixel is given all the charge, all the momentum, all the energy of that electron.
That's exactly what particles do. So when you encounter an electron, it is simply a particle. So where does
the problem come in? Where does the quantum mechanics come in? It comes in when you do the famous
double slit experiment. That's the key. The entire quantum mystery is in the double slit. Part of the
resolution is in the double slit, but the rest are a little more difficult, and I'll try to tell you. First I want to
tell you what goes wrong with Newtonian mechanics. After all, if everything is a particle, what's the big
deal, what's the problem? The double slit experiment is a problem. That's what puts the nail on the coffin
for Newtonian physics, and here it is in the basic version. You've got two slits.

By the way, I'm going to call the particle the electron. They're all doing the same thing, so what applies to
one, applies to all of them. There is a source, like an electron gun, that emits electrons. In the old days,
televisions had the electron gun. And the gun emits the electrons, they go and hit the screen, they make a
little dot, and then the dot moves around, and you see your favorite show. Okay, this is the electron gun,
and the electron gun has been engineered to send electrons off a definite momentum. That you can get by
accelerating the electrons over a definite potential, and the gain of so many electron volts will turn into
kinetic energy. As for direction, if this gun is really far away to the left, in principle 1 mile, then the only
way electrons are going to go 1 mile and hit the screen is they're all basically moving in the horizontal
direction. Then you put a row of detectors in the back, which will detect electrons. Then this is slit 1 and
this is slit 2. You block slit 1.
In fact, let me say the following thing: what do we really know when we do the experiment? Once in a
while this gun will emit an electron, and we know it's emitted the electron, because it will recoil one way,
just like a gun, rifle. It will recoil. That's when we know the electron left. Then we don't know anything,
and suddenly, one of these guys says click. That means electron's arrived here. This is what we really
know. Everything else you say about the electron is conjecture at this point. You know it was here, you
know it was there. The question is, what was it doing in between? Now if you say, "Look, things cannot go
from here to there, except by following some path, I don't know what path it is." Maybe if it's an ordinary
particle, like a Newtonian particle, it will take some straight line, hit that slit, or go through that slit and
arrive here. So you might say, "I don't know the trajectory, but it's got to be some trajectory, maybe like
that, or maybe like that." So the electron takes some path and you can label the path as either through slit
1 or through slit 2.
Okay, now here is the problem. Suppose I do the experiment with slit 2 blocked, so you cannot even get
through this one, and I sit at a certain location for a certain amount of time, maybe 1 hour and I see how
many electrons come, and I get 5 electrons, with only 1 slit open. And if I move that observation point, I
get some pattern, pretty dull, looking like that, and I'm going to call it I1. That's the count, as a function of
position up and down that wall of detectors. Then I repeat the experiment with this guy closed and that
guy open, and I get another count, looks like that. Now I'm going to pick a location. These are not drawn
to scale or anything, so I'm not responsible for any of that. Maybe I'll at least show you one thing, which is
pretty important. This graph will be big in front of the second slit, which is somewhere here. It will look
like that. So I get I1 when 1 is open and I get I2when 2 is open. This is I1, this is I2. Now I'm going to open
the two slits and I'm going to pick a particular location. It doesn't happen everywhere. I'm going to pick
one location called x. Where I used to get 5 electrons per hour with one thing open, and 5 electrons per
hour with the second thing open.
Now I want to open both and ask, what will I get? In Newtonian mechanics, there's only one possible
answer to that question, and that is 10, because we've got 5 this way and you've got 5 that way. And you
open both, whoever is going this way will keep going that way; whoever is going this way will keep going
this way. They will add up to give you 10. Now I told you, some people may say, "Well, maybe it's not 10,
because with both slits open, maybe someone from here can collide with someone from there. How do you
know that will not happen?" So I'm saying, do the experiment with such a feeble beam of electrons, there's

only one electron at a time in the whole lab. It's not going to collide with itself. Then you wait long
enough, and you have to get 5 + 5 = 10. And what I'm telling you is that if you go to the location marked x,
where you've got 5 with each one open, when you open both, you will get 0. You don't get anything. That is
a great mystery.
That is the end of Newtonian physics. And I told you that something like that never happens in your daily
life. I gave an example with machine guns. This is a machine gun. This is a concrete wall with 2 holes in it,
and there's some target here, you. And then you see how many are coming through this and how many are
coming through that. Then you go there and you wait. And both are open. Somehow, nothing comes. With
the second hole in the wall, you are safe. With one hole in the wall, you're not safe. That can never happen
with bullets. So these electrons are not following any path, because the minute you commit yourself to
saying it follows one path, either through this one or that, you cannot avoid the fact that with both of them
open, the intensity with 1 + 2 has to = I1 + I2. That's the Newtonian prediction. And I1 + I2 gives you 5 + 5
= 10 here and you get 0. In other places, instead of 10, you will get 20. Some places you get more, but
more dramatic thing is where you get less, where you get nothing.
Therefore you abandon the notion that electrons have any trajectories. You don't want to abandon it, but
you have to, because that assumption, which is very reasonable, just doesn't agree with experiment. Then
you say, "Okay, what should I do? Newtonian mechanics is wrong. What's going to take its place?" To find
that, you have to move away from this x and move up and down this row here and see what you get, and I
think I told you what you get. You get a pattern that looks like this. So the real I1 + 2 looks like this: it's got
ups and downs and ups and downs. And let's say these downs really correspond to 0. That means nobody
comes here, a lot of them come here, no one comes here, a few come here and so on. That's what you find.
I'm just telling you what happens when you do the experiment. So you put yourself in the place of a
person who did the experiment. You thought of moving away from that point x and you plot it, it makes no
sense in the language of particles.
But this is such a familiar pattern. If you're a trained physicist, which you guys are, you will say, "Hey, this
reminds me of this wave interference, with water waves or sound waves or any waves. " Obviously there's
some wavelength. The minute you give me wavelength and a slit separation, I can calculate this
pattern. dsin

= l and whatnot, and from that sin, where you get a minimum or a maximum, and if

there's a certain separation to the screen, you can find the precise location of these maxima or minima. Or
given the maxima and minima, you can work back and find the wavelength. And the wavelength happens
to be some number called , which is 10-34 joule seconds, divided by the momentum. p is the momentum.
In other words, you find that if you send more energetic electrons, accelerate them through bigger voltage,
increase the p, l goes down, the pattern gets squeezed. You slow down the electrons, p reduces, l increases,
the pattern spreads out, and the dependence on momentum is inversely proportional to momentum. And
you fool around and find out the coefficient of proportionality, which people used to call h in the old days
is now written as 2px, but it doesn't matter, it's some constant. And the number is 10-34.
So you can successfully reproduce this pattern, but what does it tell you about what's going on? What good
is that pattern? The pattern tells you that if you repeat the experiment with this electron gun a million
times or a billion times and you plotted the histogram patiently, the histogram will eventually fill out and

take this shape. So this wave is not the wave associated with a huge stream of electrons. A single electron
in the lab is controlled by this wave. You need a whole wave for 1 electron, so it's obviously not a wave of
electrons. It's not a wave of charge, like the wave of water. It's a mathematical function and you are drawn
to it, because the only way you know how to get this wiggly graph is to take something with definite
wavelength and let it interfere. So you're forced to think about this wave. And the intensity of the wave,
the brightness if you like, the square of its height, gives you what? Gives you the graph you will get if you
repeat the experiment many times. And what does it mean for the individual trial? What does it mean for
the millionth + one electron? For the millionth + one electron, it gives you the odds of where it will land
on that screen, okay?
You can never tell exactly where it will be. You can tell what the odds are, and the only way to test the
statistical theory is to do the experiment many times. And if you do it, it works, and it seems to work for
everything, for electrons, for protons, for photons, whatever it is, the wavelength and momentum are
connected in this fashion. So this wave is forced upon us, and it gives you the odds of finding the electron
somewhere. And we say that the probability I'll be a little more precise in a minute on what I mean by
the probability to find it at a location x, but let's just say, if you draw the graph of Y2, wherever it is big, the
probability is larger; wherever it is small, probability is small; wherever it is 0, probability is 0. So there
seems to be a function whose amplitude or whose square gives you the probability. That function is called
the wave function, and we know it exists, because it's the only way to calculate the result of this
experiment.
Once you tell me that the fate of a particle is controlled by a wave, you're immediately led to some other
conclusions, so I'm going to tell you what they are.
First conclusion is this: if I make a single slit, let's call this the x direction, let's call this the y direction,
and I'm sending a bunch of particles in the x direction with some momentum p0. In Newtonian
mechanics, I can manufacture for you an electron of known momentum and known position, or known to
arbitrary accuracy. If Dx is the uncertainty in my position, and Dp is the uncertainty in momentum, I can
make each of them as small as I like. So here's an actual practical way to prepare such a state. If I say,
"Give me an electron of known position, known momentum," here's what I will do. I will take a slit with a
very tiny hole in it. The width of that hole is d, and whoever comes out on the other side, what can I say
about that particle? Its position has an uncertainty of order d, because if it was not I'm sorry, Dy now,
because this is the y direction. Right? Anything who came out of hole right after it had to have a y
position, whose known to within d. It's got to come anywhere within the slit, but that's it. So that is how I
have prepared for you, that's how we filter electrons of definite position. And you can make Dy as small as
you like by making the slit as thin as you like. What about its momentum? If you had a momentum p0 in
the x direction, no momentum in the y direction, therefore py was strictly 0, no uncertainty. Dpy is 0 and
the fellow I catch here is moving horizontally with that momentum p0 whose y position is within d, and I
can make the d as small as I like. This is Newtonian physics. But we have now learned that the fate of the
particle is not in its own hands. It's contained in this wave. So what should I do in this context to find out
what it will do? Any idea what I should do to find out what will happen in this experiment, given what we
learned? Yes?

Student: [inaudible]
Professor Ramamurti Shankar: Right, but how will I calculate what will happen in this experiment?
What will decide? Yes?
Student: [inaudible]
Professor Ramamurti Shankar: Yes, this is one hole, and a light beam is coming from the left I
mean, a wave is coming from the left, and if the particle has momentum p0, it's got some wavelength,
which is 2p/p0. But if you want to know what will happen on the other side of this slit, I have to find the
fate of that wave. Yes, you can put a screen, but what will I see on a screen? Will the wave just hit this
region? You know it will spread out from diffraction. I've told you, the light will spread out. There are tiny
wiggles we ignore, and this point, where you get most of the action, that angle, , satisfies

dsin= l.

This is just wave theory. That's when you can pair up the points in the slit, in the hole, to things shifting,
differing by half a wavelength, so for every one I can find a partner that cancels it, you will get 0 here.
Beyond that, you may get a few more rises, but it's pretty much dead outside this cone. That means you
will observe this particle anywhere in this angular width.
Now a particle cannot go from this slit to there unless it had a momentum, which had a component in
the y direction. You cannot get there from here, unless you have y momentum. It's the y momentum is
uncertain to within that cone. So what's the uncertainty in the y momentum? For a vector of
lengthp0 when it gets shifted by an angle D or by an angle , it is just p0x. Or, if you like,
precisely, p0times sin. But p0sin, sin is controlled by dsin = l, so this is l/d. But l is 2p/p0 x d. You
cancel that, you find D py X d = 2p. That means D py D y is roughly forget the 2p's of our . So you
should understand this much completely without any doubt: if the future of the particle, the fate of the
particle, is controlled by the wave, you try to narrow the location of the particle by making the hole
smaller and smaller, the wave fans out more and more.
Chapter 2: Heisenberg's Uncertainty Principle [00:20:28]

That's just wave theory. People knew this about waves hundreds of years back. What is novel is that this
wave is going to tell you where the particle will end up. This wave is going to control the odds of where the
particle will end up, and the odds are pretty much concentrated in this cone, not of 0 opening angle, but
an angle , so that dsin = l. l in turn is connected to the momentum of the particle. This is where the
uncertainty principle comes in. What Heisenberg said is, "You had in your mind the classical notion that
you can have a particle of known position and known momentum. Let me see you produce that particle.
So you try to do it by putting a slit and catching guys with a very narrow range in y, but now you find out
that the momentum gets broader and broader, and that's the result of the wave associated with the
project. " It was not in the Newtonian picture. Yes?
Student: Why with the single slit, like you've drawn there, do you get the little wiggle at the outside ?
Professor Ramamurti Shankar: You mean, why would it have those wiggles? Okay, so if I look at the
single slit, I could think of many little point sources. In the forward direction, if you go very far, so you

treat them as roughly parallel, they are all in step. You've got a big maximum. In another direction where
this difference is l, dsin = l, this guy and this guy are differing by l/2.
Student: So you're assuming the slit is large enough for them to do that ________.
Professor Ramamurti Shankar: These are little mathematical dots inside the slit. You know, when
you make a slit, every point on the slit looks like a source of light, a point source of light. They are not
really light bulbs, but if you make any hole in the wall, a light comes through that hole, looks like it's a
source of light. I'm taking every point on the hole to be a source of light. And what I'm saying is, there is a
direction in which it will cancel, but if you go a little further out, it won't cancel completely. This is the
direction for perfect cancelation, where I can pair them, this one with this one, that one to that one, and so
on. They pair to give 0. But if you move further up, you no longer cancel completely, but you don't add
perfectly either, so things will get better, then they'll get worse again, and better and worse and so on. So
that's the origin of that pattern. Yes?
Student: So you're saying this is the first minimum. After that, it would increase _______.
Professor Ramamurti Shankar: Yes. For a single slit diffraction, the big thing in the middle is pretty
much all you have. It's not like double slit experiment with two holes, where you get many times the
pattern. That's because let's understand why that is true. Here, if these two differ by 1 wavelength, they
don't differ at all. I can find another angle where they differ by 2 wavelengths. They don't differ at all.
There's only 2 sources, so you can engineer them to differ by either 1 wavelength or 2 wavelength or 3
wavelengths. Here in a single slit, each point is like a source of light. You got them all to agree. You can get
them to agree only in the forward direction. In any other direction, you can get them to neutralize each
other, but never for perfect reinforcement. So you cannot get it more than this is the only real
maximum here. Everything else is tiny. Yes?
Student: Why can't you do the same trick with the double slit, where each of the one slits has its
_______?
Professor Ramamurti Shankar: No, in double slit, what's happening is, we take in double slit, the
number d I used in double slit was not the size of a slit, but the space in between the slit, okay? So there I
took the size of every slit to be vanishingly small. That means a light coming out of this slit spreads out
completely in all directions, okay? It's as if this packet became that broad. Likewise a light from this one,
we are still in the first maximum of that slit. So you've got to understand, in a single slit experiment of
diffraction, the slit size is what you are varying. In a double slit experiment, the slit is taken to be
mathematically point like, so it fans out completely. It's the interference between those two point sources
that you're adding. They can add and cancel, add and cancel, many, many times as you move along this
line.
Okay, so what we learn is it's when you combine waves and particles and go back and forth that you run
into the situation. So you cannot make a state of perfect momentum. By the way, I said one thing, I
thought about it, which is incorrect, which is, in the microscope, I said if you want to locate the position of
an electron in a microscope, take a microscope with an opening, and electron is somewhere on this line. I

said you're shining light down here. It hits the electron, but it goes in through the slit by spreading out. So
the photon that came in goes into the eyepiece with a certain uncertainty in its final angle. That means we
know the incoming momentum, but we don't know the outgoing momentum of the photon. The lens picks
up everything inside that cone. That means we don't know how much momentum it gave to the electron.
It gives an indefinite amount of momentum to the electron. Therefore the x momentum of the electron is
uncertain by that little shape, the conical shape of the momentum. And if you do the uncertainty principle
argument, you'll again find DxDp is h, h.
Now what I don't like about my experimental setup is that I had the incoming light also coming from
inside the microscope, but that means incoming light, when it comes through this hole here, will itself
spread. Then it will hit this guy, and that will go to the aperture. That will spread some more. This
uncertainty in incoming momentum is unnecessary. We can do better than that, because in other words,
when the light is picked up, it is picked up by this tiny hole. There's no reason it should also come from
the tiny hole. It can come from a source far away, say on the other side, so that it is a well defined
direction, it's not diffracting at all. So I want it to come in through a very broad hole, so it's got well
defined direction, so the light here has known momentum. It hits the electron and goes into the
microscope. It is the final momentum of the photon I don't know. And I cannot make it better. If I make it
better, I've got to open this eyepiece a lot. If I open the eyepiece a lot, I don't know where I caught this
guy. So again the problem between taking a very tiny eyepiece, so that if I see a flicker, I know the electron
was in front of it, but the light coming from the reflected electron fans out more and more.
Okay, so anyway, this is the uncertainty principle and the uncertainty principle told us something very
interesting. I asked you, what can be the function here that produces this interference pattern in the
double slit? We know the wavelength. Wavelength was 2p/p. And you know from basic physics that a
function like cosine 2px/l has got wavelength l. So let's put in the formula for l here. You get cosine 2px. l
is 2p/p. Cancel the 2 p's, you get A cosine px/. That function, when you throw it at a double slit, will
form two little wavelets, and they will interfere, that produce an interference pattern of the type you want.
Do you understand that the experiment only showed you there's a wavelength. It did not tell you what the
actual function is? That's very, very important. When Young did the experiment with the double slit, he
found the oscillations and he could read off the wavelength. It's just geometry. But he didn't know what
was oscillating. He didn't know there's an electric and magnetic field underneath all of that. But you can
always read out the wavelength without knowing what's going on. Likewise, we have the wavelength. We
know it comes from a function with a well defined wavelength, so I make my first guess to be this
function. But I told you what was wrong with this choice. You guys remember that? I said this function
violates the uncertainty principle. The uncertainty principle says if you know the position to an accuracy
Dx, and if you know the momentum to accuracy Dp, the product must be at least as big as hx some
number of order 1. We have taken the particles to have well defined momentum. If they have well defined
momentum, Dp is 0. Dp is 0, Dx is infinity. Now I told the square of the wave function is the probability to
find it somewhere, and you have no idea where it is. In other words, a particle of perfectly known
momentum has totally unknown position. So the probability should look flat, Y2 should look flat. But
the Y2, due to cosine, of course does this. It prefers some locations to another, but you're not supposed to

have any preference for any x, so we have a problem. How do I put in a wavelength into a function whose
square is flat? That's the problem we have. When you think about it, you realize your trigonometric
functions will not do the trick. If they have a wavelength, their square is not flat. The square is also
oscillating.
But then what came to the rescue is the following function, not a cosine but

Aeipx/, rather than

cosine xp/. Look at this function. This function, I've told you many, many times, if you don't know your
complex numbers, you're definitely going to have trouble. It looks like a vector of length A and angle ,
which is px/. As you vary x, this x changes and this will rotate round, but as it rotates the amplitude of
this complex number, absolute value of Y2, which is Y times Y*, which is A I'm taking A to be real here,
so A* is this, times

eipx/, times e ipx/.

That cancels out, you just get A2. In other words, the complex number describing the wave
function changes in phase but not amplitude. It's the amplitude that gives the probability. Now
there is no problem with this guy having a wavelength, because this oscillates in x. Its real part and
imaginary part both oscillate, but the square of the real + the imaginary square is 1 [A^2]. That's why the
amplitude doesn't change.
So we are driven now to the very interesting result that the wave function for a particle of definite
momentum p is this. So this is a very important lesson. Let me label this function by label p to tell you,
"Hey, I'm not talking about any old wave function." This guy has a definite momentum p. Its wave
function looks like e ipx/ times any number A you want in front of it. That's a very important thing to
know. This is called a plane wave, and a plane wave with a p right where it is describes a particle of
momentum p in particle mechanics. And I told you particles of momentum p are everywhere. Every
machine produces them, every accelerator produces them, and if you want to describe them in quantum
mechanics, you have to know complex numbers. There's just no way you can get a real answer to our
predicament. It's complex.
So that's roughly where I left you, and I want to remind you of a few other things, this further discussion
of the result we have, okay? The discussion is, if the world is really this messed up at the microscopic level,
why do I think it's the world I see in the macroscopic level? Where are all these oscillations? Why is it that
when there's a concrete wall, making another hole is bad for me and not good? Why do all these things
happen? Why do I think particles have definite momentum and position? Why do I think that if I make a
hole in the wall and I send a beam, the beam will go on the other side of the wall with a shape precisely
like the shape of the hole, no spreading out? It all has to do with the size of the object. The laws of physics
are always quantum mechanical laws, but when you apply it to an elephant, you get one kind of answer;
when you apply it to an electron, you get another kind of answer. You don't have separate laws for big and
small things. The real question is, how do these very same laws, when applied to big things by big
things, I mean things you see in daily life give the impression that the world is Newtonian? So let's look
at the double slit experiment.
Here's a double slit and we are told, "Send something. See what happens on the other side." And the
prediction is that you get these oscillations, with the peculiar property that with two holes open, you don't

get anything somewhere. We don't seem to see that in daily life, and you can ask, "Why is that so?" Well,
you remember that the condition for the next minimum is like

dsin, is l/2. So if is very small, it's

like xd is l/2, or = l/d. That's the angle you've got to go through from the central maximum. That's the
central maximum to the first minimum here. That tiny angle is given by l/d. The reason you don't see the
oscillations is when you put in the values for l and d. Let's pick a reasonable value for this angle . Do you
understand what is? In that maximum, there are some oscillations. I want to go to the first minimum
near that. The distance between these two is roughly the spacing between maxima and minima, maxima
and minima. That angle is l/2. l is 2p over the momentum and that is d. In a microscopic world, p is mv,
and let's take an object of mass 1 kilogram moving at 1 meter per second and a slit hole is 1 meter. The size
of the slit is 1 meter. Put everything equal to 1 for a typical estimate.
You find this number is 10-34 radians. That means that the angular difference between the maximum and
the minimum and the maximum and the minimum is 10-34 radians. What that means is, if you put a
screen 1 meter away, the distance between one maximum and the next maximum or one minimum to the
next minimum and so on, that spacing will be 10-34 meters, because that's how a radian is defined. If that
angle is , that distance is just the distance to the screen times . That will be 10-34 meters. That means the
wavelength of the oscillation on your screen is

10-34 meters. Can you see it? Well, make the world's

smallest detector. It's as big as 1 proton, okay. Nothing can be smaller. That's your whole detector, all the
parts, everything, 1 proton. Size of a proton is 10-15 meters. That means you will have 10 19 oscillations
inside your tiny detector. So don't be fooled by the 19. Let's take a minute to savor this.
That's how many oscillations you have, okay? You've got enough now? 3, 6, 9, 12, well, I don't have
enough time. That's a lot of oscillations. You should check the numbers though, okay? I'm saying they
typical angle will be 10-34 radians and if you put the screen 1 meter away, the spacing will be 10 -34meters,
and you look at it with an object, a detector, whose size is 10-15 meters, which looks very small, size of a
proton. But look, 1019 fit into that length, so your proton detector looks huge. In fact, I cannot even show it
here. So you don't see the oscillations; you see the average only. If you see only the average, you can show
that with 2 slits open, the sum is the sum of the two averages, so you don't see the oscillation. That's the
first reason. Now you can say, "You took a kilogram. Let me take a gram." I said, "Go ahead. Take a gram,
take a milligram and take a slit which is not 1 meter wide but 1 meter apart, but 1 millimeter apart." It
doesn't matter. You're playing around with factors like 10 and 100 and 1,000. I got 10 19 here. So nothing
you do will make any dent on that. So in the macroscopic world, you will not see this interference.
Another reason you won't see it is that the particle should have a definite momentum. It's
got an indefinite momentum, it's coming in with different momenta, then each will have its
own interference pattern and they'll get washed out. Finally, I told you, if you ever try to see
which slit the particle took by putting a light beam here, the minute you catch the electron going through
one slit or the other, this pattern is gone. It will do this "I'm not here and I'm not there" routine only if you
never catch it being anywhere. That's very interesting. The electron behaves like it does not go through
any one particular slit, as long as you don't catch it going through one slit. You put enough light to catch
every electron, then you can add the numbers and you must get the sum of the two numbers. Now for the
atomic world, it's possible for the electron to go for a long time without encountering anything, and the

interference effects come into play. In a macroscopic world, there is no way a macroscopic object can
travel for any length of time without running into something. It will run into other air molecules. It will
run into cosmic ray radiation. It can collide with black body radiation from the big bang, anything. The
minute you have any contact with it, this funny thing will disappear. So that's one reason you don't see it.
Now we can go on and on and give other numbers. I've given examples in my notes, which I will post later
on. One of them is the uncertainty principle. Why does it look like the uncertainty principle is not
important? Take again, this is 10-34. Everything is in MKS units. So take an object of mass 1 kilogram
whose location is known to the accuracy of 1 proton. Okay? So this number is 10-15 meters, do you
understand? You take an object made of 1023 protons and you know its location to the width of 1 proton.
That's all you don't know about its location. That's your Dx. What's the Dp? Well, Dp is now 10 to the,
what, 19? 10-19. Now Dp is m times Dv. That's 10-19. If this is 1 kilogram, Dv is 10-19meters per second. You
don't know its velocity to 1 part in 10 19. Now how bad is that? Well, suppose I start a particle off exactly
known velocity, I know where it will be forever. But suppose I don't know the velocity to this accuracy, and
I let it run for 1 year. So I don't know precisely where it is, but how bad is it? How badly do I not know?
Well, 1 year is 107 seconds, so if it runs for 1 year, it will be unknown to 10-12 meters. 10-12meters is what,
let's see? It's 1/100 of an atom size, 1/100 the size of an atom.
So you see, these uncertainties are not important in real life. So everything that you think has a definite
position and momentum actually has a slight uncertainty, but the uncertainties don't lead to any
measurable consequences over any distances that you can actually have. So what I'm trying to tell you is,
there are these waves. They do all kinds of things, they do interference and everything, but the condition
for them is really the microscopic world. The minute the masses become comparable to gram or kilogram
and distances and slits and so on, or like a meter or a centimeter, these effects get washed out. But in the
atomic scale, they are seen.
Chapter 3: The Probability Density Function of an Electron [00:42:34]

Now the final thing I want to mention before moving on to a completely new topic is the role of
probability in quantum mechanics. We have seen that quantum mechanics makes probabilistic
predictions. It says if you do the double slit experiment, I don't know where this guy will land, but I'll give
you the odds. Okay, now that looks like something we have done many times in classical mechanics. For
example, if you have a coin and you throw the coin, you flip it and you say, "Which way will it land?" well,
it's a very difficult calculation to do, but it can be done in principle, because a coin, once released from
your hand, can only land in one way. That's the determinism of Newtonian mechanics. If you knew exactly
how you released it with what angle or momentum, what's the viscosity of air, whatever you want, if you
give me all the numbers, I'll tell you it's head or tails. There's no need to guess. In practice, no one can do
the calculation. What you do in practice is, you throw the same coin 5,000 times and you find out the odds
for head or tails and you say, "I predict that when you throw it next time, it will be .51 chance that it will be
heads." That's how you give statistical predictions. Now you did not have to use statistics. You use it,
because you cannot really in practice do the hard calculation. In principle, you can. Secondly, if you toss a

coin and I hide it in my hand, I don't show it to you, it's either head or tails, and I say, "What do you think
it is?" you'll say, "It's .1 chance that it's heads." And I look at it, it may be head or it may be tail. Suppose I
got head. It means that it was head even before I opened my hand, right? The correct answer's already
inside my hand.
I just didn't know it. I'm using odds, but when I look at it, I get an answer. That's the answer it had even
before I looked.
So I'll give another analogy here. So this is a probability of locating me somewhere. This is my home town,
Cheshire, this is Yale, and this is the infamous Route 10. So somebody has studied me for a long time and
said, "If you look for this guy, here are the odds." Either he's working at home or he's working at Yale, and
sometimes he's driving, okay? This is the probability. First thing to understand is the spread out
probability does not mean I am myself spread out, okay? Unless I got into a terrible accident on Route 10,
I'm in only one place. So probability's being extended doesn't mean the thing you're looking at is
extended. I am in some sense a particle which can be somewhere. These are the odds.
Well, suppose you catch me here on one of your many trials. If you catch me only once, you don't know if
the prediction's even good, so you repeat it. You study me over many times and you agree the person had
got the right picture, because after observing me many, many days you in fact get the histogram that looks
like this. The important thing is, every time you catch me somewhere, I was already there; you just
happened to catch me there. I had a definite location. It was not known to you, but I had it. I had a
definite location because in the macroscopic world I'm moving in, my location is being constantly
measured. You didn't ask or you didn't find out, but I'm running through air molecules. I've slammed into
them. They know that. I ran over this ant. That was the last thing the ant knew, okay? So I'm leaving
behind a trail of destruction and they all keep track of where I am. My location is well known. You just
happened to find out.
But now let's change this picture and say this is not me. This is an electron and it's got two nuclei. This is
nucleus 1 and this is nucleus 2. It can be either near this nucleus or that nucleus, and this is the Y2 for the
electron. That means it's the probability you'll catch it here and you'll catch it there. Once again, if you
catch the electron, you will catch all of it in one place. It is wrong to think the electronic charge is
somehow spread out around the atom. It's not true. The charge is in one place. The odds are spread out.
That looks just like my case. But the difference is, if you catch the electron let's in fact simplify life and
say there are only 2 possibilities. Either it is near atom 1 nucleus 1 or nucleus 2, only 2 discrete choices.
If you catch it near 2, it is wrong to think that it was there before you got it. So where was it? It was not in
any one place. It had no location till you found its location. That's very strange.
We think of measurement as revealing a pre-existing property of the object. But in quantum theory, it's
not that you don't know where the electron is. It does not know. It is not anywhere. It's the act of
measurement that confers a location or position on the electron. That state of being, where you can be
either here or there, or simultaneously here or there, has no analog in the classical world. If anybody tries
to give you an example, don't believe it, because there are no examples in the macroscopic world that look
like this. No analogies should satisfy you, because this has no analog in the real world, okay? So this is the

interesting thing in quantum mechanics. If this is a possible wave function Y, electron near nucleus 1, and
that's a possible wave function Y, electron near nucleus 2, you can add the two functions. That's another
possible function. But what does that describe?
It describes an electron which upon measurement could be found here and could be found there. It's not
like finding me in Cheshire or finding me in New Haven, because in those cases, on a given day on a given
measurement, you can only get one answer, depending on where I am. Right now if they look for me, they
can only find me here. They cannot find me anywhere else. But in the case of the electron, the one and the
same electron, on a given trial, at a given instant, is fully capable of being here or there. It's like tossing a
coin and it's in my hand. We all know that when I reveal it to you, you can only get one answer, now that
the toss has been done, it's got one answer. If it's a quantum mechanical coin, you don't know, and it
doesn't have a value till you look. When you look, it has a definite value. Before you look, it doesn't have a
definite value. That's exactly like saying, when you looked, it goes through a definite slit. When you don't
look, it's wrong to assume it went through a definite slit. Yes?
Student: Say you had a double slit experiment but instead of having a screen that went all the way in
both directions, you just sort of had _____ screen. So then you would only be looking at the final location
of a _______ electron. The other half you would know. How would that work, because location for some
of them has to be ____________.
Professor Ramamurti Shankar: The minute you find the location to the accuracy of knowing which
slit it went through, you've got 2 slits or only 1 in the experiment?
Student: You have 2 slits but only a half screen, and nothing on the other one.
Professor Ramamurti Shankar: Oh, you've got a screen that only comes to here, you mean?
Student: Yes.
Professor Ramamurti Shankar: Yes, the real problem of location that I'm talking about here is not
when it hits the screen, but here, when you try to see which hole it went through, by putting a light source
here. I was referring to the fact that if you have the right kind of light to know which hole it went through,
if you give it enough momentum to wash out the pattern. As far as the screen goes, once it comes out, it's
the sum of these 2 waves coming from the 2 holes, and it also doesn't have a well defined position. The
probability for finding it may look like this. In fact, the probability will look like this. Forget your screen.
This is the probability. The minute you catch it, it is found there, that's the location after that
measurement. Prior to the measurement, it can really be anywhere where the function is not 0. There are
many wave functions. There's the 1 to the left of the slit and there is 1 then it becomes 2 wave functions
coming from the 2 slits. They form the interference pattern and that gives you the odds that if you looked
for it, you will find it.
Now till you find it, it's not anywhere. It can be anywhere on this line at that instant. It's only the act of
measurement, or hitting a detector that tells you that's my location. So you will have to get used to that.
You'll have to get used to the fact that things don't have position, momentum, angular momentum, energy

or anything, until you measure it. Okay, so I've got to tell you a little more now about just position. So let's
take by the way, any questions so far? Yes?
Student: Can you explain again how you can tell which hole the electron goes through with the light?
Professor Ramamurti Shankar: Well, you just see it. You see a flicker and if it's near this hole, you
know this guy went through that. And if it's near that one, you know it went through that. But to have such
good resolution, the wavelength should be much smaller than the space in between the slits. Otherwise
you'll get a big blur and you won't know which one it went through. That's a soft measurement that
doesn't do you any good. It won't destroy the pattern. That's because you don't know which hole it went
through. If you make it fine enough to know which hole it went through, you will disrupt the electron's
momentum enough to wipe out the pattern. Yes?
Student: [inaudible]
Professor Ramamurti Shankar: Oh, here? You mean what happened to the 2? Oh yeah, forget the 2s.
There are 2p's I forgot, right? There's a lot of 2p's too you've got to put in. 10 -34 is not exactly the answer,
because I got 2p there. I'm just saying, look, if it's 10-34, suppose you're talking about how much money
Bill Gates has. It's 109. Now is it 2 x 109, 3 x 109? I don't know, but I'm not worried about his financial
wellbeing, because it's up there in the 109s. So whenever I do these arguments, you should get used to this
notion, it's very common for physicist when they argue in quantum mechanics, will use the symbol that's
not quite an =. It looks like a wiggle and =. Basically it means, I'm not quite sure, but the number is in this
ballpark. And if this ballpark is 1,000 miles from that ballpark, you just have to know it's in the ballpark.
You don't have to know where it is. Some things are definitely macroscopic; some things are definitely
microscopic. But something very interesting is happening at Yale right now. People are asking the
following question: how small an object has to be before I can see its quantum mechanical fluctuations?
We know that if it's like an electron, it's completely fluctuating. You don't know anything. If it's like a
bowling ball, it seems to have a well defined position. Make the objects smaller and smaller and smaller.
How small can it be before it's first beginning to show quantum effects, like quantization of energy or
quantization of momentum, depending on the problem, or quantum fluctuations in position? So those
measurements are now being carried out at Yale. It's a very exciting time and it's so many years after the
discovery of quantum mechanics. Because we knew the really big world and we knew the really small
world, but now we're trying to go, because of nanotechnology, continuously from big to that small, and
how small is small?
That's the question? Can a little macroscopic object simultaneously be in two places? Most of them seem
to have a well defined location. Can you create a situation when it's capable of being found here and found
there? It's very hard, because you have to isolate the particle from the outside world. That's the first
condition. That's what ruins everything.
A quantum computer, you must know, has got these bits called qbits and unlike the bits in your laptop,
which are either 0 or 1, a qbit can also give you 0 or 1, but it can also be in a state where it can give either 0
or 1 on a given trial. The bits in your computer, the particular bit right now is either a 0 or a 1. Maybe you
don't know it, but it can only give you that answer because that's what it is. Because that bit is in contact

with the world and the world is constantly measuring its value. A qbit is a quantum system which can do
one of two things, but it's isolated and it's neither this nor that. It's like the electron going through both
slits. So a quantum bit can explore many possibilities. If you build a computer with 10 qbits it can be
doing 210 things at the same time. And if it's got a million bits, it's 2million operations, things it can be
exploring at the same time.
That's why, as you know, one of the ways to securely send your credit card information is to use very large
numbers, on the assumption that no one can factorize them. You can always multiply a 100 digit number
by a 100 digit number on your computer instantaneously. But if I gave you the 200 digit number and told
you to find the two factors you won't find it. You won't find it in the age of the universe. It's amazing, but
that simple problem of factorization cannot be done if the numbers are 100 digits long, and that's the
reason why people openly broadcast the product, they may broadcast one of the numbers, and only the
other person knows the second number. Now there is something somebody called Shore, Peter Shore,
showed that if you have a quantum computer, made up of these qbits, it can actually factor the number
exponentially faster, namely, instead of taking 1010 seconds, it will take 10 seconds. So if you build a
quantum computer, you have two options. Either you can become famous, or get tenure at Yale, maybe, or
you can go on the biggest shopping spree of your life, because you can get anybody's credit card number.
So that's the choice. When you come to that fork, you decide which way you want to go. Maybe you can go
through both choices, I don't know. That's something in your future.
So why is it so hard to build a quantum computer? There are many, many quantum systems which can do
one or two things, and can be the state, but they are both this and that. The problem is, they cannot be in
contact with the outside world, because single contact with them is like a dream. Think about it, it's gone.
Same thing. Any measurement destroys it. Any unintended measurement also destroys it, so you've got to
keep your system fully isolated. A system that is not talking to the outside world, unfortunately, is also not
talking to you. So you cannot ask it any questions, and if it knows the answer, it cannot tell you. So
sometimes you want it to talk. It's like relationships. Sometimes you don't want it to talk. So what do you
do? You've got to build a system where sometimes, in a controlled way, you can make contact with your
system, namely give it the problem. Then it does its quantum thing, then you've got to make a
measurement to find out what the answer is. Then you want to be able to get into it again. So the
challenge for quantum computers is how to keep them isolated long enough to do the calculation. That's a
challenge, how to keep it from how to keep it in what's called a quantum coherent state. A coherent
state is really when it's doing many things at the same time.
All right, so I want to tell you now more formally how to do more quantum mechanics. So let's take a
simple example, a particle living on a line. That's the function Y(x). Let's ask the following question: how
do we do business in Newtonian mechanics? We say, "Here's a particle. That's its x. Here's the
momentum. That's its p." Given that, I know everything I need to know right now. Angular
momentum, r x p, Kinetic energy, p2/2m. Everything is given in terms of the coordinates and momentum.
In quantum theory, you don't even tell me where it is. For every possible x, there is a function whose
square, if you now square this guy, everything will be now positive, definite. I don't know, it's something

like this. This is Y2. We say the height is proportional to finding it everywhere. What is the condition on
the function side?
The answer is, whatever you like. Anything I can draw, with no special effort, is a possible function for an
electron. There are no restrictions. It's like saying, what should the position of the particle be?x. Any x you
want is a possible x. Likewise, any Y you draw is a possible Y. That's a set of all possible ampli it's called
wave functions whose square is the set of all possible probabilities. So I said Y2, Y at the point x2, is
connected to the probability of finding it at x. That has to be in fact to be refined. That's not precisely the
story. I'll tell you why. If a statistical event has got 6 possible answers, like you throw the die, you want to
get any number from 1 to 6, you can give the probability for 1, probability for 2, 3, 4, 5 and 6. These are all
the odds for getting any number from 1 to 6. Since there are only a finite number of things that can
happen, I call them I = 1 to 6, there's a probability for each I, and if you add all the probabilities, you
should get 1. But if the set of things that can happen is a continuous variable like x, in other words, the
location of the electron is not a discrete set of numbers. It's any real number is a possible location. Then
you cannot give a finite probability for any one x. If that was finite, since the number of points is infinite,
you cannot make the total probability 1.
So what we really mean by Y2 is called the probability density. That means draw the Y2, let this be the
height of Y2. Take a little sliver of width dx. The area under the graph, P(x) dx, that is the probability of
finding the electron, or whatever particle, between x and x + dx. You understand? It's called a density. So
you don't give a finite probability to each point. For an infinitesimal region, you give it infinitesimal
probability, which is the function P(x) dx. And the statement that the particle has to be somewhere,
namely, all the probabilities add up to 0, is the statement that when you integrate this probability density
from - to + infinity, namely, Y2dx, from - to + infinity, that should be 1. This is called normalization. It's a
mathematical term. Norm is connected to length in some way, and these can be viewed as length squared.
Anyway, this is called the normalization. Now if somebody gave you a Y, which did not obey this
condition, here is a Y. This already tells you a nice story, right? It tells you the odds are pretty big here,
pretty small there, 0 here. Now take a function that's twice as tall. That gives you the same relative odds,
you understand.
So when you multiply Y by any number, you don't change the basic predictive power of the theory. It is
just that if your original Y had a square integral = 1, the new one may not have, but the information is the
same. It's really the relative height of the function. That's another shocking thing in quantum mechanics.
If Y stood for a string vibrating, 2 Y (this is Y, this is 2 Y) is a totally different configuration of the
string. But in quantum mechanics, Y and any multiple of Y are physically equivalent, because what we
extract from Y is the relative probability of finding it here and there and there and there. So scaling the
whole thing by a factor, 2 or 4 or any number, it doesn't matter. That's a very new thing. That's why
the Y is not very physical. If you took a string and you pulled it by twice as much, it's a totally different
situation. If you took the electric field and made it twice as big, that's a different situation. Forces on
electrons are doubled now. In the quantum mechanical Y, when you double it, it stands for the same
physical condition of the electron, because the odds of being here versus being there are not altered. The
only job of Y is to give you the odds.

Therefore it's like saying in 2 dimensions, that's a vector, that's a different vector. But suppose you only
care about the direction of the vector. For some reason, you don't care how long it is, you just want to
know which way it points, then of course all of these are considered equal. And that's really how it is for
quantum mechanics. Every Y and every multiple of it stands for one situation only. So what one normally
does is from all these vectors in that direction, you may pick one whose length is 1 and say, "Let me use
that member of the family to stand for the situation." That's like saying of all the Y's obtained by scaling
up and down, I'll pick one whose square integral is 1.
So let me do a concrete example, so you know what I'm talking about. So let's take a function that looks
like this. It is 0 everywhere, and it has a height A between +a and -a. That's my Y. So Y(x) = A for absolute
value(x) less than a and = 0 outside. That's a possible wave function Y. Now what does it tell you in
words? If it's a word problem, what does it tell you about the electron? Can anybody tell me? What can
you say about the electron given by this function? What do you know about it? Yes?
Student: It must be found within + or -a.
Professor Ramamurti Shankar: It must be found within + or -a and more than that.
Student: [inaudible]
Professor Ramamurti Shankar: That is correct.
Student: With the same probability.
Professor Ramamurti Shankar: With the same probability, okay? The probability is it restricted to -a
to +a, and it's the same throughout the interval. After all, if you just set it restricted to -a to +a, it's true for
this function too, but that's not the same everywhere. I've got a guy who's same everywhere. Do you agree
that this function has exactly the same property, restricted to -a to +a, and the probability's constant? So
there are many, many functions you can draw, all with the same statement that this object has got equal
likelihood to be in this interval and 0 outside. Of this family, we are going to pick one guy whose square
integral is 1. So I'm going to keep this number A, the height of the function, as a free parameter, and I'm
going to choose is so that A2 so that the Y2dxfrom - to + infinity, I want it to be 1. And I'll pick A so that
that is true. Well, we can do this integral in our head. What is this integral? This is just the square
of A times the width of this region. That's got to be 1.
That tells me that A must be chosen to be 1/(2a)1/2, where this little a, 2a, is now the width of this region.
Therefore from this whole family, the normalized Y will look like 1/(2a)1/2 for mod x < or = aand 0 outside.
And we can all see at a glance that if you squared this normalized Y and integrated from -a to +a you will
get 1. This is normalization. So sometimes, people will give you a wave function and they will say as a first
step, "Normalize this wave function." What you have to do is, you've got to square the wave function and
then put a number in front of it, and choose it so that the number makes the square integral 1.
Let me give you another example. There's a very famous function, called the Gaussian function. It looks
like this. The function e-(x (squared))dx from - to + infinity happens to have an area which is square root
of p/a. That is just one of those tabulated integrals. So here's a bell shaped function with this property.

Now I want to make a quantum mechanical wave function that looks like the Y(x) = A e- x squared over 2 D
squared. That's a possible wave function, right? Nothing funny about it, but what do you know about the
wave function? It's biggest at x = 0. It's symmetric between + and -x. And it dies off very quickly, but how
far should you go? You can easily guess that when x is much bigger than D, this function is gone,
because x/D2 is going on the exponent. So if that number's big, it's e to the - big, which is very small. So
roughly speaking the width of this graph is of order D or 2D. I'm just going to call it D, just to give you an
order of magnitude. So that's an electron whose location is roughly known to an amount D. But this is not
normalized, because if I take the square of this, I won't get 1.
So I will choose A so that 1 = Y2xdx. This being real, I don't need the absolute value of Y. That gives
me A2e -x(squared)/D2dx. The 2 went away, because I squared the function Y, so don't forget that. Now I look
at the table of integrals and what is a? When I compare these 2, a is just 1/D2. So it's square root of pD2.
This is an easy thing, because I'm already giving you the integral you need to do, but I want you to get
used to it. So this whole thing should be 1. That means A is 1/pD2 to the fourth root, the power .
Therefore the normalized wave function, Y normalized, looks like 1/pD2 to the , e x(squared)/2D2.
Normalization is just a discipline. You discipline yourself to take all functions and normalize them,
because why do you normalize them? If you normalize them to 1, then Y2 is directly the absolute
probability density. That means when you add it all up, you'll get 1. If you don't normalize it to 1, Y2is the
relative probability density. It will still tell you the relative odds of this and that, but you cannot say this
interval from here to here, the chances are 30 percent for catching it. You must take the region that you're
integrating, divide by the whole thing. But you don't have to divide by the whole thing if you've
normalized it to 1.
Okay, this is just practice in normalization. So I'm going to give you a little hint on what is going to
happen next, but I won't do it now, so you guys don't have to take down anything. Just ask the following
question and we'll come back to it on Wednesday. I've told you that in Newtonian mechanics, every
particle has an x and it has a p. In quantum theory, instead we traded for a function Y(x) and we learned
the meaning of the Y(x) is that absolute value of Y2 is the probability density, meaning P(x) dx is the
probability of finding it between x and x + dx.
Now we can say, "Okay, that's enough about position. What about momentum?" I can measure the
momentum of a particle. You talked about momentum on and off in the lecture. If I measure momentum,
what answer will I get?" What are the odds for getting this or that answer? So given Y(x)that looks like
this, you square it, you get Y(P(x)). The question is, x is not the only thing we're interested in. Even in
Newtonian mechanics, x and p were equally important. What do you think will happen now? How do I
find out what happens if, instead of being interested in where I find it, I ask, with what momentum will I
find it? Can you imagine a guess on what the answer might be or in what form the answer will be given to
you? This is a wild guess. Nobody expects you to invent quantum mechanics in 30 seconds, so make a wild
guess. Yes? Anybody there want to make a wild guess? No? Go ahead, yes, you're smiling. Make a guess. I
want the odds for different values for momentum. How do you think that information will be contained in
this theory?

Student: [inaudible]
Professor Ramamurti Shankar: Pardon me?
Student: [inaudible]
Professor Ramamurti Shankar: Maybe, based on the uncertainty principle, but I want for every value
of momentum a probability, right? I want the odds of getting this p or that p or that. So what do you think
we need to get the odds for every momentum? Yes?
Student: [inaudible]
Professor Ramamurti Shankar: Pardon me?
Student: If you have a more defined location _________ calculate the probabilities of momentum,
same way we did it for location?
Professor Ramamurti Shankar: Right. So what you will need, it seems reasonable to think, that this
guy contained all the information on where you will find it, maybe there's a different function of
momentum, whose square will give you the probability density that you get if that function looks like
this and you square that, that's the odds for getting one momentum versus another momentum. After all,
every variable in classical mechanics you can measure in the quantum theory and you can give the odds.
And for every variable, it looks like you need a function. What I will show you is that you don't need
that. Y(x) itself contains information on what happens when you measure momentum, what happens
when you measure energy, what happens when you measure anything and how do you extract it is what
we'll talk about.
[end of transcript]

S-ar putea să vă placă și