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
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 coherent state

From Wikipedia, the free encyclopedia

In physics, a squeezed coherent state is a quantum state that is usually described by two non-commuting observables having continuous spectra of eigenvalues. Examples are position and momentum of a particle, and the (dimension-less) electric field in the amplitude (phase 0) and in the mode (phase 90°) of a light wave (the wave's quadratures). The product of the standard deviations of two such operators obeys the uncertainty principle:

and , respectively.
Wigner phase space distribution of a squeezed state of light with ζ=0.5.

Trivial examples, which are in fact not squeezed, are the ground state of the quantum harmonic oscillator and the family of coherent states . These states saturate the uncertainty above and have a symmetric distribution of the operator uncertainties with in "natural oscillator units" and . (In literature different normalizations for the quadrature amplitudes are used. Here we use the normalization for which the sum of the ground state variances of the quadrature amplitudes directly provide the zero point quantum number ).

The term squeezed state is actually used for states with a standard deviation below that of the ground state for one of the operators or for a linear combination of the two. The idea behind this is that the circle denoting the uncertainty of a coherent state in the quadrature phase space (see right) has been "squeezed" to an ellipse of the same area.[1][2][3] Note that a squeezed state does not need to saturate the uncertainty principle.

Squeezed states of light were first produced in the mid 1980s.[4][5] At that time, quantum noise squeezing by up to a factor of about 2 (3 dB) in variance was achieved, i.e. . As of 2017, squeeze factors larger than 10 (10 dB) have been directly observed.[6][7][8]

YouTube Encyclopedic

  • 1/5
    Views:
    14 886
    4 595
    21 201
    2 372
    1 180
  • 4. Non-classical light, squeezing, Part 1
  • The Wave Hunters - 04: Squeezed light
  • 3. Quantum description of light, Part 1
  • 07. Quantum optics (Schrodinger equation, harmonic oscillator, coherent states, photon statistics)
  • 33/44 Squeezing obtained from NLO effects. Entanglement

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.

Mathematical definition

Animated position-wavefunction of an 2dB amplitude-squeezed coherent state of α=3.

The most general wave function that satisfies the identity above is the squeezed coherent state (we work in units with )

where are constants (a normalization constant, the center of the wavepacket, its width, and the expectation value of its momentum). The new feature relative to a coherent state is the free value of the width , which is the reason why the state is called "squeezed".

The squeezed state above is an eigenstate of a linear operator

and the corresponding eigenvalue equals . In this sense, it is a generalization of the ground state as well as the coherent state.

Operator representation

The general form of a squeezed coherent state for a quantum harmonic oscillator is given by

where is the vacuum state, is the displacement operator and is the squeeze operator, given by

where and are annihilation and creation operators, respectively. For a quantum harmonic oscillator of angular frequency , these operators are given by

For a real , (note that ,[9] where r is squeezing parameter),[clarification needed] the uncertainty in and are given by

Therefore, a squeezed coherent state saturates the Heisenberg uncertainty principle , with reduced uncertainty in one of its quadrature components and increased uncertainty in the other.

Examples

Depending on the phase angle at which the state's width is reduced, one can distinguish amplitude-squeezed, phase-squeezed, and general quadrature-squeezed states. If the squeezing operator is applied directly to the vacuum, rather than to a coherent state, the result is called the squeezed vacuum. The figures below[clarification needed] give a nice visual demonstration of the close connection between squeezed states and Heisenberg's uncertainty relation: Diminishing the quantum noise at a specific quadrature (phase) of the wave has as a direct consequence an enhancement of the noise of the complementary quadrature, that is, the field at the phase shifted by [clarification needed].

Measured quantum noise of the electric field (3π-interval shown for the first two states; 4π-interval for the last three states)
Oscillating wave packets of the five states.
Wigner functions of the five states. The ripples are due to experimental inaccuracies.
Different squeezed states of laser light in a vacuum depend on the phase of the light field.[10] Images from the top: (1) Vacuum state, (2) Squeezed vacuum state, (3) Phase-squeezed state (4) Arbitrary squeezed state (5) Amplitude-squeezed state

As can be seen in the illustrations, in contrast to a coherent state, the quantum noise for a squeezed state is no longer independent of the phase of the light wave. A characteristic broadening and narrowing of the noise during one oscillation period can be observed. The probability distribution of a squeezed state is defined as the norm squared of the wave function mentioned in the last paragraph. It corresponds to the square of the electric (and magnetic) field strength of a classical light wave. The moving wave packets display an oscillatory motion combined with the widening and narrowing of their distribution: the "breathing" of the wave packet. For an amplitude-squeezed state, the most narrow distribution of the wave packet is reached at the field maximum, resulting in an amplitude that is defined more precisely than the one of a coherent state. For a phase-squeezed state, the most narrow distribution is reached at field zero, resulting in an average phase value that is better defined than the one of a coherent state.

In phase space, quantum mechanical uncertainties can be depicted by the Wigner quasi-probability distribution. The intensity of the light wave, its coherent excitation, is given by the displacement of the Wigner distribution from the origin. A change in the phase of the squeezed quadrature results in a rotation of the distribution.

Photon number distributions and phase distributions

The squeezing angle, that is the phase with minimum quantum noise, has a large influence on the photon number distribution of the light wave and its phase distribution as well.

Experimental photon number distributions for an amplitude-squeezed state, a coherent state, and a phase squeezed state reconstructed from measurements of the quantum statistics. Bars refer to theory, dots to experimental values.[11]
Pegg-Barnett phase distribution of the three states

For amplitude squeezed light the photon number distribution is usually narrower than the one of a coherent state of the same amplitude resulting in sub-Poissonian light, whereas its phase distribution is wider. The opposite is true for the phase-squeezed light, which displays a large intensity (photon number) noise but a narrow phase distribution. Nevertheless, the statistics of amplitude squeezed light was not observed directly with photon number resolving detector due to experimental difficulty.[12]

Reconstructed and theoretical photon number distributions for a squeezed-vacuum state. A pure squeezed vacuum state would have no contribution from odd-photon-number states. The non-zero contribution in the above figure is because the detected state is not a pure state – losses in the setup convert the pure squeezed vacuum into a mixed state.[11] (source: link 1)

For the squeezed vacuum state the photon number distribution displays odd-even-oscillations. This can be explained by the mathematical form of the squeezing operator, that resembles the operator for two-photon generation and annihilation processes. Photons in a squeezed vacuum state are more likely to appear in pairs.

Classification

Based on the number of modes

Squeezed states of light are broadly classified into single-mode squeezed states and two-mode squeezed states,[13] depending on the number of modes of the electromagnetic field involved in the process. Recent studies have looked into multimode squeezed states showing quantum correlations among more than two modes as well.

Single-mode squeezed states

Single-mode squeezed states, as the name suggests, consists of a single mode of the electromagnetic field whose one quadrature has fluctuations below the shot noise level[clarification needed] and the orthogonal quadrature has excess noise. Specifically, a single-mode squeezed vacuum (SMSV) state can be mathematically represented as,

where the squeezing operator S is the same as introduced in the section on operator representations above. In the photon number basis, writing this can be expanded as,

which explicitly shows that the pure SMSV consists entirely of even-photon Fock state superpositions. Single mode squeezed states are typically generated by degenerate parametric oscillation in an optical parametric oscillator,[14] or using four-wave mixing.[4]

Two-mode squeezed states

Two-mode squeezing involves two modes of the electromagnetic field which exhibit quantum noise reduction below the shot noise level[clarification needed] in a linear combination of the quadratures of the two fields. For example, the field produced by a nondegenerate parametric oscillator above threshold shows squeezing in the amplitude difference quadrature. The first experimental demonstration of two-mode squeezing in optics was by Heidmann et al..[15] More recently, two-mode squeezing was generated on-chip using a four-wave mixing OPO above threshold.[16] Two-mode squeezing is often seen as a precursor to continuous-variable entanglement, and hence a demonstration of the Einstein-Podolsky-Rosen paradox in its original formulation in terms of continuous position and momentum observables.[17][18] A two-mode squeezed vacuum (TMSV) state can be mathematically represented as,

,

and, writing down , in the photon number basis as,[19]

If the individual modes of a TMSV are considered separately (i.e., ), then tracing over or absorbing one of the modes leaves the remaining mode in a thermal state

with an effective average number of photons .

Based on the presence of a mean field

Squeezed states of light can be divided into squeezed vacuum and bright squeezed light, depending on the absence or presence of a non-zero mean field (also called a carrier), respectively. An optical parametric oscillator operated below threshold produces squeezed vacuum, whereas the same OPO operated above threshold produces bright squeezed light. Bright squeezed light can be advantageous for certain quantum information processing applications as it obviates the need of sending local oscillator to provide a phase reference, whereas squeezed vacuum is considered more suitable for quantum enhanced sensing applications. The AdLIGO and GEO600 gravitational wave detectors use squeezed vacuum to achieve enhanced sensitivity beyond the standard quantum limit.[20][21]

Atomic spin squeezing

For squeezing of two-level neutral atom ensembles it is useful to consider the atoms as spin-1/2 particles with corresponding angular momentum operators defined as

where and is the single-spin operator in the -direction. Here will correspond to the population difference in the two level system, i.e. for an equal superposition of the up and down state . The plane represents the phase difference between the two states. This is also known as the Bloch sphere picture. We can then define uncertainty relations such as . For a coherent (unentangled) state, . Squeezing is here considered the redistribution of uncertainty from one variable (typically ) to another (typically ). If we consider a state pointing in the direction, we can define the Wineland criterion[22] for squeezing, or the metrological enhancement of the squeezed state as

.

This criterion has two factors, the first factor is the spin noise reduction, i.e. how much the quantum noise in is reduced relative to the coherent (unentangled) state. The second factor is how much the coherence (the length of the Bloch vector, ) is reduced due to the squeezing procedure. Together these quantities tell you how much metrological enhancement the squeezing procedure gives. Here, metrological enhancement is the reduction in averaging time or atom number needed to make a measurement of a specific uncertainty. 20 dB of metrological enhancement means the same precision measurement can be made with 100 times fewer atoms or 100 times shorter averaging time.

Experimental realizations

There has been a whole variety of successful demonstrations of squeezed states. The first demonstrations were experiments with light fields using lasers and non-linear optics (see optical parametric oscillator). This is achieved by a simple process of four-wave mixing with a crystal; similarly travelling wave phase-sensitive amplifiers generate spatially multimode quadrature-squeezed states of light when the crystal is pumped in absence of any signal. Sub-Poissonian current sources driving semiconductor laser diodes have led to amplitude squeezed light.[23]

Squeezed states have also been realized via motional states of an ion in a trap, phonon states in crystal lattices, and spin states in neutral atom ensembles.[24][25] Much progress has been made on the creation and observation of spin squeezed states in ensembles of neutral atoms and ions, which can be used to enhancement measurements of time, accelerations, fields, and the current state of the art for measurement enhancement[clarification needed] is 20 dB.[26][27][28][29] Generation of spin squeezed states have been demonstrated using both coherent evolution of a coherent spin state and projective, coherence-preserving measurements. Even macroscopic oscillators were driven into classical motional states that were very similar to squeezed coherent states. Current state of the art in noise suppression, for laser radiation using squeezed light, amounts to 15 dB (as of 2016),[30][7] which broke the previous record of 12.7 dB (2010).[31]

Applications

Squeezed states of the light field can be used to enhance precision measurements. For example, phase-squeezed light can improve the phase read out of interferometric measurements (see for example gravitational waves). Amplitude-squeezed light can improve the readout of very weak spectroscopic signals.[32]

Spin squeezed states of atoms can be used to improve the precision of atomic clocks.[33][34] This is an important problem in atomic clocks and other sensors that use small ensembles of cold atoms where the quantum projection noise represents a fundamental limitation to the precision of the sensor.[35]

Various squeezed coherent states, generalized to the case of many degrees of freedom, are used in various calculations in quantum field theory, for example Unruh effect and Hawking radiation, and generally, particle production in curved backgrounds and Bogoliubov transformations.

Recently, the use of squeezed states for quantum information processing in the continuous variables (CV) regime has been increasing rapidly.[36] Continuous variable quantum optics uses squeezing of light as an essential resource to realize CV protocols for quantum communication, unconditional quantum teleportation and one-way quantum computing.[37][38] This is in contrast to quantum information processing with single photons or photon pairs as qubits. CV quantum information processing relies heavily on the fact that squeezing is intimately related to quantum entanglement, as the quadratures of a squeezed state exhibit sub-shot-noise[clarification needed] quantum correlations.

See also

References

  1. ^ Loudon, Rodney, The Quantum Theory of Light (Oxford University Press, 2000), ISBN 0-19-850177-3
  2. ^ C W Gardiner and Peter Zoller, "Quantum Noise", 3rd ed, Springer Berlin 2004
  3. ^ Walls, D. F. (November 1983). "Squeezed states of light". Nature. 306 (5939): 141–146. Bibcode:1983Natur.306..141W. doi:10.1038/306141a0. ISSN 1476-4687. S2CID 4325386.
  4. ^ a b 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)
  5. ^ Wu, Ling-An (1986). "Generation of Squeezed States by Parametric Down Conversion" (PDF). Physical Review Letters (Submitted manuscript). 57 (20): 2520–2523. Bibcode:1986PhRvL..57.2520W. doi:10.1103/physrevlett.57.2520. PMID 10033788.
  6. ^ 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. hdl:11858/00-001M-0000-0013-623A-0. PMID 18232978. S2CID 3938634.
  7. ^ a b 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" (PDF). 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.
  8. ^ 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. ^ Walls, D.F. and G. J. Milburn, Quantum Optics.
  10. ^ Breitenbach, G.; Schiller, S.; Mlynek, J. (May 29, 1997). "Measurement of the quantum states of squeezed light" (PDF). Nature. 387 (6632): 471–475. Bibcode:1997Natur.387..471B. doi:10.1038/387471a0. S2CID 4259166.
  11. ^ a b G. Breitenbach, S. Schiller, and J. Mlynek, "Measurement of the quantum states of squeezed light", Nature, 387, 471 (1997)
  12. ^ kitagawa, Akira; Takeoka, Masahiro; Sasaki, Masahide; Chefles, Anthony (2006). "Entanglement evaluation with Fisher information". arXiv:quant-ph/0612099.
  13. ^ Lvovsky, A. I. (2014). "Squeezed light". arXiv:1401.4118 [quant-ph].
  14. ^ Wu, L.-A.; Xiao, M.; Kimble, H. J. (1987). "Squeezed states of light from an optical parametric oscillator". J. Opt. Soc. Am. B. 4 (10): 1465. Bibcode:1987JOSAB...4.1465W. doi:10.1364/JOSAB.4.001465.
  15. ^ Heidmann, A.; Horowicz, R.; Reynaud, S.; Giacobino, E.; Fabre, C.; Camy, G. (1987). "Observation of Quantum Noise Reduction on Twin Laser Beams". Physical Review Letters. 59 (22): 2555–2557. Bibcode:1987PhRvL..59.2555H. doi:10.1103/physrevlett.59.2555. PMID 10035582.
  16. ^ Dutt, A.; Luke, K.; Manipatruni, S.; Gaeta, A. L.; Nussenzveig, P.; Lipson, M. (2015). "On-Chip Optical Squeezing". Physical Review Applied. 3 (4): 044005. arXiv:1309.6371. Bibcode:2015PhRvP...3d4005D. doi:10.1103/physrevapplied.3.044005.
  17. ^ Ou, Z. Y.; Pereira, S. F.; Kimble, H. J.; Peng, K. C. (1992). "Realization of the Einstein-Podolsky-Rosen paradox for continuous variables" (PDF). Phys. Rev. Lett. (Submitted manuscript). 68 (25): 3663–3666. Bibcode:1992PhRvL..68.3663O. doi:10.1103/physrevlett.68.3663. PMID 10045765.
  18. ^ Villar, A. S.; Cruz, L. S.; Cassemiro, K. N.; Martinelli, M.; Nussenzveig, P. (2005). "Generation of Bright Two-Color Continuous Variable Entanglement". Phys. Rev. Lett. 95 (24): 243603. arXiv:quant-ph/0506139. Bibcode:2005PhRvL..95x3603V. doi:10.1103/physrevlett.95.243603. PMID 16384378. S2CID 13815567.
  19. ^ Schumaker, Bonny L.; Caves, Carlton M. (May 1, 1985). "New formalism for two-photon quantum optics. II. Mathematical foundation and compact notation". Physical Review A. 31 (5): 3093–3111. Bibcode:1985PhRvA..31.3093S. doi:10.1103/PhysRevA.31.3093. PMID 9895863.
  20. ^ Grote, H.; Danzmann, K.; Dooley, K. L.; Schnabel, R.; Slutsky, J.; Vahlbruch, H. (2013). "First Long-Term Application of Squeezed States of Light in a Gravitational-Wave Observatory". Phys. Rev. Lett. 110 (18): 181101. arXiv:1302.2188. Bibcode:2013PhRvL.110r1101G. doi:10.1103/physrevlett.110.181101. PMID 23683187. S2CID 3566080.
  21. ^ The LIGO Scientific Collaboration (2011). "A gravitational wave observatory operating beyond the quantum shot-noise limit". Nature Physics. 7 (12): 962. arXiv:1109.2295. Bibcode:2011NatPh...7..962L. doi:10.1038/nphys2083. S2CID 209832912.
  22. ^ Wineland, D. J.; Bollinger, J. J.; Heinzen, D. J. (July 1, 1994). "Squeezed atomic states and projection noise in spectroscopy". Physical Review A. 50 (2): 67–88. Bibcode:1994PhRvA..50...67W. doi:10.1103/PhysRevA.50.67. PMID 9910869.
  23. ^ Machida, S.; Yamamoto, Y.; Itaya, Y. (March 9, 1987). "Observation of amplitude squeezing in a constant-current driven semiconductor laser". Physical Review Letters. 58 (10): 1000–1003. Bibcode:1987PhRvL..58.1000M. doi:10.1103/PhysRevLett.58.1000. PMID 10034306.
  24. ^ O. V. Misochko, J. Hu, K. G. Nakamura, "Controlling phonon squeezing and correlation via one- and two-phonon interference," https://arxiv.org/abs/1011.2001
  25. ^ Ma, Jian; Wang, Xiaoguang; Sun, C.P.; Nori, Franco (December 2011). "Quantum spin squeezing". Physics Reports. 509 (2–3): 89–165. arXiv:1011.2978. Bibcode:2011PhR...509...89M. doi:10.1016/j.physrep.2011.08.003. S2CID 119239234.
  26. ^ Hosten, Onur; Engelsen, Nils J.; Krishnakumar, Rajiv; Kasevich, Mark A. (January 11, 2016). "Measurement noise 100 times lower than the quantum-projection limit using entangled atoms". Nature. 529 (7587): 505–8. Bibcode:2016Natur.529..505H. doi:10.1038/nature16176. PMID 26751056. S2CID 2139293.
  27. ^ Cox, Kevin C.; Greve, Graham P.; Weiner, Joshua M.; Thompson, James K. (March 4, 2016). "Deterministic Squeezed States with Collective Measurements and Feedback". Physical Review Letters. 116 (9): 093602. arXiv:1512.02150. Bibcode:2016PhRvL.116i3602C. doi:10.1103/PhysRevLett.116.093602. PMID 26991175. S2CID 29467218.
  28. ^ Bohnet, J. G.; Cox, K. C.; Norcia, M. A.; Weiner, J. M.; Chen, Z.; Thompson, J. K. (July 13, 2014). "Reduced spin measurement back-action for a phase sensitivity ten times beyond the standard quantum limit". Nature Photonics. 8 (9): 731–736. arXiv:1310.3177. Bibcode:2014NaPho...8..731B. doi:10.1038/nphoton.2014.151. S2CID 67780562.
  29. ^ Lücke, Bernd; Peise, Jan; Vitagliano, Giuseppe; Arlt, Jan; Santos, Luis; Tóth, Géza; Klempt, Carsten (April 17, 2014). "Detecting Multiparticle Entanglement of Dicke States". Physical Review Letters. 112 (15): 155304. arXiv:1403.4542. Bibcode:2014PhRvL.112o5304L. doi:10.1103/PhysRevLett.112.155304. PMID 24785048. S2CID 38230188.
  30. ^ Rini, Matteo (September 6, 2016). "Synopsis: A Tight Squeeze". Physics. 117 (11): 110801. Bibcode:2016PhRvL.117k0801V. doi:10.1103/PhysRevLett.117.110801. hdl:11858/00-001M-0000-002B-87B5-3. PMID 27661673.
  31. ^ Eberle, Tobias; Steinlechner, Sebastian; Bauchrowitz, Jöran; Händchen, Vitus; Vahlbruch, Henning; Mehmet, Moritz; Müller-Ebhardt, Helge; Schnabel, Roman (June 22, 2010). "Quantum Enhancement of the Zero-Area Sagnac Interferometer Topology for Gravitational Wave Detection". Physical Review Letters. 104 (25): 251102. arXiv:1007.0574. Bibcode:2010PhRvL.104y1102E. doi:10.1103/PhysRevLett.104.251102. PMID 20867358. S2CID 9929939.
  32. ^ Polzik, E. S. (January 1, 1992). "Spectroscopy with squeezed light" (PDF). Physical Review Letters (Submitted manuscript). 68 (20): 3020–3023. Bibcode:1992PhRvL..68.3020P. doi:10.1103/PhysRevLett.68.3020. PMID 10045587.
  33. ^ Leroux, Ian D.; Schleier-Smith, Monika H.; Vuletić, Vladan (June 25, 2010). "Orientation-Dependent Entanglement Lifetime in a Squeezed Atomic Clock". Physical Review Letters. 104 (25): 250801. arXiv:1004.1725. Bibcode:2010PhRvL.104y0801L. doi:10.1103/PhysRevLett.104.250801. PMID 20867356. S2CID 4514687.
  34. ^ Louchet-Chauvet, Anne; Appel, Jürgen; Renema, Jelmer J; Oblak, Daniel; Kjaergaard, Niels; Polzik, Eugene S (June 28, 2010). "Entanglement-assisted atomic clock beyond the projection noise limit". New Journal of Physics. 12 (6): 065032. arXiv:0912.3895. Bibcode:2010NJPh...12f5032L. doi:10.1088/1367-2630/12/6/065032. S2CID 119112907.
  35. ^ Kitagawa, Masahiro; Ueda, Masahito (June 1, 1993). "Squeezed spin states" (PDF). Physical Review A. 47 (6): 5138–5143. Bibcode:1993PhRvA..47.5138K. doi:10.1103/PhysRevA.47.5138. hdl:11094/77656. PMID 9909547.
  36. ^ Braunstein, Samuel L.; van Loock, Peter (June 29, 2005). "Quantum information with continuous variables". Reviews of Modern Physics. 77 (2): 513–577. arXiv:quant-ph/0410100. Bibcode:2005RvMP...77..513B. doi:10.1103/RevModPhys.77.513. S2CID 118990906.
  37. ^ Furusawa, A. (October 23, 1998). "Unconditional Quantum Teleportation". Science. 282 (5389): 706–709. Bibcode:1998Sci...282..706F. doi:10.1126/science.282.5389.706. PMID 9784123.
  38. ^ Menicucci, Nicolas C.; Flammia, Steven T.; Pfister, Olivier (September 22, 2008). "One-Way Quantum Computing in the Optical Frequency Comb". Physical Review Letters. 101 (13): 13501. arXiv:0804.4468. Bibcode:2008PhRvL.101m0501M. doi:10.1103/PhysRevLett.101.130501. PMID 18851426. S2CID 1307950.

External links

This page was last edited on 22 July 2023, at 16:52
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.