To install click the Add extension button. That's it.

The source code for the WIKI 2 extension is being checked by specialists of the Mozilla Foundation, Google, and Apple. You could also do it yourself at any point in time.

4,5
Kelly Slayton
Congratulations on this excellent venture… what a great idea!
Alexander Grigorievskiy
I use WIKI 2 every day and almost forgot how the original Wikipedia looks like.
Live Statistics
English Articles
Improved in 24 Hours
Added in 24 Hours
Languages
Recent
Show all languages
What we do. Every page goes through several hundred of perfecting techniques; in live mode. Quite the same Wikipedia. Just better.
.
Leo
Newton
Brights
Milds

Squeezed states of light

From Wikipedia, the free encyclopedia

In quantum physics, light is in a squeezed state[1] if its electric field strength Ԑ for some phases has a quantum uncertainty smaller than that of a coherent state. The term squeezing thus refers to a reduced quantum uncertainty. To obey Heisenberg's uncertainty relation, a squeezed state must also have phases at which the electric field uncertainty is anti-squeezed, i.e. larger than that of a coherent state. Since 2019, the gravitational-wave observatories LIGO and Virgo employ squeezed laser light, which has significantly increased the rate of observed gravitational-wave events.[2][3][4]

YouTube Encyclopedic

  • 1/5
    Views:
    22 142
    8 902
    785
    1 273
    6 428
  • 4. Non-classical light, squeezing, Part 1
  • 4. Non-classical light, squeezing, Part 2
  • 11.2 Application of squeezed states
  • Week11b: Quantum Optics
  • The Wave Hunters - 04: Squeezed light

Transcription

The following content is provided under a Creative Commons license. Your support will help MIT OpenCourseWare continue to offer high-quality educational resources for free. To make a donation, or view additional materials from hundreds of MIT courses, visit MIT OpenCourseWare at ocw.mit.edu. PROFESSOR: Good afternoon. We continue our discussion of quantum states of light. We talked at length about coherent state, and when you talk about quantum states of light, each mode of the electromagnetic field is an harmonic oscillator. We also encountered, naturally, the number states. And we realized-- yesterday, actually, in the last class-- that those number states have non-classical properties. For instance, they have a g2 function, the second order correlation function, which is smaller than 1, which is impossible for classic light, as you're proving in one of your homework assignment. So at that point, we have encountered coherent states, which are as close as possible to classical states. And we have found the number states as non-classical states. Well, are there other interesting states? I wouldn't ask you this question if the answer would not be yes, and this is what we want to discuss today. We want to talk about non-classical states of light, which we can engineer, actually, in the laboratory, by sending laser light through nonlinear crystals. Those go by the name, squeezed states. Just to give you the cartoon picture, in our two-dimensional diagram, with the quasi-probabilities, we have coherent states, where the area of this disk, delta x delta p, is h-bar over two. It's uncertainty limited. What we can do we now, is-- we cannot go beyond this. This is the fundamental limit of quantum physics. However, we can take this circle and we can squeeze it. We can squeeze it horizontally, we can squeeze it into an elongated vertical shape, or we can squeeze it at any angle. That's what we call, squeezed states. And those states have non-classical properties. They are important for metrology they are important for teleportation. There are lots and lots of reasons why you want to know about them. But again, as so often, I feel I cannot convey to you the excitement of doing squeezing in the quantum domain. And many, many physicists now, they hear about squeezing just in the quantum domain. But I want to start with classical squeezing. I will actually show you video of an experiment on classical squeezing. You can see squeezing with your own eyes. But this is just sort of to set the stage, to also get a feel of what squeezing is. And then we'll do quantum mechanical squeezing. But maybe-- tongue in cheek-- I would say, since classical harmonic oscillators and quantum harmonic oscillators have a lot in common, the step from classical squeezing to quantum of mechanical squeezing is actually rather small. It's nice to squeeze light. It's nice to have those non-classical states. But the question is, how can you detect it? If you can't detect it, you can't take advantage of it. And the detection has to be face-coherent. I will tell you what that is. And it goes by the name, homodyne detection. And finally, we can take everything we have learned together, and discuss how, in the laboratory, teleportation of a quantum state is done. There is a nice teleportation scheme, and I want to use that as an example that the language and the concepts I've introduced are useful. Concepts like, squeezing operator, displacement operator-- those methods allowing us to, in a very clear way, discuss schemes which lead to teleportation. That's the menu for today. Let's start with classical squeezing. For squeezing, we need an harmonic oscillator, means for parabolic potential, we have potential v of x. And then we study the motion of-- that should be x squared-- the motion of a particle in there. Before I even get started any equation, let me explain what the effect of squeezing will be about. If you have an harmonic oscillator, you have, actually, the motion of a pendulum. It has two quadrature components, the cosine motion and the sine motion. And they are 90 degrees out of phase. What happens now is, if you parametrically drive the harmonic oscillator-- you modulate the harmonic oscillator potential-- it's to omega. I will show you mathematically, it's very, very easy to show, that depending on the phase of the drive, you will actually exponentially amplify the sine motion, and exponentially damp the cosine motion. Or if you change, vice versa. So by driving the system, you can amplify one quadrature component, and exponentially die out the other quadrature component. And that is called, classical squeezing. Let's do the math. It's very simple. Our equation of motion has the two solutions I've just mentioned. It has a solution with cosine omega 0, and one with sine omega 0 t. And we have two coefficients. The cosine is called, c. The sine coefficient is called, s. I have to call it c 0, because I want to call that c and s. So what we have here is, we have the two quadrature components of the motion in an harmonic oscillator. And graphically, we need that for the electromagnetic field, as well. When we have our two axes, like, you know, the complex plane for the cosine of probabilities, I call one the s-axis. One is the c-axis. That's just something which confuses me. If you have only one-- just give me one second. Cosine-- Yeah. If you have only cosine motion, the s component is 0, and the harmonic oscillator would just oscillate here. If you have only a sine component, you stay on the x-axis. And now, if you have an equal amount of cosine and sine, then you can describe the trajectory to go in a circle. OK. This is just the undriven harmonic oscillator. I don't want to dwell on it any longer. But what we are doing now is, we are adding a small parametric drive. Parametric drive means we modulate the spring constant, or we replace the original harmonic potential, which was this, by an extra modulation term. So we have a small parameter, epsilon. And as I pointed out, the modulation is at twice the resonance frequency. Now we want to solve the equation of motion for the harmonic oscillator, using this added potential. The way how we want to solve it is, we assume epsilon is very small. So if the pendulum is swinging with cosine omega t, it will take a while for the epsilon term-- for the small term-- to change the motion. So therefore, we assume that we can actually go back and use our original solution. And assume that over a short term, the epsilon term is not doing anything. So for a short time, it looks like an harmonic oscillator with a sine omega 0, and cosine omega 0 t oscillation. But over any longer period of time, the small term will have an effect. And therefore, the coefficients c of t, c, and s are no longer constant, but change as a function of time. We want to solve, now, the equation of motion. That means we use this, here, as our ansatz. And we calculate the second derivative. We assume that the coefficients c and s are changing slowly. Therefore, the second derivative of c and s can be neglected. By taking the derivative of the second derivative of the cosine term and the sine term, of course we simply get, minus omega 0 squared, x of t. And now we have the second-order derivatives. Since we neglect the second-order derivative of c and s, the other terms we get when we take the second derivative is, first derivative of c times first derivative of cosine. First derivative of s times first derivative of sine. So we get two more terms, which are, minus omega 0 c dot, times sine omega 0 t. Plus omega 0 s dot, times cosine omega 0 t. This is the second derivative of our ansatz for x. This has to be equal to the force provided by the potential. So taking the potential-- We need, now, the derivative of the potential, for the potential of use across this line. The first part is the unperturbed harmonic oscillator, which gives us simply, omega 0 squared times x. And the second term, due to the parametric drive, is 2 sine omega 0 t. And now, for x, we use our ansatz for x, which is the slowly-changing amplitude c times cosine omega 0 t, plus s times sine omega 0 t. Those two terms cancel out. So now we have products of trig function. Sine 2 omega times cosine omega. Well, you know if you take the product of two trig functions, it becomes a trig function of the sum or the difference of the argument. So if you take sine 2 omega 0 times cosine omega 0, and we use trigonometric identities, we get an oscillation at 3 omega 0, which is 2 plus 1. And one at the difference, which is omega 0. Let me write down the terms which are of interest to us. Namely, the ones at omega 0. So let me factor out epsilon omega 0 squared over 2. Then we have the term c times sine omega 0 t, plus s times cosine. And then we have terms at 3 omega 0, which we are going to neglect. Now we compare the two sides of the equations. We have sine omega 0 term. We have cosine omega 0 term. And the two sides of the equations are only consistent if the two coefficients of the sine term, and the sine term, are the same. So therefore, we obtain two equations. One for c dot, one for s dot. And these are first-order differential equations. The solution is clearly an exponential. But one has a plus sign, one has a minus sign. So the c component, the c quadrature component, is exponentially amplified with this time constant. Whereas the sine component is exponentially de-amplified. This finishes the mathematical discussion of classical squeezing. We find that s of t, and c of t, are exponential functions. In one case, it's exponentially increasing. In the other case, it is exponentially decreasing. And that means that, well, if we go to our diagram, here-- and let's assume we had an arbitrary superposition of cosine and sine amplitude. This is cosine. This is sine. We had sort of a cosine oscillation, and a sine oscillation. Which means that, as a phasor, the system was moving on an ellipse. If the sine component is exponentially de-amplified, and the cosine component is exponentially amplified, that means whatever we start with is squashed horizontally, is squashed vertically. And is amplified horizontally. In the end, it will become a narrow strip. So this is classical squeezing. You may want to ask, why did I neglect the 3 omega 0 term. Well, I have to, otherwise I don't have a solution. Because I have to be consistent with my approximations. So what I did here is, I had an equation where I have the clear vision that the solution has a slowly varying c and s coefficient. And then I simply use that. I take the second-order derivative, and I have only Fourier components with omega 0, the sine, and cosine. Now I've made an approximation, here. For the derivative of the potential, the first line is exact. But in order to match the approximation I've done on the other side, I can only focus on two Fourier components resonant with omega 0, which I have here. So in other words, the 3 omega 0 term would lead to additional accelerations. Which I have not included in the treatment. So it's consistent with the ansatz. It's consistent with the assumption that we have resonant oscillations with a slowly changing amplitude. There will be a small [INAUDIBLE] for your omega 0, but it will be small. Any questions about that? Let me then show you an animation of that. Classroom files. [VIDEO PLAYBACK] -We have Dave Pritchard, professor of physics at MIT, demonstrating what squeezing is. Right now, we see a wave that's going around in a circle. What's next? What's going to happen now, Professor Pritchard? -Well, if we drive it in twice the basic period, then we will parametrically amplify one quadrature component, and we will un-amplify the other one. So now I'm going to start doing that. And then you notice that its motion turns into an ellipse. We've amplified this quadrature component, but we've un-amplifed that one. And that's squeezing. [END VIDEO PLAYBACK] PROFESSOR: Feel free to try it at home. [LAUGHTER] PROFESSOR: Actually, you may start to think about this demonstration. What he has shown was, when you have a circular pendulum which goes in a circle or an ellipse, and you start pulling on the rope with a certain phase, that one quadrature component will be de-amplified. The other one will be amplified. And as a result, no matter what the circular or the elliptical motion was, after driving it for a while, it will only swing in one direction. And this is the collection you have amplified. There is one thing which should give you pause. I have discussed, actually, a single harmonic oscillator. What Dave Pritchard demonstrated was actually two harmonic oscillators. The harmonic oscillator has an x motion and a y motion. However, you can say, this was just sort of a trick for the demonstration, because when you have a circular motion, initially, you have the sine omega and the cosine omega 0 component present simultaneously. And you can see what happens to the sine and the cosine component in one experiment. So in that sense, he did two experiments at once. He showed what happens when you have, initially, a sine component, and what happens when you initially have a cosine component. OK. So we know what classical squeezing is. And what we have learned, also-- and this helps me now a lot to motivate how we squeeze in quantum mechanics-- you have realized that what is really essential here is, to drive it to omega 0. What we need now to do squeezing in the quantum domain, if we want to squeeze light, we need something at 2 omega 0. So let's now squeeze quantum mechanically. Go back here. The second sub-section is now, squeezed quantum states. What we want to discuss is, we want to discuss a quantum harmonic oscillator. We want to have some form of parametric drive at 2 omega 0. And this will result in squeezed states. Now, what does it require, if you want to bring in 2 omega 0? Well, let's not forget our harmonic oscillators are modes of the electromagnetic field. If you now want to couple a mode of the electromagnetic field, at 2 omega 0, with our harmonic oscillator at omega 0, we need a coupling between two electromagnetic fields. So therefore, we need nonlinear interactions between photons. So this was a tautology. We need nonlinear physics, which leads to interactions between photons. Linear physics means, each harmonic oscillator is independent. So we need some nonlinear process six which will be equivalent to have interactions between photons. The device which we will provide that is an optical parametric oscillator. I could spend a long time explaining to you how those nonlinear crystals work. What is the polarization, what is the polarizability, how do you drive it, what is the nonlinearity. But for the discussion in this class, which focuses on fundamental concepts, I can actually bypass it by just saying, assume you have a system-- and this is actually what the optical parametric oscillator does, is you pump it with photons at 2 omega 0. And then the crystal generates two photons at omega 0. Which of course, is consistent with energy conservation. And if you fulfill some phase-matching condition, it's also consistent with momentum conservation. But I don't want to go into phase-matching at this point. Technically, this is done as simple as that. You have to pick the right crystal. Actually, a crystal which does mixing between three photon fields cannot have inversion symmetry, otherwise this nonlinear term is 0. What you need is a special crystal. KDP is a common choice. And this crystal will now do for us the following. You shine in laser light. Let's say, at 532 nanometer, a green light. And then this photon breaks up into two photons of omega 0. This is how it's done in the laboratory. The piece of art is, you have to pick the right crystal. It has to be cut at the right angle. You may have to heat it, and make sure that you select the temperature for which some form of resonant condition is fulfilled, to do that. But in essence, that's what you do. One laser beam, put in a crystal, and then the photon is broken into two equal parts. And these are our two photons at omega 0. OK. I hope you enjoy the elegance-- we can completely bypass all the material physics by putting operators on it. We call this mode, b. And we call this mode, a. So the whole parametric process, the down conversion process of one photon into two, is now described by the following Hamiltonian. We destroy a photon in mode b, a 2 omega 0. And now we create two photons at omega. We destroy a photon at 2 omega 0, create two photons at omega 0. And since the Hamiltonian has to be Hamiltonian, the opposite, the time-reverse process, has to be possible, too. And that means we destroy two photons at omega 0, and create one photon at 2 omega 0. So now we forget about nonlinear crystals, about non-inversion symmetry in materials. We just take this Hamiltonian and play with it. By simply looking at the Hamiltonian, what is the time evolution of a photon field under this Hamiltonian. We figure out what happens when you send light through a crystal, and what is the output. And I want to show you now that the output of that is squeezed light, which is exactly what I promised you with these quasi-probabilities. We have a coherent state, which is a nice circle. We time-evolve the coherent state, our nice round circle, with this Hamiltonian. And what we get is an ellipse. And if you want intuition, look at the classical example we did before, which really tells you in a more intuitive way what is happening. OK. We want to make one simplifying assumption, here. And this is that we pump the crystal at 2 omega 0 with a strong laser beam. So we assume that the mode, b, is a powerful laser beam. Or in other words, a strong coherent state. We assume that the mode, b, is in a coherent state. Coherent states are always labeled with a complex parameter, which I call beta, now. Well, it's mode b, therefore I call it, beta. For mode a, I've called it, alpha. The coherent state has an amplitude, which I call, r over 2. And it has a phase. We know, of course, that the operator, b, acting on beta, gives us beta times beta, because a coherent state is an eigenstate of the annihilation operator. But when we look at the action of the operator b plus, the photon creation operator, the coherent state is not an eigenstate of the creation operator. It's only an eigenstate of the annihilation operator. But what sort of happens is, the coherent state is the sum over many, many number states with n. And the creation operator goes from n, to n plus 1, and has matrix elements which are square root n plus 1. So in other words, if n is large, and if we don't care about the subtle difference between n, and n plus 1, in this limit the coherent state is also an eigenstate of the creation operator, with an eigenvalue, which is beta star. This means that we have a coherent state which is strong. Strong means, it has a large amplitude of the electric field. The photon states which are involved, n, are large. And we don't have whether it be n, or n plus 1. This is actually, also, I should mention it here, explicitly-- this is sort of the step when we have a quantum description of light. And we replace the operators, p and p dega, by a c number, then we really go back to classical physics. Then we pretend that we have a classical electric field, which is described by the imaginary part of beta. So when you have an Hamiltonian, where you write down an electric field, and the electric field is not changing-- you have an external electric field. This is really the limit of a quantum field, where you've eliminated the operator by a c number. This is essentially your electric field. And we do this approximation, here. Because we are interested in the quantum features of mode a-- a is our quantum mode, with single photons, or with a vacuum state, and we want to squeeze it. b is just, they have parametric drive. With this approximation, we have only the a operators. This is our operator. Any question? AUDIENCE: [INAUDIBLE] would give us a [INAUDIBLE], right? PROFESSOR: Yes, thank you. That means, here should be a minus sign, yes. OK. I've motivated our discussion with this nonlinear crystal, which generates pair of photons. This is the Hamiltonian which describes it. And if you want to have a time evolution by this Hamiltonian, you put this Hamiltonian into a time evolution operator. In other words, you-- e to the minus iHt is the time evolution. If you now evolve a quantum state of light for a fixed time, t, we apply the operator, e to the minus iHt, to the quantum state of light. What I've just said is now the motivation for the definition of the squeezing operator. The squeezing operator, S of r, is defined to be the exponent of minus r over 2, a squared minus a dega squared. This is related to the discussion above. You would say, hey, you want to do that time evolution, where is the i? Well, I've just made a choice of phi. If phi is chosen to be pi over 2, then the time evolution with the Hamiltonian, above, gives me the squeezing operator, below. So with that motivation we are now studying, what is the squeezing operator doing to quantum states of light? Any questions about that? I know I spent a lot of time on it. I could have taught this class by just saying, here is an operator, the squeezing operator. Trust me, it does wonderful things. And then we can work out everything. But I find his unsatisfying, so I wanted to show you what is really behind this operator. And I want you to have a feeling, where does this operator come from, and what is it doing? In essence, what I've introduced into our description is now an operator, which is creating and destroying pairs of photons. And this will actually do wonderful things to our quantum states. What are the properties of the squeezing operator? What is important is, it is unitary. It does a unitary time evolution. You may not see that immediately, so let me explain that. You know from your basic quantum mechanics course, that e to the i operator A is unitary, when A is Hermitian. So the squeezing operator-- with the definition above-- can be written as, I factor out 2 i's over 2 a squared minus a dega squared. And you can immediately verify that this part, here, is Hermitian. If you do the Hermitian conjugate, a squared turns into a dega squared. a dega squared turns into a squared. So we have a problem with a minus sign. But if you do the complex conjugate of i, this takes care of the minus sign. So this part is Hermitian. We multiply it with i, therefore this whole operator. Thus a unitary transformation in [INAUDIBLE]. Any questions? OK. So after being familiar with this operator, we want to know, what is this operator doing? I can describe, now, what this operator does, in a Schrodinger picture, or in a Heisenberg picture. I pick whatever is more convenient. And for now, this is the Heisenberg picture. In the Heisenberg picture, what is changing are the operators. Therefore, in the Heisenberg picture, this unitary transformation transforms the operators. And we can study what happens when we transform the operator, x. The unitary transformation is done by-- the operator, x, is transformed by multiplying from the left side with S, from the righthand side with S dega. You are familiar with expressions like, this, and how to disentangle them. If you have an e to the i alpha, e to the minus alpha, if you could move the alpha past x. So if A and x commute, i A, minus i A would just give unity. So therefore, this expression is just x, unless you have non-Hermitian commutators between A and x. I think you have solved, in your basic mechanics course, many such problems which involve identities of that form. Then there are higher order commutator, the commutator of A with the commutator of a x. Unless one of those commutator vanishes, you can get an infinite series. Our operator, A, is nothing else than the annihilation operator, a squared minus the creation operator, a dega squared. So we can express everything in terms of a, and a dega. The position operator in our harmonic oscillator can also be expressed by a, and a dega. By doing elementary manipulations on the righthand side, and recouping terms, you find immediately that the unitary transformation of the Heisenberg operator, x, gives you an x operator back. But multiplied with an exponential, e to the r. And if we would do the same to the momentum operator, which is a minus a dega over square root 2, we will find that the unitary transformation of the momentum operator is de-amplifying the momentum operator by an exponential factor. If we would assume that we have a vacuum state in the harmonic oscillator, and while classically, it would be at x equals 0, p equals 0, quantum mechanically, we have single-point noise in x, and single-point noise in p. Then you would find that the squeezing operator is amplifying the quantum noise in x. But it squeezes, or reduces, the noise in p. If we apply this squeezing operator to the vacuum state, we obtain what is usually called, squeezed vacuum. And it means that, in this quasi-probability diagram, the action of the squeezing operator is turning the vacuum state into an ellipse. What happens to energy, here? The vacuum state is the lowest-energy state. If you now act with a squeezing operator to it, we obtain a state which has-- the same energy? Is it energy-conserving, or very high energy? AUDIENCE: Higher [INAUDIBLE]. PROFESSOR: Yes. Why? AUDIENCE: It's no longer the [INAUDIBLE]. PROFESSOR: Sure, yeah. It's a vacuum state. We act on the vacuum state, but we get a state which is no longer the vacuum state. The reason why we have extra energy-- the squeezed vacuum is very, very energetic. Because the squeezing operator had a dega squared, a squared. Well a squared, the annihilation operator acting on the vacuum, gives 0. But what we are creating now, we are acting on the vacuum, and we are creating pairs of photons. So we are adding, literally, energy to the system. And the energy, of course, comes from the drive laser, from the laser 2 omega 0, which delivers the energy in forms of photons which are split into half, and they go into our quantum field. In the limit of infinite squeezing-- I will show it to you, mathematically, but it's nice to discuss it already here. In the limit of infinite squeezing, what is the state we are getting? AUDIENCE: Eigenstate of momentum. PROFESSOR: Eigenstate momentum. We get the p equals 0 eigenstate. What is the energy of the p equals 0 eigenstate? AUDIENCE: Infinite. It has to contain all number states. PROFESSOR: It contains all number states? OK, you think immediately into number states, which is great. But in a more pedestrian way, the p equals 0 state has no kinetic energy. But if a state is localized in momentum, p equals 0, it has to be infinitely smeared out on the x-axis. And don't forget, we have an harmonic oscillator potential. If you have a particle which is completely delocalized in x, it has infinite potential energy at the wings. So therefore in the limit of extreme squeezing, we involve an extreme number of number states. Actually, I want to be more specific-- of photon pairs. We have states with 2n, and n can be infinitely large. But we'll see in the classical picture, what we get here when we squeeze it is, we get the p equals 0 eigenstate, which has infinite energy, due to the harmonic oscillator potential. If we would allow with the system now, after we have squeezed it, to evolve for a quarter period in the harmonic oscillator, then the ellipse would turn into an vertical ellipse. So this is now an eigenstate of x. It's the x equals 0 eigenstate. But the x equals 0 eigenstate has also infinite energy, because due to Heisenberg's uncertainty relation, it involves momentum states of infinite momentum. Questions? AUDIENCE: [INAUDIBLE] a is the photon field, right? So p is roughly the electrical field, right? PROFESSOR: Yes. AUDIENCE: So it's kind of that the electric field counts 0, and x is kind of the a, the-- and it-- because of [INAUDIBLE]. The electrical field is squeezed? PROFESSOR: Yes. AUDIENCE: It means we have no electrical field? PROFESSOR: We'll come to that in a moment. I want to do a little bit more math, to show you. I wanted to derive for you an expression of the squeeze state, in number basis, and such. Your question mentioned something which is absolutely correct. By squeezing that, we have now the p-axis is the electric field axis. So now we have, actually, in the limit of infinite squeezing, we have an electric field which has no uncertainty anymore. By squeezing the coherent state into a momentum eigenstate, we have created a sharp value for the electric field. We have created an electric field eigenstate. Well you would say, it's pretty boring, because the only electric field state we have created is electric field e equals 0. But in the next half-hour, we want to discuss the displacement operator, and I will tell you what it is. That we can now move the ellipses, and move the circles, anywhere where we want. So once we have an electric field state which has a sharp value of the electric field at e equals 0, we can just translate it. But before you get too excited about having an eigenstate of the electric field, I want you to think about what happened after one quarter-period it of the harmonic oscillator frequency. It turns upside down, and your electric field has an infinite variance. That's what quantum mechanics tells us. We can create electric fields which are very precise, but only for a short moment. So in other words, this electric field state which we have created would have a sharp value. A moment later, it would be very smeared out, then it has a sharp value again, and then it's smeared out again. I mean, that's what squeezed states are. Other questions? AUDIENCE: That's why [INAUDIBLE]. PROFESSOR: That's why we need homodyne detection. Yes, exactly. If we have squeezed something, which is sort of narrow, that's great for measurement. Now we can do a measurement of, maybe, a LIGO measurement for gravitational waves with higher precision, because we have a more precise value in our quantum state. But we have to look at it at the right time. We have to look at it synchronized with the harmonic motion. Homodyne detection means we look only at the sine component, or at the cosine component. Or if I want to simplify it, what you want to do is, if you have a state like this, you want to measure the electric field, so to speak, stroboscopically. You want to look at your system always when the ellipse is like this. The stroboscopic measurement is, as I will show you, in essence, a lock-in measurement, which is phase-sensitive. And this will be homodyne detection. So we can only take advantage of the squeezing, of having less uncertainty in one quadrature component, if you do phase-sensitive detection, which is homodyne detection. But now I'm already an hour ahead of the course. OK. Back to basics. We want to explicitly calculate, now, how does a squeezed vacuum look like. We actually want to do it twice, because it's useful. We have to see it in two different basis. One is, I want to write down the squeezed vacuum for you in a number representation. And then in a coherent state representation. The squeezing operator is an exponential function involving a squared, and a dega squared. And of course, we're now using the Taylor expansion of that. We are acting on the vacuum state. I will not do the calculation. It's again, elementary. You have n factorial, you have terms with a dega acts on c, well, you pay 2 photons. If it acts again, it adds 2 more photons, and the matrix element of a dega acting on n is square root n plus 1. You just sort of rearrange the terms. And what you find is, what I will write you down in the next line. The important thing you should immediately realize is, the squeeze state is something very special. It is the superposition of number states, but all number states are even because our squeezing operator creates pairs of photons. This is what the parametric down-conversion does. We inject photons into the vacuum, but always exactly in pairs. And therefore, it's not a random state. It's a highly correlated state with very special properties. OK. If you do the calculation and recoup the terms, you get factorials, you get 2 to the n, you get another factorial. You get hyperbolic tangent-- sorry, to the power n. And the normalization is done by the square root of the cos function. And the more we squeeze, the larger are the amplitudes at higher and higher n. But this is also obvious from the graphic representation I've shown you. Let me add the coherent state representation. The coherence states are related to the number states in that way. If we transform now from number states to coherent states, the straightforward calculation gives, now, superposition over coherent states. Coherent states require an integral. e to the minus e to the r over 2, divided by-- Anyway, all this expressions, they may not be in its general form, too illuminating. But those things can be done analytically. I just want to mention the interesting limiting case of infinite squeezing. AUDIENCE: When you do the integral over alpha, is this over like, a magnitude of alpha, or a real part, or [INAUDIBLE]? PROFESSOR: I remember, but I'm not 100% sure that alpha is real, here. I mean, it sort of makes sense, because we start with the vacuum state. And if we squeeze it, we are not really going into the imaginary direction. So I think what is involved here are only real alpha. AUDIENCE: For negative r, we should get [INAUDIBLE]. PROFESSOR: For negative r, we need imaginary state. AUDIENCE: So we should [INAUDIBLE]. PROFESSOR: Let me double-check. I don't remember that. You know that, sometimes, I admit it, the issue-- if you research material, prepare a course some years ago, you forget certain things. If I prepared the lecture, and everything worked out yesterday, I would know that. But certain things you don't remember. As far as I know, it's the real axis. But I have to double-check. The limiting case is interesting. If r goes to infinity, you can show that this is simply the integral, d alpha over coherent states. We have discussed, graphically, the situation where we had-- so these are quasi-probabilities. In that case of infinite squeezing, we have the momentum eigenstate, p equals 0. This is the limit of the infinitely squeezed vacuum, and in a coherent state representation, it is the integral over coherent state alpha. I'm pretty sure alpha is real here, seeing that now. There is a second limit, which happens simply-- you can say, by rotation, or by time evolution-- which is the x equals 0 eigenstate. And this is proportional to the integral over alpha when we take the coherent state i alpha, and we integrate from minus to plus infinity. OK. So we have connected our squeezed states, the squeezed vacuum, with number states, with coherent states. Now we need one more thing. So far we've only squeezed the vacuum, and we have defined the squeezing operator that it takes a vacuum state and elongates it. In order to generate more general states, we want to get away from the origin. And this is done by the displacement operator. The definition of the displacement operator is given here. The displacement by a complex number, alpha, is done by putting alpha, and alpha star, into an exponential function. In many quantum mechanic courses, you show very easily the elementary properties. If the displacement operator is used to transform the annihilation operator, it just does that. If you take the complex conjugate of it-- so in other words, what that means is, it's called the displacement operator, I just take that as the definition. But you immediately see why it's called the displacement operator when we do the unitary transformation of the annihilation operator, we get the annihilation operator displaced by a complex number. So the action, the transformation of the annihilation operator is the annihilator operator itself, minus a c number. So therefore, we say, the annihilation operator data has been displaced. So this is the action of the displacement operator on an operator-- on the annihilation operator. The question is now, how does the displacement operator act on quantum states? And the simplest quantum state we want to test out is the vacuum state. And well, not surprisingly, the displacement operator, displacing the vector state by alpha, is creating the coherent state, alpha. This can be proven in one line. We take our displaced vacuum, and we act on it with the annihilation operator. If we act with the annihilation operator on something, and we get the same thing back times an eigenvalue, we know it's a coherent state. Because this was the definition of coherent states. So therefore, in order to show that this is a coherent state, we want to show that it's an eigenstate of the annihilation operator. So this is what we want to do. The proof is very easy. By multiplying this expression with unity, which is DD dega, we have this. And now we can use the result for the transformation of operators. Namely, that this is simply the annihilation operator, plus alpha. If the annihilation operator acts on the vacuum state, we get 0. If alpha acts on the vacuum state, we get alpha times 0. So therefore, what we obtain is that. When the annihilation operator acts on this state, we get alpha times the state, and therefore the state is a coherent state with eigenvalue alpha. In a graphical way, if you have a vacuum state the displacement operator, D alpha, takes a vacuum state and creates a coherent state alpha. If you want to have squeezed states with a finite value, well, we just discussed the electric field. Related to the harmonic oscillator, we want squeeze states, which are not centered at the origin, which have a finite value of x or p. We can now create them by first squeezing the vacuum, and then displacing the state. AUDIENCE: What's the physical realization of the displacement operator? PROFESSOR: What is the physical realization of the displacement operator? Just one second. The physical representation of the displacement operator-- we'll do that on Monday-- is the following. If you pass an arbitrary state through a beam splitter-- but it's a beam splitter which has very, very high transmission-- and then, from the-- I'll just show that. If you have a state-- this is a beam splitter-- which has a very high transmission, T is approximately 1, then the state passes through. But then from the other side of the beam splitter, you come with a very strong coherent state. You have a coherent state which is characterized by a large complex number, beta. And then there is a reflection coefficient, r, which is very small. It sort of reflects the coherent state with an amplitude r beta. If you mix together the transmitted state and r beta-- I will show that to you explicitly, by doing a quantum treatment of the beam splitter-- what you get is, the initial state is pretty much transmitted without attenuation. But the reflected part of the strong coherent state-- you compensate for the small r by a large beta-- does actually an exact displacement of r beta. It's actually great. The beam splitter is a wonderful device. You think you have a displacement operator formulated with a's and a dega's, it looks like something extract. But you can go to the lab, simply get one beam splitter, take a strong laser beam, and whatever you send through the beam splitter gets displaced, gets acted upon by the beam splitter. Yes. AUDIENCE: You showed the displacement operator, when you acted on the vacuum state, will displace the vacuum state to a state alpha. Does it still hold if you acted on, like another coherent state. Or in this case, a squeeze state like that? PROFESSOR: Yes. I haven't shown it, but it's really-- it displaces-- When we use this representation with quasi-probabilities, it simply does a displacement in the plane. But no. To be honest, when I say it does a displacement on the plane, it reminds me that we have three different ways of defining quasi-probabilities. The w, the p, and the q representation. I know we use it all the time, that we displace things in the plane. But I'm wondering if the displacement operator does an exact displacement of all representations, or only of the q representation. That's something I don't know for sure. AUDIENCE: I was thinking it could also, like-- I mean, are you going to be able to displace all types of light, like thermal light, or any representation of light that you could put in, is the same displacement operator going to work? Or is its domain just the vacuum and coherent states? PROFESSOR: The fact is, the coherent states-- I've shown you that it's a vacuum state. I know that's the next thing to show, the displacement operator if you have a displacement by alpha followed by displacement by beta, it is equal to displacement by alpha plus beta. So displacement operator forms a group, and if you do two displacements, they equal into one area of displacement, which is the sum of two complex numbers. What I'm just saying, if you do the first displacement you can get an arbitrarily coherent state. So therefore, the displacement operator is exactly displacing a coherent state by the argument of the displacement operator. And now if you take an arbitrary quantum state and expand it into coherent states-- coherent states are not only complete, they are over-complete. All you have done is, you've done a displacement. Now the over-completeness, of course, means you have to think about it, because you can represent states in more than one way by coherent states. But if you have your representation, you just displace it, and this is the result of the displacement operator. So since the q representation is simply, you take the statistical operator and look for the elements in alpha, and if you displace alpha, the q representation has been moved around. So I'm sure that for the q representation, for the q quasi-probabilities, the displacement operator shifted around in this place. For the w and p representation, I'm not sure. Maybe there's an expert in the audience who knows more about it than I do. OK. We have just five minutes left. I want to discuss now the electric field of squeezed states. And for that, let me load a picture. Insert picture. Classroom files. Let us discuss, now, the electric field of squeezed states. Just as a reminder, we can discuss the electric field by using the quasi-probability representation. And the electric field is the projection of the quasi-probabilities on the vertical axis. And then the time evolution is, that everything rotates with omega in this complex plane. We discussed it already. For coherent state, we have a circle which rotates. Therefore, the projected fuzziness of the electric field is always the same. And as time goes by, we have a sinusoidal-bearing electric field. Let me just make one comment. If you look into the literature, some people actually say, the electric field is the projection on the horizontal axis. So there are people who say, the electric field is given by the x-coordinate of the harmonic oscillator, whereas I'm telling you, it's the p-coordinate. Well, if you think one person is wrong, I would suggest you just wait a quarter-period of the harmonic oscillator, and then the other person is right. It's really just a phase convention. What do you assume to be t equals 0-- it's really arbitrary. But here in this course, I will use the projections on the vertical axis. OK. If you project the number state, we get always, 0 electric field, with a large uncertainty. So that's just a reminder. But now we have a squeezed state. It's a displaced squeezed state. If you project it onto the y-axis, we have first some large uncertainty. I think this plot assumes that we rotate with negative time, so I apologize for that. You can just invert time, if you want. So after a quarter-period, the ellipse is now horizontal, and that means the electric field is very sharp. As time goes by, you see that the uncertainty of the electric field is large, small, large, small-- it modulates. It can become very extreme, when you do extreme squeezing, so you have an extremely precise value of the electric field, here, but you've a large uncertainty, there. Sometimes you want to accurately measure the 0 crossing of the electric field. This may be something which interests you, for an experiment. In that case, you actually want to have an ellipse which is horizontally squeezed. Now, whenever the electric field is 0, there is very little noise. But after a quarter-period, when the electric field reaches its maximum, you have a lot of noise. So it's sort of your choice which way you squeeze. Whether you want the electric field to be precise, have little fluctuations when it goes through 0, or when it goes through the maximum. So what we have done here is, we have first created the squeezed vacuum, and then we have acted on it with a displacement operator. OK. I think that's a good moment to stop. Let me just say what I wanted to take from this picture. The fact that the electric field is precise only at certain moments means that we can only take advantage of it when we do a phase-sensitive detection. We only want to sort of, measure, the electric field when it's sharp. Or-- this is equivalent-- we should regard light is always composed of two quadrature components. You can say, the cosine, the sine oscillation, the x, and the p. And the squeezing is squeezed in one quadrature, by it is elongated in the other quadrature. Therefore, we want to be phase-sensitive. We want to pick out either the cosine omega T, or the sine omega T oscillation. This is sort of, homodyne detection. We'll discuss it on Monday. Any question? OK. Good.

Quantum physical background

An oscillating physical quantity cannot have precisely defined values at all phases of the oscillation. This is true for the electric and magnetic fields of an electromagnetic wave, as well as for any other wave or oscillation (see figure right). This fact can be observed in experiments and is described by quantum theory. For electromagnetic waves usually just the electric field is considered, because it is the one that mainly interacts with matter.

Fig. 1. shows five different quantum states that a monochromatic wave could be in. The difference of the five quantum states is given by different electric field excitations and by different distributions of the quantum uncertainty along the phase . For a displaced coherent state, the expectation (mean) value of the electric field shows an oscillation, with an uncertainty independent of the phase (a). Also the phase- (b) and amplitude-squeezed states (c) show an oscillation of the mean electric field, but here the uncertainty depends on phase and is squeezed for some phases. The vacuum state (d) is a special coherent state and is not squeezed. It has zero mean electric field for all phases and a phase-independent uncertainty. It has zero energy on average, i.e. zero photons, and is the ground state of the monochromatic wave we consider. Finally, a squeezed vacuum state has also a zero mean electric field but a phase-dependent uncertainty (e).

Generally, quantum uncertainty reveals itself through a large number of identical measurements on identical quantum objects (here: modes of light) that, however, give different results. Let us again consider a continuous-wave monochromatic light wave (as emitted by an ultra-stable laser). A single measurement of Ԑ is performed over many periods of the light wave and provides a single number. The next measurements of Ԑ will be done consecutively on the same laser beam. Having recorded a large number of such measurements we know the field uncertainty at . In order to get the full picture, and for instance Fig.1(b), we need to record the statistics at many different phases .

Quantitative description of (squeezed) uncertainty

The measured electric field strengths at the wave's phase are the eigenvalues of the normalized quadrature operator , defined as[5]

where and are the annihilation and creation operators, respectively, of the oscillator representing the photon. is the wave's amplitude quadrature, equivalent to the position in optical phase space, and is the wave's phase quadrature, equivalent to momentum. and are non-commuting observables. Although they represent electric fields, they are dimensionless and satisfy the following uncertainty relation:[6]
,

where stands for the variance. (The variance is the mean of the squares of the measuring values minus the square of the mean of the measuring values.) If a mode of light is in its ground state (having an average photon number of zero), the uncertainty relation above is saturated and the variances of the quadrature are . (Other normalizations can also be found in literature. The normalization chosen here has the nice property that the sum of the ground state variances directly provide the zero point excitation of the quantized harmonic oscillator ).

Definition: Light is in a squeezed state, if (and only if) a phase exists for which .[6][7]

While coherent states belong to the semi-classical states, since they can be fully described by a semi-classical model,[8] squeezed states of light belong to the so-called nonclassical states, which also include number states (Fock states) and Schrödinger cat states.

Squeezed states (of light) were first produced in the mid-1980s.[9][10] At that time, quantum noise squeezing by up to a factor of about 2 (3 dB) in variance was achieved, i.e. . Today, squeeze factors larger than 10 (10 dB) have been directly observed.[11][12][13] A limitation is set by decoherence, mainly in terms of optical loss.[8]

The squeeze factor in Decibel (dB) can be computed in the following way:

, where is the smallest variance when varying the phase from 0 to . This particular phase is called the squeeze angle.

Representation of squeezed states by quasi-probability densities

Fig. 1(f): Left: Wigner function of a squeezed vacuum state. Right: Connection to Fig. 1 (e).

Quantum states such as those in Fig. 1 (a) to (e) are often displayed as Wigner functions, which are quasi-probability density distributions. Two orthogonal quadratures, usually and , span a phase space diagram, and the third axes provides the quasi probability of yielding a certain combination of . Since and are not precisely defined simultaneously, we cannot talk about a 'probability' as we do in classical physics but call it a 'quasi probability'. A Wigner function is reconstructed from time series of and . The reconstruction is also called 'quantum tomographic reconstruction'. For squeezed states, the Wigner function has a Gaussian shape, with an elliptical contour line, see Fig.: 1(f).

Physical meaning of measurement quantity and measurement object

Quantum uncertainty becomes visible when identical measurements of the same quantity (observable) on identical objects (here: modes of light) give different results (eigen values). In case of a single freely propagating monochromatic laser beam, the individual measurements are performed on consecutive time intervals of identical length. One interval must last much longer than the light's period; otherwise the monochromatic property would be significantly disturbed. Such consecutive measurements correspond to a time series of fluctuating eigen values. Consider an example in which the amplitude quadrature was repeatedly measured. The time series can be used for a quantum statistical characterization of the modes of light. Obviously, the amplitude of the light wave might be different before and after our measurement, i.e. the time series does not provide any information about very slow changes of the amplitude, which corresponds to very low frequencies. This is a trivial but also fundamental issue, since any data taking lasts for a finite time. Our time series, however, does provide meaningful information about fast changes of the light's amplitude, i.e. changes at frequencies higher than the inverse of the full measuring time. Changes that are faster than the duration of a single measurement, however, are invisible again. A quantum statistical characterization through consecutive measurements on some sort of a carrier is thus always related to a specific frequency interval, for instance described by with Based on this insight, we can describe the physical meaning of the observable more clearly:[8]

Fig. 2: Normalized variances of modulation states of the same carrier light beam versus modulation frequency . Here, the measurement band width is about 10 kHz. Each trace therefore describes about 200 mutually independent modulation modes.

The quantum statistical characterization using identical consecutive modes carried by a laser beam confers to the laser beam's electric field modulation within a frequency interval. The actual observable needs to be labeled accordingly, for instance as . is the amplitude (or depth) of the amplitude modulation and the amplitude (or depth) of the phase modulation in the respective frequency interval. This leads to the doggerel expressions 'amplitude quadrature amplitude' and 'phase quadrature amplitude'.

Within some limitations, for instance set by the speed of the electronics, and can be freely chosen in course of data acquisition and, in particular, data processing. This choice also defines the measurement object, i.e. the mode that is characterized by the statistics of the eigen values of and . The measurement object thus is a modulation mode that is carried by the light beam. – In many experiments, one is interested in a continuous spectrum of many modulation modes carried by the same light beam.[14] Fig. 2 shows the squeeze factors of many neighboring modulation modes versus . The upper trace refers to the uncertainties of the same modes being in their vacuum states, which serves as the 0 dB reference.

The observables in squeezed light experiments correspond exactly to those being used in optical communication. Amplitude modulation (AM) and frequency modulation (FM) are the classical means to imprint information on a carrier field. (Frequency modulation is mathematically closely related to phase modulation). The observables and also correspond to the measurement quantities in laser interferometers, such as in Sagnac interferometers measuring rotation changes and in Michelson interferometers observing gravitational waves. Squeezed states of light thus have ample applications in optical communication and optical measurements. The most prominent and important application is in gravitational-wave observatories.[15][16][8] Arguably, it is the first end-user driven application of quantum correlations.[17] Squeezed light originally was not planned to be implemented in either Advanced LIGO nor in Advanced Virgo, but now it contributes a significant factor towards the observatories design sensitivities and increases the rate of observed gravitational-wave events.[2][3]

Frequency-dependent squeezing

Strain noise spectra of the LIGO (Hanford) detector in amplitude spectral density units for frequency-dependent squeezing (purple), frequency-independent squeezing (green) and no squeezing (black)[18][19]

Frequency-dependent squeezing is a method being implemented at the LIGO–Virgo–KAGRA collaboration to improve sensitivity using its 300 m long filter cavities to handle light differently according to frequencies which allows to improve accuracy of phases at high frequencies at the cost of more inaccuracy in amplitudes at low frequencies and equivalently better amplitudes at low frequencies but worse phases at high frequencies, manipulating the uncertainty relation by the measurement of interest.[20][21]

Noise at high frequencies is dominated by shot noise while at low frequencies is dominated by radiation pressure noise so when one source is reduced the other increases.[22]

Applications

Optical high-precision measurements

Fig. 3: Schematic of a laser interferometer for the detection of gravitational waves. Here, squeezed vacuum states are injected and overlapped with the bright field at the central beam splitter to improve the sensitivity.
Fig. 4: Photo voltages of a photo diode detecting light.

Squeezed light is used to reduce the photon counting noise (shot noise) in optical high-precision measurements, most notably in laser interferometers. There are a large number of proof-of-principle experiments.[23][24] Laser interferometers split a laser beam in two paths and overlap them again afterwards. If the relative optical path length changes, the interference changes, and the light power in the interferometer's output port as well. This light power is detected with a photo diode providing a continuous voltage signal. If for instance the position of one interferometer mirror vibrates and thereby causes an oscillating path length difference, the output light has an amplitude modulation of the same frequency. Independent of the existence of such a (classical) signal, a beam of light always carries at least the vacuum state uncertainty (see above). The (modulation) signal with respect to this uncertainty can be improved by using a higher light power inside the interferometer arms, since the signal increases with the light power. This is the reason (in fact the only one) why Michelson interferometers for the detection of gravitational waves use very high optical power. High light power, however, produces technical problems. Mirror surfaces absorb parts of the light, become warmer, get thermally deformed and reduce the interferometer's interference contrast. Furthermore, an excessive light power can excite unstable mechanical vibrations of the mirrors. These consequences are mitigated if squeezed states of light are used for improving the signal-to-noise-ratio. Squeezed states of light do not increase the light's power. They also do not increase the signal, but instead reduce the noise.[8]

Laser interferometers are usually operated with monochromatic continuous-wave light. The optimal signal-to-noise-ratio is achieved by either operating the differential interferometer arm lengths such that both output ports contain half of the input light power (half fringe) and by recording the difference signal from both ports, or by operating the interferometer close to a dark fringe for one of the output ports where just a single photodiode is placed.[7] The latter operation point is used in gravitational-wave (GW) detectors.

For improving an interferometer sensitivity with squeezed states of light, the already existing bright light does not need to be fully replaced. What has to be replaced is just the vacuum uncertainty in the difference of the phase quadrature amplitudes of the light fields in the arms, and only at modulation frequencies at which signals are expected. This is achieved by injecting a (broadband) squeezed vacuum field (Fig. 1e) into the unused interferometer input port (Fig. 3). Ideally, perfect interference with the bright field is achieved. For this the squeezed field has to be in the same mode as the bright light, i.e. has to have the same wavelength, same polarisation, same wavefront curvature, same beam radius, and, of course, the same directions of propagation in the interferometer arms. For the squeezed-light enhancement of a Michelson interferometer operated at dark fringe, a polarising beam splitter in combination with a Faraday rotator is required. This combination constitutes an optical diode. Without any loss, the squeezed field is overlapped with the bright field at the interferometer's central beam splitter, is split and travels along the arms, is retro-reflected, constructively interferes and overlaps with the interferometer signal towards the photo diode. Due to the polarisation rotation of the Faraday rotator, the optical loss on signal and squeezed field is zero (in the ideal case). Generally, the purpose of an interferometer is to transform a differential phase modulation (of two light beams) into an amplitude modulation of the output light . Accordingly, the injected vacuum-squeezed field is injected such that the differential phase quadrature uncertainty in the arms is squeezed. On the output light amplitude quadrature squeezing is observed. Fig. 4 shows the photo voltage of the photo diode in the interferometer output port. Subtracting the constant offset provides the (GW) signal.

A source of squeezed states of light were integrated in the gravitational-wave detector GEO600 in 2010,[16] as shown in Fig. 4. The source was built by the research group of R. Schnabel at Leibniz Universität Hannover (Germany).[25] With squeezed light, the sensitivity of GEO600 during observational runs has been increased to values, which for practical reasons were not achievable without squeezed light.[26] In 2018, squeezed light upgrades are also planned for the gravitational wave detectors Advanced LIGO and Advanced Virgo.

Going beyond squeezing of photon counting noise, squeezed states of light can also be used to correlate quantum measurement noise (shot noise) and quantum back action noise to achieve sensitivities in the quantum non-demolition (QND) regime.[27][28]

Radiometry and calibration of quantum efficiencies

Squeezed light can be used in radiometry to calibrate the quantum efficiency of photo-electric photo detectors without a lamp of calibrated radiance.[12] Here, the term photo detector refers to a device that measures the power of a bright beam, typically in the range from a few microwatts up to about 0.1 W. The typical example is a PIN photo diode. In case of perfect quantum efficiency (100%), such a detector is supposed to convert every photon energy of incident light into exactly one photo electron. Conventional techniques of measuring quantum efficiencies require the knowledge of how many photons hit the surface of the photo detector, i.e. they require a lamp of calibrated radiance. The calibration on the basis of squeezed states of light uses instead the effect, that the uncertainty product increases the smaller the quantum uncertainty of the detector is. In other words: The squeezed light method uses the fact that squeezed states of light are sensitive against decoherence. Without any decoherence during generation, propagation and detection of squeezed light, the uncertainty product has its minimum value of 1/16 (see above). If optical loss is the dominating decoherence effect, which usually is the case, the independent measurement of all optical losses during generation and propagation together with the value of the uncertainty product directly reveals the quantum uncertainty of the photo detectors used.[12]

When a squeezed state with squeezed variance is detected with a photo detector of quantum efficiency (with ), the actually observed variance is increased to

Optical loss mixes a portion of the vacuum state variance to the squeezed variance, which decreases the squeeze factor. The same equation also describes the influence of a non-perfect quantum efficiency on the anti-squeezed variance. The anti-squeezed variance reduces, however, the uncertainty product increases. Optical loss on a pure squeezed state produces a mixed squeezed state.

Entanglement-based quantum key distribution

Fig. 5: Measurement results on two EPR entangled light fields. The measurement values taken on one subsystem (at A) and on the other subsystem (at B) vary a lot, i.e. show a large local uncertainty. Comparing the data as shown here reveals correlations (top, blue) or anti-correlations (bottom, blue). In this example, correlations as well as anti-correlations are stronger than the vacuum state uncertainty (black).

Squeezed states of light can be used to produce Einstein-Podolsky-Rosen-entangled light that is the resource for a high quality level of quantum key distribution (QKD), which is called 'one-sided device independent QKD'.[29]

Superimposing on a balanced beam splitter two identical light beams that carry squeezed modulation states and have a propagation length difference of a quarter of their wavelength produces two EPR entangled light beams at the beam splitter output ports. Quadrature amplitude measurements on the individual beams reveal uncertainties that are much larger than those of the ground states, but the data from the two beams show strong correlations: from a measurement value taken at the first beam (), one can infer the corresponding measurement value taken at the second beam (). If the inference shows an uncertainty smaller than that of the vacuum state, EPR correlations exist, see Fig. 5.

The aim of quantum key distribution is the distribution of identical, true random numbers to two distant parties A and B in such a way that A and B can quantify the amount of information about the numbers that has been lost to the environment (and thus is potentially in hand of an eavesdropper). To do so, sender (A) sends one of the entangled light beams to receiver (B). A and B measure repeatedly and simultaneously (taking the different propagation times into account) one of two orthogonal quadrature amplitudes. For every single measurement they need to choose whether to measure or in a truly random way, independently from each other. By chance, they measure the same quadrature in 50% of the single measurements. After having performed a large number of measurements, A and B communicate (publicly) what their choice was for every measurement. The non-matched pairs are discarded. From the remaining data they make public a small but statistically significant amount to test whether B is able to precisely infer the measurement results at A. Knowing the characteristics of the entangled light source and the quality of the measurement at the sender site, the sender gets information about the decoherence that happened during channel transmission and during the measurement at B. The decoherence quantifies the amount of information that was lost to the environment. If the amount of lost information is not too high and the data string not too short, data post processing in terms of error correction and privacy amplification produces a key with an arbitrarily reduced epsilon-level of insecurity. In addition to conventional QKD, the test for EPR correlations not only characterizes the channel over which the light was sent (for instance a glas fibre) but also the measurement at the receiver site. The sender does not need to trust the receivers measurement any more. This higher quality of QKD is called one-sided device independent. This type of QKD works if the natural decoherence is not too high. For this reason, an implementation that uses conventional telecommunication glas fibers would be limited to a distance of a few kilometers.[29]

Generation

Fig. 6: Schematic of a squeezing resonator. The pumped nonlinear crystal inside the resonator attenuates the electric field at optical frequency . This leads to perfect destructive interference for one quadrature angle that is carried by the optical frequency and propagates towards the left (left side of resonator). The pump light enters from the right and is simply retro-reflected. If the pump light intensity is kept below the resonator's oscillation threshold, its input and output powers are basically identical.
Timeline of experimentally achieved light squeezing values in the laboratory. Since the first demonstration in 1985 values have steadily improved.

Squeezed light is produced by means of nonlinear optics. The most successful method uses degenerate type I optical-parametric down-conversion (also called optical-parametric amplification) inside an optical resonator. To squeeze modulation states with respect to a carrier field at optical frequency , a bright pump field at twice the optical frequency is focussed into a nonlinear crystal that is placed between two or more mirrors forming an optical resonator. It is not necessary to inject light at frequency . (Such light, however, is required for detecting the (squeezed) modulation states). The crystal material needs to have a nonlinear susceptibility and needs to be highly transparent for both optical frequencies used. Typical materials are lithium niobate (LiNbO3) and (periodically poled) potassium titanyl phosphate (KTP). Due to the nonlinear susceptibility of the pumped crystal material, the electric field at frequency is amplified and deamplified, depending on the relative phase to the pump light. At the pump's electric field maxima, the electric field at frequency is amplified. At the pump's electric field minima, the electric field at frequency is squeezed. This way, the vacuum state (Fig. 1e) is transferred to a squeezed vacuum state (Fig. 1d). A displaced coherent state (Fig. 1a) is transferred to a phase squeezed state (Fig. 1b) or to an amplitude squeezed state (Fig. 1c), depending on the relative phase between coherent input field and pump field. A graphical description of these processes can be found in.[8]

The existence of a resonator for the field at is essential. The task of the resonator is shown in Fig. 6. The left resonator mirror has a typical reflectivity of about . Correspondingly of the electric field that (continuously) enters from the left gets reflected. The remaining part is transmitted and resonates between the two mirrors. Due to the resonance, the electric field inside the resonator gets enhanced (even without any medium inside). of the steady-state light power inside the resonator gets transmitted towards the left and interferes with the beam that was retro-reflected directly. For an empty loss-less resonator, 100% of the light power would eventually propagate towards the left, obeying energy conservation.

The principle of the squeezing resonator is the following: The medium parametrically attenuates the electric field inside the resonator to such a value that perfect destructive interference is achieved outside the resonator for the attenuated field quadrature. The optimum field attenuation factor inside the resonator is slightly below 2, depending on the reflectivity of the resonator mirror.[8] This principle also works for electric field uncertainties. Inside the resonator, the squeeze factor is always less than 6 dB, but outside the resonator it can be arbitrarily high. If quadrature is squeezed, quadrature is anti-squeezed – inside as well as outside the resonator. It can be shown that the highest squeeze factor for one quadrature is achieved if the resonator is at its threshold for the orthogonal quadrature. At threshold and above, the pump field is converted into a bright field at optical frequency . Squeezing resonators are usually operated slightly below threshold, for instance, to avoid damage to the photo diodes due to the bright down-converted field.

A squeezing resonator works efficiently at modulation frequencies well inside its linewidth. Only for these frequencies highest squeeze factors can be achieved. At frequencies the optical-parametric gain is strongest, and the time delay between the interfering parts negligible. If decoherence was zero, infinite squeeze factors could be achieved outside the resonator, although the squeeze factor inside the resonator was less than 6 dB. Squeezing resonators have typical linewidths of a few tens of MHz up to GHz.[30]

Due to the interest in the interaction between squeezed light and atomic ensemble, narrowband atomic resonance squeezed light have been also generated through crystal[31] and the atomic medium.[32]

Detection

Fig. 7: Balanced homodyne detector. LO: local oscillator; PD: photo diode.

Squeezed states of light can be fully characterized by a photo-electric detector that is able to (subsequently) measure the electric field strengths at any phase . (The restriction to a certain band of modulation frequencies happens after the detection by electronic filtering.) The required detector is a balanced homodyne detector (BHD). It has two input ports for two light beams. One for the (squeezed) signal field, and another for the BHDs local oscillator (LO) having the same wavelength as the signal field. The LO is part of the BHD. Its purpose is to beat with the signal field and to optically amplify it. Further components of the BHD are a balanced beam splitter and two photo diodes (of high quantum efficiency). Signal beam and LO need to be overlapped at the beam splitter. The two interference results in the beam splitter output ports are detected and the difference signal recorded (Fig. 7). The LO needs to be much more intense than the signal field. In this case the differential signal from the photo diodes in the interval is proportional to the quadrature amplitude . Changing the differential propagation length before the beam splitter sets the quadrature angle to an arbitrary value. (A change by a quarter of the optical wavelength changes the phase by  .)

The following should be stated at this point: Any information about the electro-magnetic wave can only be gathered in a quantized way, i.e. by absorbing light quanta (photons). This is also true for the BHD. However, a BHD cannot resolve the discrete energy transfer from the light to the electric current, since in any small time interval a vast number of photons are detected. This is ensured by the intense LO. The observable therefore has a quasi-continuous eigenvalue spectrum, as it is expected for an electric field strength. (In principle, one can also characterize squeezed states, in particular squeezed vacuum states, by counting photons, however, in general the measurement of the photon number statistic is not sufficient for a full characterization of a squeezed state and the full density matrix in the basis of the number states has to be determined.)

See also

References

  1. ^ Walls, D. F. (1983). "Squeezed states of light". Nature. 306 (5939): 141–146. Bibcode:1983Natur.306..141W. doi:10.1038/306141a0. ISSN 1476-4687. S2CID 4325386.
  2. ^ a b Tse, M.; Yu, Haocun; Kijbunchoo, N.; Fernandez-Galiana, A.; Dupej, P.; Barsotti, L.; Blair, C. D.; Brown, D. D.; Dwyer, S. E.; Effler, A.; Evans, M. (December 5, 2019). "Quantum-Enhanced Advanced LIGO Detectors in the Era of Gravitational-Wave Astronomy". Physical Review Letters. 123 (23): 231107. Bibcode:2019PhRvL.123w1107T. doi:10.1103/physrevlett.123.231107. hdl:1721.1/136579.2. ISSN 0031-9007. PMID 31868462.
  3. ^ a b Acernese, F.; Agathos, M.; Aiello, L.; Allocca, A.; Amato, A.; Ansoldi, S.; Antier, S.; Arène, M.; Arnaud, N.; Ascenzi, S.; Astone, P. (December 5, 2019). "Increasing the Astrophysical Reach of the Advanced Virgo Detector via the Application of Squeezed Vacuum States of Light". Physical Review Letters. 123 (23): 231108. Bibcode:2019PhRvL.123w1108A. doi:10.1103/physrevlett.123.231108. hdl:11573/1343006. ISSN 0031-9007. PMID 31868444.
  4. ^ Abbott, R. (2021). "GWTC-2: Compact Binary Coalescences Observed by LIGO and Virgo during the First Half of the Third Observing Run". Physical Review X. 11 (2): 021053. arXiv:2010.14527. Bibcode:2021PhRvX..11b1053A. doi:10.1103/PhysRevX.11.021053. S2CID 225094244.
  5. ^ Schleich, Wolfgang P. (2001). Quantum Optics in Phase Space (1 ed.). Berlin: WILEY-VCH Verlag Berlin GmbH. ISBN 3-527-29435-X.
  6. ^ a b Gerry, Christopher; Knight, Peter (2004). Introductory Quantum Optics. Cambridge University Press. doi:10.1017/cbo9780511791239. ISBN 9780521527354. S2CID 209908717.
  7. ^ a b Bachor, Hans-Albert; Ralph, Tim C. (2004). A Guide to Experiments in Quantum Optics, Second Edition - Wiley Online Library. doi:10.1002/9783527619238. ISBN 9783527619238. S2CID 118649942.
  8. ^ a b c d e f g Schnabel, Roman (2017). "Squeezed states of light and their applications in laser interferometers". Physics Reports. 684: 1–51. arXiv:1611.03986. Bibcode:2017PhR...684....1S. doi:10.1016/j.physrep.2017.04.001. S2CID 119098759.
  9. ^ R. E. Slusher et al., Observation of squeezed states generated by four wave mixing in an optical cavity, Phys. Rev. Lett. 55 (22), 2409 (1985)
  10. ^ Wu, Ling-An (1986). "Generation of Squeezed States by Parametric Down Conversion" (PDF). Physical Review Letters. 57 (20): 2520–2523. Bibcode:1986PhRvL..57.2520W. doi:10.1103/physrevlett.57.2520. PMID 10033788.
  11. ^ Vahlbruch, Henning; Mehmet, Moritz; Chelkowski, Simon; Hage, Boris; Franzen, Alexander; Lastzka, Nico; Goßler, Stefan; Danzmann, Karsten; Schnabel, Roman (January 23, 2008). "Observation of Squeezed Light with 10-dB Quantum-Noise Reduction". Physical Review Letters. 100 (3): 033602. arXiv:0706.1431. Bibcode:2008PhRvL.100c3602V. doi:10.1103/PhysRevLett.100.033602. PMID 18232978. S2CID 3938634.
  12. ^ a b c Vahlbruch, Henning; Mehmet, Moritz; Danzmann, Karsten; Schnabel, Roman (September 6, 2016). "Detection of 15 dB Squeezed States of Light and their Application for the Absolute Calibration of Photoelectric Quantum Efficiency". Physical Review Letters. 117 (11): 110801. Bibcode:2016PhRvL.117k0801V. doi:10.1103/PhysRevLett.117.110801. hdl:11858/00-001M-0000-002B-87B5-3. PMID 27661673.
  13. ^ Schönbeck, Axel; Thies, Fabian; Schnabel, Roman (January 1, 2018). "13 dB squeezed vacuum states at 1550 nm from 12 mW external pump power at 775 nm". Optics Letters. 43 (1): 110–113. arXiv:2005.09891. Bibcode:2018OptL...43..110S. doi:10.1364/OL.43.000110. ISSN 1539-4794. PMID 29328207. S2CID 23252788.
  14. ^ G. Breitenbach, F. Illuminati, S. Schiller, J. Mlynek (October 15, 1998), "Broadband detection of squeezed vacuum: A spectrum of quantum states", Europhysics Letters (EPL) (in German), vol. 44, no. 2, pp. 192–197, arXiv:quant-ph/9901044, Bibcode:1998EL.....44..192B, doi:10.1209/epl/i1998-00456-2, ISSN 0295-5075, S2CID 13119216{{citation}}: CS1 maint: multiple names: authors list (link)
  15. ^ Caves, Carlton M. (April 15, 1981). "Quantum-mechanical noise in an interferometer". Physical Review D. 23 (8): 1693–1708. Bibcode:1981PhRvD..23.1693C. doi:10.1103/physrevd.23.1693. ISSN 0556-2821.
  16. ^ a b The LIGO Scientific Collaboration (2011). "A gravitational wave observatory operating beyond the quantum shot-noise limit". Nature Physics. 7 (12): 962–965. arXiv:1109.2295. Bibcode:2011NatPh...7..962L. doi:10.1038/nphys2083. ISSN 1745-2481. S2CID 209832912.
  17. ^ Schnabel, Roman (January 29, 2020). ""Quantum Weirdness" in Exploitation by the International Gravitational-Wave Observatory Network". Annalen der Physik. 532 (3): 1900508. arXiv:1909.13723. Bibcode:2020AnP...53200508S. doi:10.1002/andp.201900508. ISSN 0003-3804.
  18. ^ LIGO O4 Detector Collaboration; Ganapathy, D.; Jia, W.; Nakano, M.; Xu, V.; Aritomi, N.; Cullen, T.; Kijbunchoo, N.; Dwyer, S. E.; Mullavey, A.; McCuller, L.; Abbott, R.; Abouelfettouh, I.; Adhikari, R. X.; Ananyeva, A. (October 30, 2023). "Broadband Quantum Enhancement of the LIGO Detectors with Frequency-Dependent Squeezing". Physical Review X. 13 (4): 041021. Bibcode:2023PhRvX..13d1021G. doi:10.1103/PhysRevX.13.041021. hdl:2440/140376.{{cite journal}}: CS1 maint: numeric names: authors list (link)
  19. ^ "LIGO Surpasses the Quantum Limit". LIGO Lab | Caltech. Retrieved November 15, 2023.
  20. ^ Virgo Collaboration; Acernese, F.; Agathos, M.; Ain, A.; Albanesi, S.; Alléné, C.; Allocca, A.; Amato, A.; Amra, C.; Andia, M.; Andrade, T.; Andres, N.; Andrés-Carcasona, M.; Andrić, T.; Ansoldi, S. (July 25, 2023). "Frequency-Dependent Squeezed Vacuum Source for the Advanced Virgo Gravitational-Wave Detector". Physical Review Letters. 131 (4): 041403. Bibcode:2023PhRvL.131d1403A. doi:10.1103/PhysRevLett.131.041403. hdl:11568/1196710. PMID 37566847. S2CID 260185660.
  21. ^ Polini, E (August 1, 2021). "Broadband quantum noise reduction via frequency dependent squeezing for Advanced Virgo Plus". Physica Scripta. 96 (8): 084003. Bibcode:2021PhyS...96h4003P. doi:10.1088/1402-4896/abfef0. ISSN 0031-8949. S2CID 235285860.
  22. ^ Zhao, Yuhang; Aritomi, Naoki; Capocasa, Eleonora; Leonardi, Matteo; Eisenmann, Marc; Guo, Yuefan; Polini, Eleonora; Tomura, Akihiro; Arai, Koji; Aso, Yoichi; Huang, Yao-Chin; Lee, Ray-Kuang; Lück, Harald; Miyakawa, Osamu; Prat, Pierre (April 28, 2020). "Frequency-Dependent Squeezed Vacuum Source for Broadband Quantum Noise Reduction in Advanced Gravitational-Wave Detectors". Physical Review Letters. 124 (17): 171101. arXiv:2003.10672. Bibcode:2020PhRvL.124q1101Z. doi:10.1103/PhysRevLett.124.171101. PMID 32412296. S2CID 214623227.
  23. ^ Xiao, Min; Wu, Ling-An; Kimble, H. J. (July 20, 1987). "Precision measurement beyond the shot-noise limit" (PDF). Physical Review Letters. 59 (3): 278–281. Bibcode:1987PhRvL..59..278X. doi:10.1103/PhysRevLett.59.278. PMID 10035719. S2CID 120065263.
  24. ^ Grangier, P.; Slusher, R. E.; Yurke, B.; LaPorta, A. (November 9, 1987). "Squeezed-light-enhanced polarization interferometer". Physical Review Letters. 59 (19): 2153–2156. Bibcode:1987PhRvL..59.2153G. doi:10.1103/PhysRevLett.59.2153. PMID 10035438.
  25. ^ Vahlbruch, Henning; Khalaidovski, Alexander; Lastzka, Nico; Gräf, Christian; Danzmann, Karsten; Schnabel, Roman (2010). "The GEO 600 squeezed light source". Classical and Quantum Gravity. 27 (8): 084027. arXiv:1004.4975. Bibcode:2010CQGra..27h4027V. doi:10.1088/0264-9381/27/8/084027. S2CID 53991930.
  26. ^ Grote, H.; Danzmann, K.; Dooley, K. L.; Schnabel, R.; Slutsky, J.; Vahlbruch, H. (May 1, 2013). "First Long-Term Application of Squeezed States of Light in a Gravitational-Wave Observatory". Physical Review Letters. 110 (18): 181101. arXiv:1302.2188. Bibcode:2013PhRvL.110r1101G. doi:10.1103/PhysRevLett.110.181101. PMID 23683187. S2CID 3566080.
  27. ^ Jaekel, M. T.; Reynaud, S. (1990). "Quantum Limits in Interferometric Measurements". EPL (Europhysics Letters). 13 (4): 301–306. arXiv:quant-ph/0101104. Bibcode:1990EL.....13..301J. doi:10.1209/0295-5075/13/4/003. ISSN 0295-5075. S2CID 119056243.
  28. ^ Kimble, H. J.; Levin, Yuri; Matsko, Andrey B.; Thorne, Kip S.; Vyatchanin, Sergey P. (December 26, 2001). "Conversion of conventional gravitational-wave interferometers into quantum nondemolition interferometers by modifying their input and/or output optics". Physical Review D. 65 (2): 022002. arXiv:gr-qc/0008026. Bibcode:2001PhRvD..65b2002K. doi:10.1103/PhysRevD.65.022002. S2CID 15339393.
  29. ^ a b Gehring, Tobias; Händchen, Vitus; Duhme, Jörg; Furrer, Fabian; Franz, Torsten; Pacher, Christoph; Werner, Reinhard F.; Schnabel, Roman (October 30, 2015). "Implementation of continuous-variable quantum key distribution with composable and one-sided-device-independent security against coherent attacks". Nature Communications. 6: 8795. Bibcode:2015NatCo...6.8795G. doi:10.1038/ncomms9795. PMC 4640132. PMID 26514280.
  30. ^ Ast, Stefan; Mehmet, Moritz; Schnabel, Roman (June 3, 2013). "High-bandwidth squeezed light at 1550 nm from a compact monolithic PPKTP cavity". Optics Express. 21 (11): 13572–13579. arXiv:1303.1925. Bibcode:2013OExpr..2113572A. doi:10.1364/oe.21.013572. ISSN 1094-4087. PMID 23736610. S2CID 9258803.
  31. ^ Hétet, G.; Glöckl, O.; Pilypas, K. A.; Harb, C.C.; Buchler, B.C.; Bachor, H.-A.; Lam, P.K. (2006). "Squeezed light for bandwidth-limited atom optics experiments at the rubidium D1 line". Journal of Physics B. 40 (1): 221–226. arXiv:quant-ph/0611204. Bibcode:2007JPhB...40..221H. doi:10.1088/0953-4075/40/1/020. ISSN 0953-4075. S2CID 1882589.
  32. ^ Kim, Saesun; Marino, Alberto M. (December 10, 2018). "Generation of 87Rb resonant bright two-mode squeezed light with four-wave mixing". Optics Express. 26 (25): 33366–33375. arXiv:1806.04797. Bibcode:2018OExpr..2633366K. doi:10.1364/OE.26.033366. ISSN 1094-4087. PMID 30645489. S2CID 51685095.
This page was last edited on 11 March 2024, at 18:34
Basis of this page is in Wikipedia. Text is available under the CC BY-SA 3.0 Unported License. Non-text media are available under their specified licenses. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc. WIKI 2 is an independent company and has no affiliation with Wikimedia Foundation.