MIT8.04 Lecture 2: Overview of quantum mechanics (cont.). Interaction-free measurements.

L2.1 More on superposition. General state of a photon and spin states (17:10)

MITOCW | watch?v=0xNmc2tJ-YM
We spoke about superposition, and we showed how, when you have two states that are superimposed, the
resulting state that is built up doesn’t have properties that are intermediate between the two states that you’re
superimposing. But rather, when you do a measurement, you obtain the result that you would sometimes-- you
sometimes obtain the result that you would have with one of the states, and some other times with different
probabilities, the result as if you had the other state. So it’s a strange kind of way in which things are combined in
quantum mechanics.
So the next thing we have to say is a physical assumption that is made here. And it is that if you have a state and
you superimpose it to itself, you haven’t done anything. So the superposition of a state with itself has no physical
import. So we can say this. A physical assumption superimposing a state to itself does not change the physics.
So if I have a state, this is physically equivalent-- I’ll write physically equivalent with this symbol-- to the state a plus
a, which would be 2 times a. It’s physically equivalent to the state minus a. It’s physically equivalent to the state ia
on anything. It’s not equivalent to 0a, because that would be the zero state. So it’s physically equivalent as long as
you have a non-zero coefficient. All these states are supposed to be physically equivalent.
And that will eventually mean that we sometimes choose a particular one in those collection of states that is one
that is convenient to work with. And that will be called a normalized state, a state that satisfies other properties
having to do with the norm squared of the state. That will come later. But it’s important that the number that is
multiplying the physical state of your system has no relevance.
And you could say, well, why all of the sudden you tell us this. Could this be shown to be necessary? Or it’s a
physical assumption, so can we test it? Does it make some sense? And we can make some sense of this
assumption at this level.
And we do it with states of light. So remember, we spoke about photons hitting a polarizer. And we could speak of
two independent kind of photons-- photons polarized along the x-axis and a photon polarized along the y-axis.
And those are two quantum mechanical states.
Now suppose I decide to superimpose those states to create the most general photon state. I would have an
alpha, which is a number here, a complex number, and a beta there. And I would say, OK, here is my most
general photon state. And how many parameters does this state have? It has two complex parameters, alpha and
beta, and therefore, four real parameters.
And then you think about polarization states, how many parameters they have. And as we’ll review in a second,
it’s well known that photons-- their polarization state can be expressed with just two real parameters. So some
counting is not going very well here.
But here comes the help. If the overall coefficient here doesn’t matter-- if I can change it, I can multiply everything
by 1 over alpha, and therefore get that the state is just the same, physically equivalent to this state, beta over
alpha photon y. So all the physics is contained in this state as well.
And if all the physics is contained in that state, I must look how many parameters it has. It still looks like there’s
two numbers here, but only the ratio appears. So if you call beta over alpha, the number gamma is just one
complex parameter. And therefore, thanks to this assumption, you now get that the most general photon
polarization state has just one complex parameter, or just equivalently, two real parameters. And that is the
correct number.
Indeed, if you have a polarization, a wave that has some polarization, the most general polarization state of a
wave is an elliptical polarization. You probably did study a lot about circular polarizations, or maybe you also heard
about the elliptical one in which the electric field-- in a circular polarization, the electric field at any point traces a
circle. But if you have an elliptical polarization, the electric field traces an ellipse.
And that ellipse has an angle that is one parameter. And for an ellipse, the other-- the size doesn’t matter. The
size depends just on the magnitude of the electric field. It’s not a parameter of the polarization of the wave. Since
the size doesn’t matter, it’s the shape of the ellipse that matters. And that’s characterized by the eccentricity or by
the ratio a over b of the semi-major axis, so parameters, two parameters, and they are a over b and theta.
So an elliptically polarized wave, which is the most general state of polarization of a wave, has two real
parameters. And now, thanks to this physical assumption, we get this right. And this is important because that’s
something we’re going to use all the time, that the overall factor in a wave function does not matter.
So if we have superpositions, I want to emphasize one more thing about superpositions. And for that, I’m going to
use spins. So what is spin? Spin is a property of elementary particles that says that actually, even if they’re not
rotating around some other particle, they have angular momentum. They have intrinsic angular momentum, as if
they would be made of a tiny little ball that is spinning.
I say as if because nobody has ever constructed a model of an elementary particle where you can really make it
spin and calculate how it works. Somehow, this elementary particle has angular momentum is born. Even if it is a
point particle, it has angular momentum, and it’s spin.
And spin is very quantum mechanical. And we can’t quite understand it without it. So what happens is that you can
measure the spin of a particle.
And then if you measure it, you have to decide, however, since angular momentum is a vector, what direction you
should use. And suppose you use the z direction to measure the spin of a particle. You may find that the particle
has either spin up or the particle has spin down. Spin.
And the spin is the direction of the angular momentum. And that’s a funny thing that happens with most matter
particles. These are spin 1/2 particles. The spin can be up or it can be down along the z direction that you
measure. You measure it, and you never find it’s 0 or a little bit. It’s just either up with the full magnitude or down
with the full magnitude. That is a spin 1/2 particle.
And the state where it is up, we sometimes denote it with an arrow up and call it z because it’s up along z. And this
would be down, an arrow down along z. If those are possible quantum states, you could build a new quantum
state by superposition which would be up along z plus down along z.
Now, if I wish to normalize it, I would put the factor in front of this. I will not talk about normalizations at this
moment. They’re not so important.
If you are faced with this quantum state-- so suppose you have an electron that is not in this state nor in this state,
but is in this state, in a quantum superposition. So you go and you decide to try to measure it. Now, since you
cannot predict what that electron is going to be doing-- we cannot predict things in quantum mechanics with
certainty-- we, since we’re going to do this experiment, avail ourselves of 1,000 copies of this electron, all of them
in this peculiar quantum state.
So you have the 1,000 copies, and you start measuring. And you decide to measure the spin in the z direction.
And now what do you get? Well, we mentioned last time that you don’t get an average, or since this is up and this
is down, you get 0. You measure the first particle and you find it up.
Measure the second, up, the third, up, the fourth, down, five, down. And then you get a series of measurements.
At the end of the 1,000 particles, you find about 495 up and 505 down, about half and half. And if you did it with
10,000 particles, maybe it would be closer. Eventually, you’ll find 50% in this state and find 50% in this state.
And if you think this is strange, which you probably do, well, you could be justified. But here would come Einstein
along and would say all this stuff of this superposition is not quite right. You had this 1,000 particles. But actually,
those 1,000 particles, half of them were with a spin up and half of them were with a spin down.
So here you have your 1,000 particles, your quantum state, this. But Einsten says, no, let’s make an ensemble of
1,000 particles, 500 up, 500 down, and do the same experiment. And the result is going to be the same. So how
do you know you really have this as opposed to somebody has given you 1,000 particles, 500 up, 500 down? How
can you tell?
And in fact, he would say even more-- whenever Einstein used the word realism to say if I measure a spin and I
find it up, it’s because before I measured it, the spin was up. It’s almost like learning something about an object. If
I look at this page and I find the color red, it’s because before I looked at it, it was red.
But then in quantum mechanics, that doesn’t seem to be the case. The state is this mix. And it was this mix before
you measured. And after you measure, it’s this. So there is no such thing as you learn by doing one measurement
what the state of the particle was.
So we will not resolve Einstein’s paradox completely here because we would have to learn more about spins,
which you will do soon enough. But here’s the catch that actually happens. If, instead of having an ensemble of
quantum states, you would have an ensemble of those states that half of them are up and half of them are down,
you could now decide to measure the spin of the particle along the x direction.
You take these particles, and you measure along x. And what you will calculate with quantum mechanics later in
this course-- if you measure along x, in this state, you will find all of them to be pointing along plus, up along x, all
of them. While on this Einstein ensemble of 50% up and 50% down, you would find 50% up along x and 50% down
along x.
So there is an experiment that can tell the difference, but you have to look in another direction. And that
experiment, of course, can be done. And it’s a calculation that can be done, and you can decide whether these
quantum states exist. And they really seem to exist.

L2.2 Entanglement (13:07)

MITOCW | watch?v=G3HSP3qMgKI
BARTON
ZWIEBACH:
Let’s talk now about entanglement. So we talk about entanglement when we have two noninteracting particles. You don’t need a strong interaction between particles to produce
entanglement, the particles can be totally non-interacting. Suppose particle 1 can be in any of
these states-- u 1, u 2. Let’s assume just u 1 and u 2. And particle 2 can be in states v 1 and v
2.
And you have these two particles flying around, these are possible states of particle 1 and
possible states of particle 2. Now you want to describe the full system, the quantum state of
the two particles. States of the two particles. Two particles. Well, it seems reasonable that to
describe the state of the two particles that are not interacting, I should tell you what particle 1
is doing and what particle 2 is doing.
OK, so particle 1 could be doing this. Could be u 1. And particle 2 could be doing v 1. And in a
sense, by telling you that, we’ve said what everything is doing. Particle 1 is doing u 1, particle 2
is doing u 2. And mathematically, we like to make this look like a state and we want to write it
in a coherent way. And we sort of multiply these two things, but we must say sort of multiply,
because this strange multiplication, this, you know, we think of them as vectors or states, so
how do you multiply states? So you put something called the tensor product, a little
multiplication like this.
So you could say, don’t worry, it’s kind of like a product, and it’s the way we do it. We don’t
move things across, the first state here, the second state here, and that’s a possible state.
Now, I could have a different state. Because particle 1, in fact, could be doing something a
little different. Could be doing alpha 1 u 1 plus alpha 2 u 2, and maybe particle 2 is doing beta
1 v 1 plus beta 2 v 2. And this would be all right. I’m telling you what particle 1 is doing and I’m
telling you what particle 2 is doing and the rules of tensor multiplication or this kind of
multiplication to combine those states are just like a product, except that as I said, you never
move the states across. So you just distribute, so you have alpha 2, beta 1, the number goes
out, u 1 v 1-- that’s the first factor-- plus alpha 1 beta 2 u 1 v 2 plus alpha 2 beta 1 u 2 v 1 plus
alpha 2 beta 2 u 2 v 2. I think I got it right. Let me know. I just multiplied and got the numbers
out. The numbers can be move out across this product.
OK, so that’s a state and that’s a superposition of states, so actually, I could try to write a
different state now. You see, we’re just experimenting, but here is another state. u 1 v 1 plus u
2 v 2. Now this is a state that actually seems different. Quite different. Because I don’t seem to
be able to say that what particle 1 is doing and what particle 2 is doing separately. You see, I
can say when particle 1 is doing u 1, particle 2 is doing v 1. And if when particle 1 is doing u 2,
this is v 2. But can I write this as some state of the first particle times some state of the second
particle?
Well, let’s see. Maybe I can and can write it in this form. This is the most general state that you
can say, particle 1 is doing this, and particle 2 is doing that. So can they do that? Well, I can
compare these two terms with those and they conclude that alpha 1 beta 1 must be 1. Alpha 2
beta 2 must be also 1. But no cross products exist, so alpha 1 beta 2 must be 0 and alpha 2
beta 1 must be 0. And that’s a problem because either alpha 1 is 0, which is inconsistent, or
beta 2 is 0, which is inconsistent with that, so now this state is un-factorizable. It’s a funny state
in which you cannot say that this quantum state can be described by telling what the first
particle is doing and what the second particle is doing. What the first particle is doing depends
on the second and what the second is doing depends on the first. This is an entangled state.
And then we can build entangled states and our very strange states. So with two particles with
spins, for example, we can build an entangled states of 2 spin 1/2 particles. And this state
could look like this-- the first particle is up along z and the second particle is down along z, plus
a particle that is down along z for the first particle, but the second is up along z.
And these are 2 spin 1/2 particles and in the usual notation, these experiments in quantum
mechanics and black hole physics, people speak of Alice and Bob. Alice has one particle, Bob
has the other particle. Maybe Alice is in the moon and has her electron and Bob is on earth
and has his electron, and the two electrons, one on the earth and in the moon, are in this
state. So then we say that Alice and Bob share an entangled pair. And all kinds of strange
things happen. People can do those things in the lab-- not quite one in the earth and one in
the moon, but one photon at one place and another photon entangled with it at 100 kilometers
away, that’s pretty doable.
And they are in this funny state in which their properties are currently that in surprising ways.
So what happens here? Suppose Alice goes-- or let’s say Bob goes along and measures his
spin and he finds his spin down. So-- oh, you look here, oh, here is down for Bob. So at this
moment, the whole state collapses into this. Because up with Bob didn’t get realized. So once
Bob measures and he finds down, the whole state goes into this. So if Alice-- on the moon or
in another galaxy-- at that instant looks at her spin, she will find it’s up before light has had
time to get there. Instantaneously. It will go into this state.
People were sure somehow this violates special relativity. It doesn’t. You somehow when you
think about this carefully, you can’t quite send information, but the collapse is instantaneous in
quantum mechanics. Somehow, Bob and Alice cannot communicate information by sharing
this entangled pair, but it’s an interesting thing why it cannot happen.
Einstein again objected to this. And he said, this is a fake thing. You guys are going to share–
and now, of course, they have to share many entangled pairs to do experiments, so maybe
1,000 entangled pairs. And Einstein would say, no, that’s not what’s happening. What’s
happening is that some of your entangled pairs are this. That is, Bob is down, Alice is up,
some of them are this-- and there’s no such thing as this entanglement and indeed, if you
measure and you find down, she will find up, and if you measure and you find up, she will find
down, and there’s nothing too mysterious here.
But then came John Bell in 1964 and discovered his Bell inequalities that demonstrated that if
Alice and Bob can measure in three different directions, they will find correlations that are
impossible to explain with classical physics. It took a lot of originally for Bell to discover this,
that you have to measure in three directions, and therefore, the kind of correlations that
appear in entangled states are very subtle and pretty difficult to disentangle. So that’s why
entanglement is a very peculiar subject. People think about it a lot because it’s very
mysterious. It somehow violates classical notions, but in a very subtle way.

L2.3 Mach-Zehnder interferometers and beam splitters (15:32)

MITOCW | watch?v=0USje5vTIKs
PROFESSOR: Mach-Zehnder–
interferometers.
And we have a beam splitter.
And the beam coming in, it splits into 2. A mirror–
another mirror. The beams are recombined into another beam splitter. And then, 2 beams
come out. One to a detector d0-- and a detector d1.
We could put here any kind of devices in between. We could put a little piece of glass, which is
a phase shifter. We’ll discuss it later. But our story is a story of a photon coming in and
somehow leaving through the interferometer.
And we want to describe this photon in quantum mechanics. And we know that the way to
describe it is through a wave function. But this photon can live in either of 2 beams. If a photon
was in 1 beam, I could have a number that tells me the probability to be in that beam.
But now, it can be in either of 2 beams. Therefore, I will use two numbers. And it seems
reasonable to put them in a column vector. Two complex numbers that give me the probability
amplitudes–
for this photon to be somewhere. So you could say, oh, look here. What is the probability that
we’ll find this photon over here? Well, it may depend on the time. I mean, when the photon is
gone, it’s gone.
But when it’s crossing here, what is the probability? And I have 2 numbers. What is the
probability here, here, here, here? And in fact, you could even have 1 photon that is coming in
through 2 different channels, as well.
So I have 2 numbers. And I want, now, to do things in a normalized way. So this will be the
probability amplitude to be here. This is the probability amplitude to be down. And therefore,
the probability to be in the upper one-- you do norm squared.
The probability to be in the bottom one, you do norm squared. And you get 1. Must get 1. So if
you write 2 numbers, they better satisfy that thing. Otherwise, you are not describing
probabilities.
On the other hand, I may have a state that is like this. Alpha-- oh, I’ll mention other states.
State 1-0 is a photon in the upper beam.
No probability to be in the lower beam. And state 0-1 is a photon in the lower beam.
So these are states. And indeed–
think of superposition. And we have that the state, alpha beta-- you know how to manipulate
vectors-- can be written as alpha 1-0. Because the number goes in and becomes alpha 0. Plus
beta 0-1.
So the state, alpha beta, is a superposition with coefficient alpha of the state in the upper
beam plus the superposition with coefficient beta of the state in the lower beam. We also had
this little device, which is called the beam shifter of face delta.
If the probability amplitude completing is alpha to the left of it, it’s alpha e to the i delta to the
right of it, with delta a real number. So this is a pure phase. And notice that alpha is equal to e
to the i delta.
The norm of a complex number doesn’t change when you multiply it by a phase. The norm of
a complex number times a phase is the norm of the complex number times the norm of the
phase. And the norm of any phase is 1.
So actually, this doesn’t absorb the photon, doesn’t generate more photons. It preserves the
probability of having a photon there, but it changes the phase.
How does the beam splitter work, however? This is the first thing we have to model here. So
here is the beam splitter.
And you could have a beam coming–
A 1-0 beam hitting it. So nothing coming from below. And something coming from above. And
then, it would reflect some and transmit some. And here is a 1-- is the 1 of the 1-0. And here’s
an s and a t.
Which is to mean that this beam splitter takes the 1-0 photon and makes it into an st photon.
Because it produces a beam with s up and t down.
On the other hand, that same beam splitter-- now, we don’t know what those numbers s and t
are. That’s part of designing a beam splitter. You can ask the engineer what are s and t for the
beam splitter.
But we are going to figure out what are the constraints. Because no engineer would be able to
make a beam splitter with arbitrary s and t.
In particular, you already see that if 1-0-- if a photon comes in, probability conservation, there
must still be a photon. You need that s squared plus t squared is equal to 1 because that’s a
photon state.
Now, you may also have a photon coming from below and giving you uv. So this would be a 0-
1 photon, giving you uv.
And therefore, we would say that 0-1–
gives you uv. And you would have u plus v norm squared is equal to 1. So we need,
apparently, 4 numbers to characterize the beam splitter. And let’s see how we can do that.
Well, why do we need, really, 4 numbers? Because of linearity. So let’s explore that a little
more clearly. And suppose that I ask you, what happens to an alpha beta state–
alpha beta state if it enters a beam splitter? What comes out? Well, the alpha beta state, as
you know, is alpha 1-0 plus beta 0-1. And now, we can use our rules.
Well, this state, the beam splitter is a linear device. So it will give you alpha times what it
makes out of the 1-0. But out of the 1-0 gives you st.
And the beta times 0-1 will give you beta uv. So this is alpha s plus beta u times alpha t plus
beta v. And I can write this, actually, as alpha beta times the matrix, s u t v.
And you get a very nice thing, that the effect of the beam splitter on any photon state, alpha
beta, is to multiply it by this matrix, s u t v. So this is the beam splitter. The beam splitter acts
on any photon state. And out comes the matrix times the photon state.
This is matrix action, something that is going to be pretty important for us.
How do we get those numbers? After all, the beam splitter is now determined by these 4
numbers and we don’t have enough information. So the manufacturer can tell you that maybe
you’ve got-- you bought a balanced beam splitter.
Which means that if you have a beam, half of the intensity goes through, half of the intensity
gets reflected. That’s a balanced beam splitter. That simplifies things because the intensity
here, the probability, [INAUDIBLE] must be the same as that.
So each norm squared must be equal to 1/2, if you have a balanced–
beam splitter.
And you have s squared equal t squared equal u squared equal v squared equal 1/2. But
that’s still far from enough to determine s, t, u, and v. So rather than determining, them at this
moment, might as well do a guess.
So can it be that the beam splitter matrix-- Could it be that the beam splitter matrix is 1 over
square root of 2, 1 over square root of 2, 1 over the square root of 2, and 1 over square root
of 2. That certainly satisfies all of the properties we’ve written before.
Now, why could it be wrong? Because it could be pluses or minuses or it could be i’s or
anything there. But maybe this is right. Well, if it is right, the condition that it be right is that, if
you take a photon state, 1 photon-- after the beam splitter, you still have 1 photon.
So conservation of probability. So if you act on a normalized photon state that satisfies this
alpha squared plus beta squared equal 1, it should still give you a normalized photon state.
And it should do it for any state.
And presumably, if you get any numbers that satisfy that, some engineer will be able to build
that beam splitter for you because it doesn’t contradict any physical principle. So let’s try acting
on this with on this state-- 1 over square root of 2, 1 over square root of 2.
Let’s see. This is normalized-- 1/2 plus 1/2 is 1. So I multiply. I get 1/2 plus 1/2 is 1, and 1.
Sorry, this is not normalized. 1 squared plus 1 squared is 2, not 1. So this can’t be a beam
splitter. No way.
We try minus 1 over square root of 2. Actually, if you try this for a few examples, it will work. So
how about if we tried in general. So if I try it in general, acting on alpha beta, I would get 1 over
square root of 2 alpha plus beta and alpha minus beta.
Then, I would check the normalization. So I must do norm of this 1 squared. So it’s 1/2 alpha
plus beta squared plus 1/2 alpha minus beta norm squared.
Well, what is this? Let me go a little slow for a second. [INAUDIBLE] plus beta star.
Plus alpha minus beta. Alpha star minus beta star.
Well, the cross terms vanish. And alpha alpha star, alpha alpha star, beta beta star, beta beta
star add. So you do get alpha squared plus beta squared. And that’s 1 by assumption because
you started with a photon.
So this works. This is a good beam splitter matrix.
It does the job. So actually–
Consider this beam splitters. Actually, it’s not the unique solution by all means.
But we can have 2 beam splitter that differ a little bit. So I’ll call beam splitter 1 and beam
splitter 2. Beam splitter or 1 will have this matrix. And beam splitter 2 will have the matrix were
found here, which is a 1 1 1 minus 1.
So both of them work, actually. And both of them are good beam splitters. I call this–
beam splitter 1.
And this, beam splitter 2. And we’ll keep that. And so we’re ready, now, to think about our
experiments with the beam splitter.

L2.4 Interferometer and interference (12:26)

MITOCW | watch?v=37-GdFJGSXs
PROFESSOR: And let me I assume, for example, that I’ll put the state alpha beta in. Alpha and beta. What do
I get out? So you have this state, alpha beta. What do you get out? Well, state comes in and is
acted by beam splitter 1. So you must put the beam splitter, 1 matrix.
And then it comes the mirrors. And lets assume mirrors do nothing. In fact, mirrors-- the two
mirrors would multiply by minus 1, which will have no effect. So lets ignore mirrors.
And then you get to beam splitter 2 and you must multiply by the matrix of beam squared 2.
And that’s the output. And that output is a two-component vector. That gives you the amplitude
up and the amplitude down.
So I should put BS2 here, BS1 over here, alpha beta. The numbers move away, 1 over square
root of 2 and 1 over square root of 2. Commute in matrix multiplication. Then you multiply
these two matrices. You get 0, 2, minus 2, and 0, alpha beta. And you put the 2 in so you get
beta minus alpha.
So here is the rule. If you have alpha and beta, you get, here, beta and minus alpha here, or a
beta minus alpha photon at the end. Good.
So let’s do our first kind of experiment. Our first experiment is to have the beam splitters here.
D0-- detector D0 and detector D1 over there. And let’s send in a photon over here only-- 1 or
input 0, 1.
Well this photon, 0, 1, splits here. You act with BS1-- the matrix BS1. You get two things. You
act with the matrix BS2, and it gives you this. But we have the rule already. If you have an
alpha beta, out comes a beta minus alpha. So it should have as 1, here, and minus 0, here,
which is 0. So you get a 1, 0.
So what is really happening? What’s really happening is that your photon that came in divided
in two, recombined, and, actually, there was a very interesting interference here. From the top
beam came some amplitude and gave some reflected and some transmitted. From the bottom
beam, there was some transmission and some reflection. The transmission from the top and
reflection from the bottom interfered, to give 0. And this, too, the reflection from the top and
transmission from the bottom, were coherent and added up to 1. And every single photon
ends up in D0.
If you would put the beam-- well, Mach and Zehnder were working in the late 1800s, 1890s.
And they would shine light. They had no ability to manipulate photons. But they could put
those beam splitters and they could get this interference effect, where everything goes to D0.
So far, so good.
Now let me do a slightly different experiment. I will now put the same thing, a BS1 and a beam
going in, mirror, mirror, BS2 here. But now, I will put a block of concrete here on the way. I’ll
put it like this. So that if there is any photon that wants to come in this direction, it will be
absorbed. Photon could still go like this, but nothing would go through here. And here, of
course, there might be D0 and D1. And here are the mirrors, M and M.
Now the bottom mirror is of no use anymore because there is a big block of concrete that will
stop any photon from getting there. And we are asked, again, what happens? What do the
detectors see? And this time, we still have a 01.
Now I would be tempted to use this formula, but this formula was right under the wrong
assumption-- that there was no block here. So I cannot use that formula. And certainly, things
are going to be different.
So I have to calculate things. And we’re doing a quantum mechanical calculation. Well, up to
here, before it reaches here, I can you do my usual calculation. Certainly, we have BS1 acting
on the state, 01, and this is 1 over square root of 2, I think, minus 1, 1, 1, 1. Yup, that B is 1,
acting on 01. And that gives me 1 over square root of 2, 1 over square root of 2. So, yes, here
I have one over square root of 2 amplitude. And here I also have 1 over square root of 2
amplitude.
OK. Now that’s the end of this amplitude. It doesn’t follow. But on the other hand, in this
branch, the mirror doesn’t change the amplitude, doesn’t absorb. So you still have 1 over
square root of 2 here. And now you’re reaching BS2.
Now what is the input for BS2? The input is a one over square root 2 in the top beam, and
nothing in the lower beam because nothing is reaching BS2 from below. This is blocked. So
yes, there was some times when something reached from below, but nothing here.
So to figure out the amplitudes, here, I must do BS2 acting on 1 over the square root of 2, 0.
Because 1 over square root of 2 is coming in, but nothing is coming in from below. And,
therefore, I get 1 over the square root of 2, 1, 1, 1, minus 1, 1 over square root of 2, 0. This
time, I get 1/2 and 1/2. OK, we must trust the math. 1/2 here and 1/2 there, so 1/2 a column
vector, 1/2, 1/2.
OK, let me maybe tabulate this result, which is somewhat strange, really. So what is strange
about it is the following.
In the first case, where the interferometer was totally clear, nothing in the middle, everything
went to D2. And nothing went into D1. But now, you do something that should block some
photons. You block some photons in the lower path, and yet, now you seem to be able to get
something into D1. There is an amplitude to get into the D1. So by blocking a source, you’re
getting more somewhere. It’s somewhat counterintuitive. You will see by the end of the lecture
in 10 minutes, that it’s not just somewhat counterintuitive, it’s tremendously counterintuitive.
Let’s summarize the result here-- the outcome in the blocked lower branch case and the
probability for those events. So photon at the block-- the photon can end in three places. It can
end on the block. It can end on the D0. Or it can end on D1. So photon at the block-- well, the
amplitude to be here is one over square root of 2. The probability should be 1/2. Photon at D0,
probability amplitude, 1/2, probability, 1/4-- photon at D1, probability, 1/4.
You could put another table here-- outcome all open, probability. And in this case, there’s just
photon at D0. That’s 1. And photon at D1 was 0.

L2.5 Elitzur-Vaidman bombs (10:29)

MITOCW | watch?v=vFZeh8bMx58
PROFESSOR: So far so good. So here is the kind of very entertaining thing that happens when you try to do
some physics with this. And this was done by two physicists, Elitzur and Vaidman in Tel Aviv,
they invented or fantasized about some sort of bombs-- things that explode. So they’re called
Elitzur-Vaidman bombs.
And you could invent different things, but here is what an Elitzur-Vaidman bomb is-- some sort
of bomb, and the way it works is with a photon detector. So there’s a little tube in the bomb,
and there’s a photon detector.
And you have your bomb, and you want to detonate it-- you send the photon in, you send the
photon in through the tube. And the photon, it’s detected by the detector. And the bomb
explodes.
On the other hand, if the bomb is defective, the photon goes in, and the detector doesn’t work.
The photon goes out. Just goes through.
So that’s an Elitzur-Vaidman bomb. And here is the puzzle for you-- suppose you have those
bombs, and unfortunately, those bombs, after time, they decay. And sometimes the detectors
go wrong, and they don’t work anymore.
So you have 10 bombs, and you know, maybe five have gone wrong. And now, you have
maybe a very important mission and you need the bomb that really works. So what do you do?
Let’s assume you cannot break apart the detector-- it’s just too complicated. So you have the
bombs, and you want to test them. If you send in a photon and nothing happens, the bomb is
not working. But if you send in a photon, and the bomb is working-- explodes. So you cannot
use it anymore.
So the question that Elitzur and Vaidman pose, is there a way to certify that the bomb is
working without exploding it? Can you do that?
The answer looks absolutely impossible. And certainly, in classical physics, it’s completely
impossible. You either do the measurement to see if the detector works, and if it works, your
lab goes off. It’s totally destroyed. And if it doesn’t work, well, OK, it’s not a good bum
anyways. So there’s no way out.
But there is a way out, and it is to insert this bomb in the mass [? center ?] interferometer. So
here we go. We put the mass [? center ?] interferometer, and we insert the bomb here with
the detector along this place. D 0 and D 1 are still here.
And now, you put this, and you send in a photon. So let’s see what happens if you send in a
photon.
Suppose the bomb is defective-- bomb is defective. So what are the possible outcomes?
Outcome and probability.
Photon goes to D 0-- 0. Photon to D 1, bomb explodes. Well, we said the bomb is defective.
So if the bomb is defective, we said it’s like a detector that doesn’t work, and lets the photons
go through. So if the bomb is defective, it’s as if there no bomb here, and you have the
situation where all is open.
So there will be a probability of 1 to get the photon to D 0-- a probability of 0 to get the photon
to D 1. And the bomb, of course, doesn’t explode-- probability of 0.
On the other hand, suppose the bomb is good-- bomb is good. And then, what are the
outcomes? And what are the probabilities?
Well, you know, more or less, what’s happening already. The bomb is good means there is a
detector that never fails to detect the photon. And if a photon comes in, it will capture it-- it will
block it. The bomb will explode. So you have your mass [? center ?] interferometer, and you’ve
really done the equivalent of this-- if the bomb is really working. You’ve put a block of
concrete-- it’s going to absorb the photon.
So if the bomb is really working, the outcome are the following-- well, I’m sorry to say, your lab
will explode half of the times, because the photon on the block happens, and bomb explodes
with probability 1/2. On the other hand, in this situation, it is possible that the photon-- photon–
at D 0, and bomb doesn’t explode-- not explode. And there is a probability 1/4.
And there is a probability, 1/4, that the photon is at D 1, and the bomb does not explode. But
here is the catch now-- yes, half of the bombs exploded, we’re sorry about that. But if the
bomb doesn’t work, there is no way a photon can reach D 1, because if a bomb doesn’t work,
all photons go to D 0.
So the fact that some photons, a quarter percent of the time, 1/4-- 25% of the time-- reach D 1
implies that photon is at D 1, and bomb did not explode. But the bomb is good.
So look what has happened-- it’s really strange. The photon went-- the bomb was here, it was
ready to explode. The photons kept the bomb, and ended at D 1, and you still know that the
bomb works now-- even though the photon never went through the detector. It never touched
here, it never went inside and get detected. Somehow it went through the other way, but you
know that the bomb is working, with a quarter percent efficiency.
We will do exercises, and it’s possible to raise the efficiency to 50%. And if you put the bomb
inside a cavity, a resonant cavity with photons going around, you can reach 99% efficiency. So
the probability of blowing up MIT goes down to 1%.
[LAUGHTER]
I don’t know if we can live with that-- I don’t think so. But anyway, this is a true fact–
experiments without bombs have been done, and it shows that in quantum mechanics, you
can do very surprising measurements.

  • 8
    点赞
  • 8
    收藏
    觉得还不错? 一键收藏
  • 0
    评论
评论
添加红包

请填写红包祝福语或标题

红包个数最小为10个

红包金额最低5元

当前余额3.43前往充值 >
需支付:10.00
成就一亿技术人!
领取后你会自动成为博主和红包主的粉丝 规则
hope_wisdom
发出的红包
实付
使用余额支付
点击重新获取
扫码支付
钱包余额 0

抵扣说明:

1.余额是钱包充值的虚拟货币,按照1:1的比例进行支付金额的抵扣。
2.余额无法直接购买下载,可以购买VIP、付费专栏及课程。

余额充值