Next Article in Journal
Modified Godard Timing Recovery for Non Integer Oversampling Receivers
Previous Article in Journal
Remaining Useful Life Prediction of Hybrid Ceramic Bearings Using an Integrated Deep Learning and Particle Filter Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Higher‐Order Interactions in Quantum Optomechanics: Revisiting Theoretical Foundations

1
School of Electrical Engineering, Sharif University of Technology, P. O. Box 11365‐9363, Tehran, Iran
2
École Polytechnique Fédérale de Lausanne, Lausanne CH‐1015, Switzerland
Appl. Sci. 2017, 7(7), 656; https://doi.org/10.3390/app7070656
Submission received: 15 May 2017 / Revised: 13 June 2017 / Accepted: 23 June 2017 / Published: 24 June 2017
(This article belongs to the Section Optics and Lasers)

Abstract

:
The theory of quantum optomechanics is reconstructed from first principles by finding a Lagrangian from light’s equation of motion and then proceeding to the Hamiltonian. The nonlinear terms, including the quadratic and higher-order interactions, do not vanish under any possible choice of canonical parameters, and lead to coupling of momentum and field. The existence of quadratic mechanical parametric interaction is then demonstrated rigorously, which has been so far assumed phenomenologically in previous studies. Corrections to the quadratic terms are particularly significant when the mechanical frequency is of the same order or larger than the electromagnetic frequency. Further discussions on the squeezing as well as relativistic corrections are presented.

1. Introduction

The general field of quantum optomechanics is based on the standard optomechanical Hamiltonian, which is expressed as the simple product of photon number n ^ and the position x ^ operators, having the form OM = g 0 n ^ x ^ [1,2,3,4] with g 0 being the single-photon coupling rate. This is mostly referred to in a classical paper by Law [5], where the non-relativistic Hamiltonian is obtained through Lagrangian dynamics of the system. This basic interaction is behind numerous exciting theoretical and experimental studies, which demonstrate a wide range of applications. The optomechanical interaction OM is inherently nonlinear by its nature, which is quite analogous to the third-order Kerr optical effect in nonlinear optics [6,7]. These for instance include the optomechanical arrays [8,9,10,11,12,13,14,15,16,17], squeezing of phonon states [18,19,20], Heisenberg’s limited measurements [21], non-reciprocal optomechanical systems [22,23,24,25,26,27,28], sensing [29,30,31], engineered dissipation [32], engineered states [33], and non-reciprocal acousto-optical effects in optomechanical crystals [34,35,36].
Recent ideas in this field such as microwave-optical conversion [37,38,39,40], cavity electrooptics [41,42,43], optomechanical induced transparency [44,45,46,47], and optomechanical verification of Bell’s inequality [48] all emerge as closely related duals of quantum optomechanical systems, which are described within a completely identical framework. Furthermore, quantum chaos which had been predicted in quantum optomechanics [2,49] and cavity quantum electro-dynamical systems [50,51,52], has been recently observed in optomechanics [53,54].
Usually, the analysis in all these above works is ultimately done within the linearized approximation of ladder operators, thus being limited in accuracy where OM interaction is non-existent or vanishingly small. For applications where quadratic or even quartic effects are primarily pursued, g 0 may be designed to be identically zero [55,56,57,58,59,60,61], which urges need for accurate knowledge of higher-order interaction terms. Similar situation also could arise in trapped ultracold atomic gases [62], where linear interactions identically vanish. Relevant Physical phenomena in optomechanics such as four-wave mixing, also is suitably described by higher order interaction terms [63]. Moreover, significance and prominent role of such nonlinear interactions was observed in few recent experiments [64,65].
A careful review of the theory of this subject [5], however, reveals that there are a number of physical approximations in formulation of the problem such as the non-relativistic limit [66,67], which makes the unified description of relativistic photon momenta and non-relativistic mirror motion inaccurate. As it is being shown here, a full treatment of the latter will yield higher-order multi-particle interactions. Such quadratic interactions have been recently used phenomenologically [58] without a theoretical basis. Similar yet weaker interactions may be also drawn from relativistic corrections [68] as discussed here. In that sense, the quadratic interactions are shown to receive contributions from both non-relativistic and relativistic terms, which become quite significant when the mechanical frequency is comparable or larger than the electromagnetic frequency.

2. Classical Hamiltonian

The basic theory to be discussed is based on two important and basic assumptions. We first assume that the cavity mode decomposition is valid independent of the mirror motion. Second, the electric fields vanish at the mirror surface in the frame that the Lagrangian is constructed. These assumptions seem reasonable in the usual discussions with much lower mechanical frequencies when the cavity mode change can be treated adiabatically. However, when the end mirror oscillates at a very high frequency, it dresses the cavity modes so that the mode frequency might become undefined invalidating the assumption of mode decomposition. At higher frequencies, it has also been known and demonstrated that the mirror undergoing relativistic motion could produce photons from vacuum, known as the dynamical Casimir effect [69,70]. This is of course beyond the regime being explored in this article.
The focus of the first two Section 2.1 and Section 2.2 is to assert the claim, and demonstrate what term is missing and why it happens. As it will be shown and rigorously proven, even for the simplest case of interaction with a single-optical mode, a new term of the type q ˙ 2 Q 2 representing quadratic momentum q ˙ and optical field Q interactions is found, the origin of which is also identified. For the more general case of multi-mode optical fields, the situation is even much more complex and there are a few more missing terms to consider. Once the Lagrangian is known, the Hamiltonian is subsequently constructed in Section 2.3.

2.1. The Equations of Motion

The one-dimensional wave equation for transverse component of the magnetic potential A ( x , t ) in the dimensional form is expressed as [5]
c 2 A x x ( x , t ) = A t t ( x , t ) ,
where x and t are respectively the position and time coordinates, and c is the speed of light in free space. Suppose the Fourier series relations for the magnetic vector potential are defined as [5]
Q k ( t ) = 1 c 2 S μ 0 q ( t ) 0 q ( t ) A ( x , t ) sin [ κ k ( t ) x ] d x , A ( x , t ) = c 2 μ 0 S q ( t ) k = 1 Q k ( t ) sin [ κ k ( t ) x ] , ω k ( t ) = c π k / q ( t ) = c κ k ( t ) ,
where S is the cross-sectional area, μ 0 is the permeability of vacuum. This arrangement ensures that the definition of canonical variables can be used later, so that Q k ˙ 2 simply takes on the dimension of energy.
One may furthermore define the functions f k = 2 / q sin ( κ k x ) and g k = 2 / q cos ( κ k x ) , and hence A ( x , t ) = s Q k f k where s = c μ 0 / S . Here, the inner product is also defined as ( a | b ) = 0 q a b d x such that the following relations may be found
( f k | f j ) = δ k j ,   ( f k | κ j x | g j ) = α k j ,   ( f k | κ j 2 x 2 | g j ) = β k j .
After straightforward calculations one obtains
α k j = 1 2 δ k j + g k j , β k j = ( k 2 π 2 3 1 2 ) δ k j + h k j = { k 2 π 2 3 1 2 , k = j , 8 ( 1 ) k + j k j 3 ( k 2 j 2 ) 2 , k j .
Here, the anti-symmetric coefficients g k j and h k j are
g k j = g j k = { 0 , k = j , 2 ( 1 ) k + j k j j 2 k 2 , k j , h k j = { 0 , k = j , 8 ( 1 ) k + j k j 3 ( k 2 j 2 ) 2 , k j .
Differentiating f k and g k with respect to t , noting that κ ˙ k = q ˙ κ k / q , gives
f ˙ k = q ˙ q ( 1 2 f k + x κ k g k ) , g ˙ k = q ˙ q ( 1 2 g k x κ k f k ) , f ¨ k = q ¨ q q ˙ 2 q 2 ( 1 2 f k + x κ k g k ) + q ˙ 2 q 2 ( 1 4 f k + 2 x κ k g k x 2 κ k 2 f k ) .
It is possible now to differentiate A ( x , t ) with respect to position and time to obtain the relations
c 2 s A x x ( x , t ) = ω k 2 Q k f k , 1 s A t ( x , t ) = Q ˙ k f k q ˙ q Q k ( 1 2 f k + x κ k g k ) , 1 s A t t ( x , t ) = Q ¨ k f k 2 q ˙ q Q ˙ k ( 1 2 f k + x κ k g k ) q ¨ q q ˙ 2 q 2 Q k ( 1 2 f k + x κ k g k ) + q ˙ 2 q 2 Q k ( 1 4 f k + 2 x κ k g k x 2 κ k 2 f k ) .
Therefore, the wave equation reads
ω k 2 Q k f k = Q ¨ k f k q ˙ q Q ˙ k ( f k + 2 x κ k g k ) q ¨ q Q k ( 1 2 f k + x κ k g k ) + q ˙ 2 q 2 Q k ( 3 4 f k + 3 x κ k g k x 2 κ k 2 f k ) .
Now, the inner product relationships (3) and some further simplification (Appendix A) help us to obtain the equation of motion as
Q ¨ k = ω k 2 Q k + r k q ˙ 2 q 2 Q k + 2 q ˙ q g k j Q ˙ j + q ¨ q g k j Q j + q ˙ 2 q 2 ( h k j 3 g k j ) Q j ,
where all summations are nonzero only for j k , and
r k = k 2 π 2 3 + 1 4 .
The related equation in Law’s paper [5]
Q ¨ k = ω k 2 Q k + 2 q ˙ q g k j Q ˙ j + q ¨ q g k j Q j + q ˙ 2 q 2 g k j g l j Q l q ˙ 2 q 2 g k j Q j ,
is completely equivalent, but apparently different in presentation. Similarly, one may directly deduce from the Newton’s equation of motion
m q ¨ = x V ( x ) + S 2 μ 0 B 2 ( q , t ) = x V ( x ) + 1 q ( 1 ) k + j ω k ω j Q k Q j .

2.2. Lagrangian

The associated Lagrangian which leads to the above set of Euler equations is given by
= 1 2 m q ˙ 2 V ( q ) + 1 2 ( Q ˙ k 2 ω k 2 Q k 2 ) q ˙ q g k j Q j Q ˙ k + q ˙ 2 2 q 2 ( h k j 2 g k j + r k δ k j ) Q k Q j .
It should be again noticed that h k k = 0 by (5), which together with the antisymmetry of coefficients g k j from (5), allows further simplification to reach
= 1 2 m q ˙ 2 V ( q ) + 1 2 ( Q ˙ k 2 ω k 2 Q k 2 ) + q ˙ 2 2 q 2 d k j Q k Q j q ˙ q g k j Q j Q ˙ k ,
where d k j = 1 2 ( h k j + h j k ) + r k δ k j is related to the symmetric part of h k j , given by
d k j = d j k = { r k , k = j , 4 ( 1 ) k + j k j ( k 2 + j 2 ) ( k 2 j 2 ) 2 , k j .
This is to be compared with the quite different expression by Law [5] given by
= 1 2 m q ˙ 2 V ( q ) + 1 2 ( Q ˙ k 2 ω k 2 Q k 2 ) + q ˙ 2 2 q 2 g l k g l j Q k Q j q ˙ q g k j Q j Q ˙ k ,
which indeed consistently satisfies the Euler’s pair of equations.
Now, it is a matter of speculation whether (9) is equivalent to (11) or not. Firstly, the diagonal terms are equal if
k g k j g k j = r j = d j j ,
which holds true because of the identity
j = 1 , j k 4 k 2 j 2 ( j 2 k 2 ) 2 = k 2 π 2 3 + 1 4 .
Now, for a single-mode system with only one radiation mode, we may easily notice that (9) and (11) become readily identical. This becomes more clear by noticing that g k k = h k k = 0 , and hence for a system with only one electromagnetic radiation mode, (9) becomes Q ¨ = ω 2 Q + ( r q 2 q ˙ 2 ) Q , with r = r 1 . This is while Law’s expression (11) is Q ¨ = ω 2 Q + q 2 q ˙ 2 ( g 1 j g 1 j ) Q . However, (17) requires that r = g 1 j g 1 j , which confirms the equivalency for single-mode systems.
As for a multimode system, and by comparing (13) and (16) one would need
g k l g j l = d k j = r k δ k j 2 g k j + h k j ,
in order to (9) and (11) be identical. This can be put to numerical tests (here done by the author by coding a simple Mathematica program), and is in fact accurately satisfied. Hence, the single-mode Lagrangian in the non-relativistic limit can be written in the form
= 1 2 m q ˙ 2 V ( q ) + 1 2 [ Q ˙ 2 ω 2 ( q ) Q 2 ] + q ˙ 2 2 q 2 r Q 2 .
The last term has been usually ignored so far in the literature, and will result in momentum-field coupling. In the remainder of the paper, we focus on nonlinear terms arising from this interaction, and then also add up the relativistic corrections in the end.

2.3. Hamiltonian

The definition of canonical momenta taken here is
P k = Q ˙ k q ˙ q A k j Q j ,    p = m q ˙ 1 q B k j P k Q j ,
with A k j and B k j being some transformation coefficients to be determined later. One here may take advantage of the degree of freedom in choice of A k j and B k j to get rid of unwanted summation terms in the Hamiltonian. It has to be here noticed again that the existence of the last term of (20), being completely new even under the single-mode operation, has nothing to do with the choice of canonical momenta. That implies the final resulting single-mode (and therefore multi-mode) Hamiltonian will be inevitably different, incorporating a few new terms. The Hamiltonian may be now derived from the Lagrangian by iterated use of (21) and through the relationship = p q ˙ + P k Q ˙ k as
= [ m q ˙ 1 q B k j ( Q ˙ k q ˙ q A k l Q l ) Q j ] q ˙ + ( Q ˙ k q ˙ q A k j Q j ) Q ˙ k   .
Law [5] arbitrates the choice A k j = B k j = g k j . However, by going further with this choice for a single-mode optical field, it will be evident that P = Q ˙ and p = m q ˙ , hence, still resulting in an extra nonlinear term proportional to p 2 Q 2 / q 2 in the Hamiltonian, leading to a fourth-order momentum-field interaction. This will be discussed shortly in the subsection below.
Hence, the Lagrangian found in the above yields the Hamiltonian below after some algebra
= 1 2 m q ˙ 2 + V ( q ) + 1 2 ( Q ˙ k 2 + ω k 2 Q k 2 ) q ˙ q ( A k j + B k j g k j ) Q ˙ k Q j + q ˙ 2 2 q 2 ( 2 B k j A k l d j l ) Q l Q j .
Here is readily evident now by (19), that the last term can be made identically zero, only if d j l = g k j g k l = 2 B k j A k l . This not only cannot be satisfied by Law’s choice A k j = B k j = g k j , but also the second summation term nonlinear in q ˙ Q ˙ k will also survive, further complicating the Hamiltonian formulation.
Now, further elimination of q ˙ and Q ˙ k from (21) gives
= 1 2 m ( p + 1 q B k j P k Q j ) 2 + V ( q ) + 1 2 ω k 2 Q k 2 + 1 2 [ P k + 1 m q ( p + 1 q B l m P l Q m ) A k j Q j ] 2 1 m q ( p + 1 q B k j P k Q j ) C k j [ P k + 1 m q ( p + 1 q B k j P k Q j ) A k j Q j ] Q j + 1 2 m 2 q 2 ( p + 1 q B k j P k Q j ) 2 D j l Q l Q j .
Here, C k j = A k j + B k j g k j and D j l = 2 B k j A k l d j l . Dealing directly with such an intractable and long expression is without doubt too tough. Instead, we may tweak Law’s choice slightly as A k j = B k j = 1 2 g k j , which allows (23) be greatly simplified as
= 1 2 m q ˙ 2 + V ( q ) + 1 2 ( Q ˙ k 2 + ω k 2 Q k 2 ) q ˙ 2 4 q 2 d k j Q k Q j .
This can be further eventually expanded and simplified as
= 1 2 m ( p + 1 2 q g k j P k Q j ) 2 + V ( q ) + 1 2 ( P k 2 + ω k 2 Q k 2 ) + 1 4 m q ( p + 1 2 q g k j P k Q j ) g k j P k Q j 1 8 m 2 q 2 ( p + 1 2 q g k j P k Q j ) 2 d k j Q k Q j .
The last two terms of this Hamiltonian can be expanded to obtain multiple orders of interactions. These include higher-order tripartite phonon/two-photon, and quadpartite two-phonon/two-photon interactions, which does not exist in the Law’s Hamiltonian [5], given by
= 1 2 m ( p + 1 q g k j P k Q j ) 2 + V ( q ) + 1 2 ( P k 2 + ω k 2 Q k 2 ) .
As it appears, (27) is missing two very different types of momentum field interaction as the last two summation terms of (26). This fact becomes evident in below.

2.4. Single Optical Mode

The interesting difference between these two Hamiltonians becomes quite clear with consideration of only one optical mode in the cavity. This simplifies our derived Hamiltonian to
= 1 2 m p 2 + V ( q ) + 1 2 [ P 2 + π 2 q 2 Q 2 ] r 8 m 2 q 2 p 2 Q 2 ,
where ω ( q ) = π / q and r 3.8 , while the Law’s Hamiltonian [5] gives rise to the significantly different form
= 1 2 m p 2 + V ( q ) + 1 2 [ P 2 + π 2 q 2 Q 2 ] .
As it will be shown below, (28) and (29) agree only to the first order, and hence up to the standard optomechanical Hamiltonian.

3. Field Quantization

When the obtained Hamiltonian is moved to the realm of quantum mechanics, it is first needed to define the non-commutation rules [ q ^ , p ^ ] = i and [ Q ^ k , P ^ j ] = i δ k j , with the commutation rules [ q ^ , Q ^ k ] = [ Q ^ k , p ^ ] = [ q ^ , P ^ j ] = [ P ^ j , p ^ ] = 0 . This allows us to introduce the field creation and annihilation operators according to
Q ^ k = 2 ω k ( q ^ ) ( a ^ k + a ^ k ) = ω k ( q ^ ) k , P ^ k = i ω k ( q ^ ) 2 ( a ^ k a ^ k ) = ω k ( q ^ ) k ,
where ω k ( q ^ ) = c π k / q ^ is defined according to (2). Also for a mechanical resonant frequency Ω and a spring restoring potential
V ( q ^ ) = 1 2 m Ω 2 ( q ^ l ) 2 ,
the displacement operator may be defined as q ^ = l + x ^ , where l is the reference position of mirror, and hence the phonon ladder operators as
x ^ = 2 m Ω ( b ^ + b ^ ) = m Ω X ,    p ^ = i m Ω 2 ( b ^ b ^ ) = m Ω P ,
with [ b ^ , b ^ ] = 1 and [ X , P ] = i .
Now, it is necessary first to symmetrize [67] the classical Hamiltonian prior to insertion of operators, to ensure correct quantization of parameters. The process of symmetrization is done according to [71,72,73]
S { A B } = 1 2 ( A B + B A ) ,    S { A B } = 1 3 ( A S { B } + B S { A } + S { A B } ) ,
etc. Therefore, after symmetrization the final form of the Hamiltonian is given by
= 1 2 m p ^ 2 + V ( q ^ ) + 1 2 ( P ^ k 2 + ω k 2 Q ^ k 2 ) + 1 4 m S { 1 q ^ 2 ( p ^ q ^ + 1 2 g k j P ^ k Q ^ j ) g k j P ^ k Q ^ j } r 8 m 2 S { 1 q ^ 4 ( p ^ q ^ + 1 2 g k j P ^ k Q ^ j ) 2 d k j Q ^ k Q ^ j } .
For a single-optical mode, (34) greatly simplifies and one gets
= 1 2 m p ^ 2 + V ( q ^ ) + 1 2 ( P ^ 2 + π 2 q ^ 2 Q ^ 2 ) r 8 m 2 S { 1 q ^ 2 p ^ 2 } Q ^ 2 .
This has to be applied to the last interacting term, which involves
S { 1 q ^ 2 p ^ 2 } = S { p ^ 2 q ^ 2 } .
However, symmetrization of a term which contains n non-commuting terms, results in n! terms, which for this case sum up to a total of 4 ! = 24 different expressions. The direct way to get around this situation is to first make an estimate of which terms are the strongest in the limit of linearized interaction and ignore the rest. It is possible furthermore to use the approximate replacement
1 q ^ n 1 l n ( 1 n x ^ l ) ,
to obtain
= 1 2 m p ^ 2 + V ( q ^ ) + 1 2 [ P ^ 2 + ω 2 ( 1 2 x ^ l + 4 l 2 x ^ 2 + ) Q ^ 2 ] r 8 m 2 l 2 S { p ^ 2 ( 1 2 x ^ l ) } Q ^ 2 ,
where further substitutions should be taken as
P ^ ω ( 1 1 2 l x ^ + 3 8 l 2 x ^ 2 + ) ,    Q ^ ω ( 1 + 1 2 l x ^ 1 8 l 2 x ^ 2 + ) .
This can be decomposed to the terms
= 012 + 3 + 4 + 5 + ,    012 = 1 2 Ω P 2 + U ( X ) + 1 2 ω ( 2 + 2 ) .
Hence, there are several distinct types of nonlinear optomechanical multi-phonon/multi-photon interactions, which by defining R = r / 4 0.95 are respectively given by
3 = ω 2 l m Ω X ( 2 + 2 ) ,    4 = 2 2 l 2 m [ R Ω ω P 2 2 + ω Ω X 2 ( 2 + 2 ) ] , 5 = 5 2 2 m 3 2 l 3 Ω { 2 R Ω ω S { P 2 X } 2 + ω Ω X 3 ( 2 + 2 ) } ,
and so on for higher order interactions. Here, the expansion of symmetrized terms, for instance, gives
S { P 2 X } = 1 3 ( P 2 X + X P 2 + P X P ) .
Now, it is noted that since usually ω Ω , it may observed that the first terms are much weaker than the second terms. Hence, using the identity 2 + 2 = 1 2 n ^ + 1 4 where n ^ is photon number operator, the following is obtained
3 = ω l m Ω X ( n ^ + 1 2 ) = α X ( n ^ + 1 2 ) , 4 = 2 l 2 m ω Ω X 2 ( n ^ + 1 2 ) = + β X 2 ( n ^ + 1 2 ) , 5 = ω Ω 5 2 m 3 2 l 3 Ω X 3 ( n ^ + 1 2 ) = γ X 3 ( n ^ + 1 2 ) ,
where 3 OM is the simple optomechanical interaction, and 4 is known as the quadratic interaction. It has to be emphasized that while 3 OM is actually nonlinear in the exact mathematical sense, it is the quadratic interaction 4 which is mostly referred to as the nonlinear interaction in the literature [55,56,57]. Since it is possible to make g 0 and therefore 3 identically vanish by appropriate optomechanical design [55,56,57,58] in which the overlap integral of optical and mechanical modes sums up to zero, hence the quadratic interactions 4 can then find physical significance.
The quadratic interaction has been a subject of growing importance in the recent years in optomechanical systems [59,60,61] and beyond [62]. In [59] the photon statistics and blockade under 4 interactions has been studied and analytical expressions were derived. The quantum dissipative master function has been numerically solved and the corresponding correlation functions were obtained. Interestingly, quadratic optomechanical interactions can arise at the single-photon level, too, where rigorous analytical solutions have been devised [60]. Such type of interactions can be also well described using equivalent nonlinear electrical circuits, where a Josephson junction brings in the desired nonlinearity of quadratic interactions and terminate a pair of lumped transmission lines [61]. Finally, ultracold atoms also can exhibit interactions of a comparable type which is mathematically equivalent to the quadratic interaction [62].
The single-photon multi-particle rates are given by
α = ω l m Ω 2 g 0 ,    β = l 2 m ω Ω ,    γ = ω Ω 3 2 m 3 2 l 3 Ω .
This summarizes the Hamiltonian as
= 0 + i n t .
in which 0 and i n t are respectively the non-interacting and interacting Hamiltonians
0 = 012 1 2 ( α X β X 2 + γ X 3 + ) ,    i n t = ( α X β X 2 + γ X 3 + ) n ^ ,
when ω Ω . Now, the dimensionless constant θ is defined as
θ = 1 l m Ω = x z p l ,
with x z p being the r.m.s. value of zero-point fluctuations, by which the following is deduced
β = θ α α ,    γ = θ β = θ 2 α β .
This implies that every kind of higher-order interaction is typically θ times weaker than the interaction of the preceding-order. It should be noted that while such interactions are normally expected to rapidly vanish with the order increasing, is a well-known fact that certain physical phenomena such as magnetism in solid 3He cannot be understood without inclusion of four-particle interaction terms [74,75]. It is worth here to mention that a detailed theory of optomechanics in superfluid. It has been developed [76], but no expression for the nonlinear terms has been reported.
In general, the interaction of mechanical and optical modes is not strictly one-dimensional, implying that the overlap integral of normalized modes should also be taken into account. For instance, odd mechanical modes with even optical modes have zero interaction. In that sense, tuning the interaction to an odd mode and then shining an even optical mode, or vice versa, makes the optomechanical interaction identically zero by setting α 2 g 0 = 0 . Then the lowest order surviving interaction would be the 4 term. This method has been used in [55,56,57] to highlight the quadratic interaction and make its measurement much easier. It has been shown that these quadratic terms may be exploited for direct observation of mechanical eigenmode jumps [55,56], as well as two-phonon cooling and squeezing [57], while the coupling strength β could be increased by three orders of magnitude [56].
Moreover, the origin of mechanical parametric coupling which has recently been phenomenologically hypothesized [58] for the associated physical interactions cannot be understood without the presented analysis, although based on some earlier experimental evidence [77].
It must be added that the condition ω Ω may be violated in carefully designed superconducting microwave circuits and also the recently demonstrated molecular optomechanics [78], which signifies the importance of the P 2 2 term in 4 . It is furthermore worthwhile to point out that the regime ω = Ω can be indeed be accessed and investigated, as it has been shown experimentally for superconducting circuit optomechanics [79]. The proposal of light propagation in a cylinder with rotating walls [80] also requires accessing regimes where ω and Ω fall within the same order of magnitude. Alternatively, in situations where ω Ω , the scaling will be then given as
θ = R Ω 2 x z p ω 2 l ,
which shows a significant enhancement in this type of interactions.

3.1. Conditions for Observation of Momentum-Field Quadratic Interactions

In summary, two general criteria should be satisfied in a carefully designed experiment to allow investigation of momentum-field quadratic interactions:
  • The optomechanical interaction 3 must vanish to allow easier study of quadratic interaction 4 . This is quite possible by design as extensively has been discussed in the above and literature [55,56,57,58,59,60,61,62].
  • The mechanical frequency Ω must be of the same order of magnitude or exceeding the electromagnetic frequency ω . This is also possible and at least one experiment using superconducting optomechanics [79] has accessed this regime. Other possibilities are molecular optomechanics [78] as well as a rotating cylinder [80].
Evidently, such momentum-field quadratic interactions might be more difficult to observe under normal experimental conditions compared to the regular optomechanical setups. However, progressive developments in the precision and accuracy of optomechanics experiments, such as what happened for the case of Laser Interferometric Gravitational Observatory (LIGO) [81], could make it eventually possible to realize and probe such unexplored domains.

3.2. Linearized Quantization

3.2.1. Optical Field

The standard method to linearize the interaction Hamiltonian can be now used by making the substitutions a ^ a ¯ + a ^ where the new a ^ operator from now on stands for the non-classical perturbations and a ^ is a measure of optical field amplitude. Then ignoring higher-order terms and retaining only the lowest-order interacting terms, we get
3 = g 3 ( b ^ + b ^ ) ( e i φ a ^ + e i φ a ^ ) ,
as well as
4 = + g 4 + ( b ^ + b ^ ) 2 ( e i φ a ^ + e i φ a ^ ) , ω Ω ,   4 = + g 4 ( b ^ b ^ ) 2 ( a ^ + a ^ ) , ω Ω .
Here, φ = a ¯ and the coupling frequencies are defined as
g 3 = α 2 | a ¯ | g 0 | a ¯ | G ,    g 4 + = β 2 | a ¯ | = θ G ,    g 4 =   R Ω 2 ω 2 g 4 + .

3.2.2. Mechanical Field

Following the same method to linearize the mechanical motions, with the replacement b ^ b ¯ + b ^ where the new b ^ operator denotes the perturbations, gives rise to the expressions
4 = + G 4 + ( b ^ + b ^ ) ( e i φ a ^ + e i φ a ^ ) , ω Ω , 4 = + G 4 ( b ^ b ^ ) ( a ^ + a ^ ) , ω Ω ,
where ϑ = b ¯ is set to zero without loss of generality, G 4 + = 2 | b ¯ | g 4 + cos ϑ , and G 4 = 2 | b ¯ | g 4 sin ϑ . In general, when ω Ω is violated, one would expect the momentum of mirror be coupled to the first quadrature of the radiation field. This type of interaction can be compared to the normal optomechanical interaction (50), in which the position is coupled to the first quadrature of the field.

3.2.3. Squeezing Hamiltonian

When the optical and mechanical frequencies do not differ by orders of magnitude so that neither ω Ω nor ω Ω hold, then the linearized Hamiltonian could be recast as
4 = G 4 + ( b ^ + b ^ ) ( e i φ a ^ + e i φ a ^ ) + G 4 ( b ^ b ^ ) ( a ^ + a ^ ) ,
This can be written as
4 = G ( a ^ B ^ + a ^ B ^ ) = G ( b ^ A ^ + b ^ A ^ ) ,
where
G 4 = G 4 + G 4 , A ^ = a ^ sin h ρ + a ^ cos h ρ , B ^ = b ^ cos h ρ + b ^ sin h ρ , ρ = tan h 1 ( G 4 + G 4 e i φ G 4 + + G 4 e i φ ) .
It is here to be noticed that B ^ and A ^ are in the standard form of Bogoliubov squeezing operator [54,55]. It may be noted that the equation
G 4 = R Ω 2 ω 2 G 4 + ,
is actually a function of ϑ by definition of G 4 + and G 4 . Simplifying the above gives the expression for squeeze ratio as
ρ = ln ( ω R Ω ) i φ 2 .
This shows that quadratic interactions give rise to squeezed mechanical or optical states unless ω = R Ω and of course φ = 0 .

3.2.4. Special Case

As discussed above, the Hamiltonian 3 can be made identically zero [56,57,82,83] to access the quadratic interaction terms 4 directly. There is an interesting condition on the ratio of optical to mechanical frequencies, which could be sought here. Let
ω = η R Ω ,
in which η is a constant to be determined later. This allows the 4 to be written as
4 = 2 β [ 1 η P 2 2 + X 2 ( 2 + 2 ) ]                      = β [ 1 2 η ( b ^ b ^ ) 2 ( a ^ + a ^ ) 2 + ( b ^ + b ^ ) 2 ( a ^ a ^ + a ^ a ^ ) ] .
Further expansion of results in (Appendix B)
4 , i n t = 2 β | a ¯ | [ ( 1 + 1 2 η ) ( b ^ 2 + b ^ 2 ) + ( 1 1 2 η ) m ^ ] × ( a ^ + a ^ ) .
Then, for the choice of η = 1 2 , that is ω 0.69 Ω , one may reach the desired interaction quadratic Hamiltonian, linearized in the electromagnetic operators
4 , i n t = 2 J ( b ^ 2 + b ^ 2 ) ( a ^ + a ^ ) ,
where the interaction rate is J = λ = 2 β | a ¯ | . When expanded in its four terms and after the replacement c ^ = 1 2 b ^ 2 (Appendix C), one may immediately recognize the Hamiltonian of the type
4 , i n t = J ( c ^ a ^ + c ^ a ^ ) + λ ( c ^ a ^ + c ^ a ^ ) .
The first parenthesis represents the Hopping or Beam-Splitter term, while the second is normally referred to as the dissipation. Interestingly, the above could have been further linearized in mechanical operators to obtain
4 , i n t = i J ( e i ϑ b ^ + e i ϑ b ^ ) ( a ^ a ^ ) = i J ( e i ϑ b ^ a ^ e i ϑ b ^ a ^ ) + i J ( e i ϑ b ^ a ^ e i ϑ b ^ a ^ ) .
where J = 2 J | b ¯ | . This latter form, may find application in non-reciprocal optomechanics [84].

4. Relativistic Considerations

As a final remark, the approximate nature of the Lagranian formulation by Law [5] has not been left unnoticed. It could be attributed first to the non-relativistic description of mirror’s motion which ultimately ignores higher-order interactions, and then to the relativistic nature of radiation friction force and the associated Doppler shift [66]. As a result, in a subsequent paper by Cheung and Law [67], it has been made clear that the non-relativistic optomechanical Hamiltonian is correct only to the first-order in q ˙ .
The nature of the relativistic corrections can be quite different, as follows:
  • relativistic Doppler shift [66], which causes corrections in q ˙ / c ,
  • relativistic correction in radiation pressure term [67], the lowest-order of which being proportional to q ˙ / c ,
  • length contraction [68], due to the moving mirror boundary, resulting again in corrections as q ˙ / c .
Not surprisingly, all these relativistic terms vanish in the limit of infinite light speed c . These altogether could be taken into account in a fully relativistic formulation of the Lagrangian and equations of motions for the mirror and optical field [85], which has been recently carried out in an extensive research by Castaños and Weder [68].
As shown in Appendix D, the total relativistic correction terms added to the Hamiltonian takes the form
Δ = ( b ^ b ^ ) 2 k w k j ( a ^ k + a ^ k ) ( a ^ j + a ^ j ) .
For the single-mode cavity, w = χ 0 π d Ω / 4 m c l 2 , to be compared with β = ω / m Ω l 2 in (48). Hence, the relativistic correction to the quadratic Hamiltonian 4 is
w β = χ 0 π d Ω 2 4 c ω .
Again, it is seen that when ω Ω is violated, the relativistic corrections might be quite significant. In any case, there is no relativistic correction to 3 .

5. Conclusions and Future Work

The derivation of the optomechanical Hamiltonian has been carefully examined from the modal expansions, equations of motion, all the way to the Lagrangian, and ultimately the Hamiltonian and relativistic considerations. A set of correction terms added to the nonlinear terms has been identified, which do not eliminate under any choice of canonical momenta. With the careful system design which allows g 0 = 0 , these type of interactions are particularly interesting and now being actively pursued. It was shown that under these conditions one may expect coupling of mechanical momentum to the field position. Other sorts of interactions emerge under various conditions. In general, when the optical frequency is not much larger than the mechanical frequency, novel nonlinear interactions may appear.
A future work of the author [86] discusses a semi-analytical method based on the Langevin equations, which will enable easier study of quadratic interactions in quantum mechanical systems. This modified method of Langevin equations could in principle greatly simplify the study of quadratic and higher-order interactions, which would otherwise need either a full numerical solution using the master equation approach or full expansion unto the infinite set of orthogonal basis number kets.

Acknowledgments

This work been supported in part by the Laboratory of Photonics and Quantum Measurements (LPQM) at École Polytechnique Fédérale de Lausanne (EPFL). The author wishes to thank Christophe Galland at EPFL for reading the manuscript and useful suggestions, as well as Alexey Feofanov, Itay Shomroni, Amir Hossein Ghadimi, and Liu Qiu at LPQM for discussions. The author would like to thank the referees for their invaluable and constructive comments. The author is indebted to Kirstin Friedrich for overwhelming encouragement and enthusiasm regarding this work.

Conflicts of Interest

The author declares no conflict of interest.

Appendix A. Equations of Motion

In this section, we present the step-by-step details of the derivation of (9) from the previous equations, as it constitutes the most critical part of this article.
Starting from (8), one has first to rename the dummy index from k to j , multiply both sides by f k , and then take the inner product. This will yield the expression
ω j 2 Q j ( f k | f j ) = Q ¨ j ( f k | f j ) q ˙ q Q ˙ j [ ( f k | f j ) + 2 ( f k | κ j x | g j ) ] q ¨ q Q j [ 1 2 ( f k | f j ) + ( f k | κ j x | g j ) ] + q ˙ 2 q 2 Q j [ 3 4 ( f k | f j ) + 3 ( f k | κ j x | g j ) ( f k | κ j 2 x 2 | g j ) ] .
Using (3), we trivially get
ω j 2 Q j δ k j = Q ¨ j δ k j q ˙ q Q ˙ j [ δ k j + 2 α k j ] q ¨ q Q j [ 1 2 δ k j + α k j ] + q ˙ 2 q 2 Q j [ 3 4 δ k j + 3 α k j β k j ] ,
which after rearrangement takes the form
Q ¨ k = ω k 2 Q k + q ˙ q Q ˙ k + q ¨ 2 q Q k 3 q ˙ 2 4 q 2 Q k + 2 q ˙ q α k j Q ˙ j + q ¨ q α k j Q j q ˙ 2 q 2 ( 3 α k j β k j ) Q j .
We may furthermore use α k j = 1 2 δ k j + g k j and β k j = ( 1 3 k 2 π 2 1 2 ) δ k j + h k j from (4) to simplify and rewrite (A3) as
Q ¨ k = ω k 2 Q k 3 q ˙ 2 4 q 2 Q k + 3 q ˙ 2 2 q 2 Q k + q ˙ 2 q 2 ( k 2 π 2 3 1 2 ) Q k + 2 q ˙ q g k j Q ˙ j + q ¨ q g k j Q j q ˙ 2 q 2 ( 3 g k j h k j ) Q j ,
which by plugging in the definition for r k from (10) takes the form
Q ¨ k = ω k 2 Q k + r k q ˙ 2 q 2 Q k + 2 q ˙ q g k j Q ˙ j + q ¨ q g k j Q j + q ˙ 2 q 2 ( h k j 3 g k j ) Q j .
This is exactly Equation (9).

Appendix B. Special Case

Expansion of (60) results in
4 =   β [ 1 2 η ( b ^ b ^ ) 2 ( a ^ 2 + a ^ 2 ) + ( 1 + 1 2 η ) ( m ^ + 1 2 ) ( n ^ + 1 2 ) + ( b ^ 2 + b ^ 2 ) ( n ^ + 1 2 ) ] =   β [ 1 2 η ( b ^ 2 + b ^ 2 ) ( a ^ 2 + a ^ 2 ) + ( 1 + 1 2 η ) ( m ^ + 1 2 ) ( n ^ + 1 2 ) + ( b ^ 2 + b ^ 2 ) ( n ^ + 1 2 ) 1 2 η ( m ^ + 1 2 ) ( a ^ 2 + a ^ 2 ) ] ,
where n ^ = a ^ a ^ and m ^ = b ^ b ^ are respectively photon and photon number operators. Retaining only the interacting terms, gives the expression
4 , i n t =   β [ 1 2 η ( b ^ 2 + b ^ 2 ) ( a ^ 2 + a ^ 2 ) + ( 1 + 1 2 η ) m ^ n ^ + ( b ^ 2 + b ^ 2 ) n ^ 1 η m ^ ( a ^ 2 + a ^ 2 ) ] .
In the limit ω Ω with η , 4 as in (53) is recovered. With linearization of the electromagnetic field operators, and removing the non-interacting terms the following is found
4 , i n t = 2 β [ 1 2 η ( b ^ 2 + b ^ 2 ) ( a ¯ * a ^ + a ¯ a ^ ) + ( 1 + 1 2 η ) m ^ ( a ¯ a ^ + a ¯ * a ^ ) + ( b ^ 2 + b ^ 2 ) ( a ¯ a ^ + a ¯ * a ^ ) 1 η m ^ ( a ¯ * a ^ + a ¯ a ^ ) ] ,
By continuing the work on the linearized quadratic interaction one obtains the expression
4 , i n t = 2 β | a ¯ | [ 1 2 η ( b ^ 2 + b ^ 2 ) ( e i φ a ^ + e i φ a ^ ) + ( 1 + 1 2 η ) m ^ ( e i φ a ^ + e i φ a ^ ) + ( b ^ 2 + b ^ 2 ) ( e i φ a ^ + e i φ a ^ ) 1 η m ^ ( e i φ a ^ + e i φ a ^ ) ] ,
which for φ = 0 simplifies to
4 , i n t = 2 β | a ¯ | [ ( 1 + 1 2 η ) ( b ^ 2 + b ^ 2 ) + ( 1 1 2 η ) m ^ ] × ( a ^ + a ^ ) .

Appendix C. Squared Annihilator

The operator c ^ has clearly a simple solution for its eigenkets, which is the same as coherent states such as | z where c ^ | z = 1 2 z 2 | z . Hence, the eigenvalue is simply the complex number 1 2 z 2 . Meanwhile, one has
| z = e 1 2 | z | 2 m = 0 z m m ! | m .
It is furthermore easy to check that [ c ^ , c ^ ] = m ^ + 1 2 = b ^ b ^ + 1 2 [86]. When, the mean phonon number is m ^ = 1 2 , then [ c ^ , c ^ ] = 1 , which is quite similar to the commutator [ b ^ , b ^ ] = 1.

Appendix D. Relativistic Correction

The relativistic Lagrangian density for a light field with normal incidence to a fully reflective and non-compressible moving mirror, correct to the first-order in q ˙ and q ¨ , reads [68]
V = 1 2 ( E · D H · B ) + Γ 2 ϵ 0 2 χ | E · z ^ c Β B · y ^ | 2 ,
where E = z ^ t A , B = × ( A z ^ ) = y ^ x A , D = ϵ 0 E , and B = μ 0 H . Furthermore,
Β = v c = q ˙ c = p m c ,   Γ = 1 1 Β 2 ,
and χ is a dimensionless shape function independent of v , being zero outside mirror and relative susceptibility of the mirror’s dielectric χ 0 inside, and ϵ 0 is the permittivity of vacuum. By expanding in the powers of Β , this Lagrangian gives the first- and second-order corrections to the quadratic Hamiltonian density as
V Δ = V Δ ( 1 ) + V Δ ( 2 ) = Β V | Β = 0 Β 1 2 2 Β 2 V | Β = 0 Β 2 .
Hence, one may obtain
Δ ( 1 ) = ϵ 0 S 0 q Β [ 1 2 ( 1 1 Β 2 ) 2 χ | A t + c Β A x | 2 ] | Β = 0 Β d x ,
which further simplifies as
Δ ( 1 ) = ϵ 0 S c 0 q χ A t A x Β d x .
It is appropriate to assume the approximation of conducting interface [87,88,89] for the mirror, such as the thickness is let to approach zero, while it susceptibility increases proportionally. In that limit, one may set
χ ( x , t ) χ 0 d δ [ x q ( t ) ] ,
where d is the mirror’s thickness. This is similar to the assumption of the locality of interaction by Gardiner & Zoller [90], too. Hence, one gets
Δ ( 1 ) ϵ 0 V c χ 0 A t ( q , t ) A x ( q , t ) Β ,
with V = S d is the cavity volume. Now, one has from (2), A x ( q , t ) = s π 2 / q 3 k Q k , A t ( q , t ) = q ˙ A x ( q , t ) , and thus
Δ ( 1 ) 2 π 2 d χ 0 q ˙ 2 q 3 k j Q k Q j
It is quite remarkable that (A19) is purely relativistic, and vanish in the limit of infinite c , as shown below. Here, the dependence on t is hidden for convenience. This term translates after symmetrization into
Δ ( 1 ) = 2 ( b ^ b ^ ) 2 k w k j ( a ^ k + a ^ k ) ( a ^ j + a ^ j ) ,
where w k j = j k χ 0 π d Ω / 4 m c l 2 are the coupling rates. Now, the quadratic correction Δ ( 2 ) is given by
Δ ( 2 ) = 1 2 ϵ 0 S 0 q 2 Β 2 { 1 2 ( 1 1 Β 2 ) 2 χ | A t + c Β A x | 2 } d x | Β = 0 Β 2 .
Simplifying and using the conducting interface approximation gives
Δ ( 2 ) = 1 2 S ϵ 0 0 q χ [ A t 2 + c 2 A x 2 ] Β 2 d x 1 2 V ϵ 0 χ 0 [ q ˙ 2 + c 2 ] A x 2 ( q , t ) Β 2 π 2 d χ 0 q ˙ 2 q 3 k j Q k Q j .
This one after insertion of operators gives Δ ( 2 ) = 1 2 Δ ( 1 ) and thus the total relativistic correction is found as
Δ ( 1 ) + Δ ( 2 ) = ( b ^ b ^ ) 2 k w k j ( a ^ k + a ^ k ) ( a ^ j + a ^ j ) .

References

  1. Kippenberg, T.J.; Vahala, K.J. Cavity optomechanics. Science 2008, 321, 1172. [Google Scholar] [CrossRef] [PubMed]
  2. Aspelmeyer, M.; Kippenberg, T.; Marquardt, F. Cavity Optomechanics; Springer: Berlin, Germany, 2014. [Google Scholar]
  3. Marquardt, F.; Girvin, S.M. Optomechanics Retarded radiation forces. Physics 2009, 2, 40. [Google Scholar] [CrossRef]
  4. Meystre, P. A short walk through quantum optomechanics. Ann. Phys. 2013, 525, 215. [Google Scholar] [CrossRef]
  5. Law, C.K. Interaction between a moving mirror and radiation pressure: A Hamiltonian formulation. Phys. Rev. A 1995, 51, 2537. [Google Scholar] [CrossRef] [PubMed]
  6. Gong, Z.R.; Ian, H.; Liu, X.Y.; Sun, C.P.; Nori, F. Effective Hamiltonian approach to the Kerr nonlinearity in an optomechanical system. Phys. Rev. A 2009, 80, 065801. [Google Scholar] [CrossRef]
  7. Lü, X.Y.; Zhang, W.M.; Ashhab, S.; Wu, Y.; Nori, F. Quantum-criticality-induced strong Kerr nonlinearities in optomechanical systems. Sci. Rep. 2013, 3, 2943. [Google Scholar] [CrossRef] [PubMed]
  8. Mahboob, I.; Mounaix, M.; Nishiguchi, K.; Fujiwara, A.; Yamaguchi, H. A multimode electromechanical parametric resonator array. Sci. Rep. 2014, 4, 4448. [Google Scholar] [CrossRef] [PubMed]
  9. Heinrich, G.; Ludwig, M.; Qian, J.; Kubala, B.; Marquardt, F. Collective dynamics in optomechanical arrays. Phys. Rev. Lett. 2011, 107, 043603. [Google Scholar] [CrossRef] [PubMed]
  10. Gan, J.H.; Xiong, H.; Si, L.G.; Lu, X.Y.; Wu, Y. Solitons in optomechanical arrays. Opt. Lett. 2016, 41, 2676. [Google Scholar] [CrossRef] [PubMed]
  11. Chen, W.; Clerk, A.A. Photon propagation in a one-dimensional optomechanical lattice. Phys. Rev. A 2014, 89, 033854. [Google Scholar] [CrossRef]
  12. Xuereb, A.; Genes, C.; Pupillo, G.; Paternostro, M.; Dantan, A. Reconfigurable long-range phonon dynamics in optomechanical arrays. Phys. Rev. Lett. 2014, 112, 133604. [Google Scholar] [CrossRef] [PubMed]
  13. Houhou, O.; Aissaoui, H.; Ferraro, A. Generation of cluster states in optomechanical quantum systems. Phys. Rev. A 2015, 92, 063843. [Google Scholar] [CrossRef]
  14. Peano, V.; Brendel, C.; Schmidt, M.; Marquardt, F. Topological phases of sound and light. Phys. Rev. X 2015, 5, 031011. [Google Scholar] [CrossRef]
  15. Ludwig, M.; Marquardt, F. Quantum many-body dynamics in optomechanical arrays. Phys. Rev. Lett. 2013, 111, 073603. [Google Scholar] [CrossRef] [PubMed]
  16. Liao, J.-Q.; Wu, Q.-Q.; Nori, F. Entangling two macroscopic mechanical mirrors in a two-cavity optomechanical system. Phys. Rev. A 2014, 89, 014302. [Google Scholar] [CrossRef]
  17. Johansson, J.R.; Johansson, G.; Nori, F. Optomechanical-like coupling between superconducting resonators. Phys. Rev. A 2014, 90, 053833. [Google Scholar] [CrossRef]
  18. Kronwald, A.; Marquardt, F.; Clerk, A.A. Arbitrarily large steady-state bosonic squeezing via dissipation. Phys. Rev. A 2013, 88, 063833. [Google Scholar] [CrossRef]
  19. Lü, X.Y.; Liao, J.Q.; Tian, L.; Nori, F. Steady-state mechanical squeezing in an optomechanical system via Duffing nonlinearity. Phys. Rev. A 2015, 91, 013834. [Google Scholar] [CrossRef]
  20. Lü, X.Y.; Wu, Y.; Johansson, J.R.; Jing, H.; Zhang, J.; Nori, F. Squeezed Optomechanics with Phase-matched Amplification and Dissipation. Phys. Rev. Lett. 2015, 114, 093602. [Google Scholar] [CrossRef] [PubMed]
  21. Schilling, R.; Schütz, H.; Ghadimi, A.H.; Sudhir, V.; Wilson, D.J.; Kippenberg, T.J. Field Integration of a SiN Nanobeam and a SiO2 Microcavity for Heisenberg-Limited Displacement Sensing. Phys. Rev. Appl. 2016, 5, 054019. [Google Scholar] [CrossRef]
  22. Ruesink, F.; Miri, M.A.; Alu, A.; Verhagen, E. Nonreciprocity and magnetic-free isolation based on optomechanical interactions. Nat. Commun. 2016, 7, 13662. [Google Scholar] [CrossRef] [PubMed]
  23. Xu, X.-W.; Li, Y.; Chen, A.-X.; Liu, Y.-X. Nonreciprocal conversion between microwave and optical photons in electro-optomechanical systems. Phys. Rev. A 2016, 93, 023827. [Google Scholar] [CrossRef]
  24. Fang, K.; Luo, J.; Metelmann, A.; Matheny, M.H.; Marquardt, F.; Clerk, A.A.; Painter, O. Generalized nonreciprocity in an optomechanical circuit via synthetic magnetism and reservoir engineering. Nat. Phys. 2017, 13, 465. [Google Scholar] [CrossRef]
  25. Xu, H.; Mason, D.; Jiang, L.; Harris, J.G.E. Topological dynamics in an optomechanical system with highly non-degenerate modes. arxiv, 2017; 1703.07374. [Google Scholar]
  26. Bernier, N.R.; Tóth, L.D.; Koottandavida, A.; Ioannou, M.; Malz, D.; Nunnenkamp, A.; Feofanov, A.K.; Kippenberg, T.J. Nonreciprocal reconfigurable microwave optomechanical circuit. arxiv, 2017; 1612.08223. [Google Scholar]
  27. Peterson, G.A.; Lecocq, F.; Cicak, K.; Simmonds, R.W.; Aumentado, J.; Teufel, J.D. Demonstration of efficient nonreciprocity in a microwave optomechanical circuit. arxiv, 2017; 1703.05269. [Google Scholar]
  28. Malz, D.; Tóth, L.D.; Bernier, N.R.; Feofanov, A.K.; Kippenberg, T.J.; Nunnenkamp, A. Quantum-limited directional amplifiers with optomechanics. arxiv, 2017; 1705.00436. [Google Scholar]
  29. Zippilli, S.; Li, J.; Vitali, D. Steady-state nested entanglement structures in harmonic chains with single-site squeezing manipulation. Phys. Rev. A 2015, 92, 032319. [Google Scholar] [CrossRef]
  30. Barzanjeh, S.; Guha, S.; Weedbrook, C.; Vitali, D.; Shapiro, J.H.; Pirandola, S. Microwave quantum illumination. Phys. Rev. Lett. 2015, 114, 080503. [Google Scholar] [CrossRef] [PubMed]
  31. Zhang, K.; Bariani, F.; Dong, Y.; Zhang, W.; Meystre, P. Proposal for an optomechanical microwave sensor at the subphoton level. Phys. Rev. Lett. 2015, 114, 113601. [Google Scholar] [CrossRef] [PubMed]
  32. Toth, L.D.; Bernier, N.R.; Nunnenkamp, A.; Glushkov, E.; Feofanov, A.K.; Kippenberg, T.J. A dissipative quantum reservoir for microwave light using a mechanical oscillator. arxiv, 2017; 1602.05180. [Google Scholar]
  33. Galland, C.; Sangouard, N.; Piro, N.; Gisin, N.; Kippenberg, T.J. Heralded single-phonon preparation, storage, and readout in cavity optomechanics. Phys. Rev. Lett. 2014, 112, 143602. [Google Scholar] [CrossRef] [PubMed]
  34. Khorasani, S. Coupled mode theory of optomechanical crystals. IEEE J. Quant. Electron. 2016, 52, 6100406. [Google Scholar] [CrossRef]
  35. Aram, M.H.; Khorasani, S. Efficient analysis of confined guided modes in phoxonic crystal slabs. arxiv, 2017; 1703.08416. [Google Scholar]
  36. Aram, M.H.; Khorasani, S. Optical wave evolution due to interaction with elastic wave in a phoxonic crystal slab waveguide. arxiv, 2017; 1704.00895. [Google Scholar]
  37. Bochmann, J.; Vainsencher, A.; Awschalom, D.D.; Cleland, A.N. Nanomechanical coupling between microwave and optical photons. Nat. Phys. 2013, 9, 712. [Google Scholar] [CrossRef]
  38. Andrews, R.W.; Peterson, R.W.; Purdy, T.P.; Cicak, K.; Simmonds, R.W.; Regal, C.A.; Lehnert, K.W. Bidirectional and efficient conversion between microwave and optical light. Nat. Phys. 2014, 10, 321. [Google Scholar] [CrossRef]
  39. Javerzac-Galy, C.; Plekhanov, K.; Bernier, N.R.; Toth, L.D.; Feofanov, A.K.; Kippenberg, T.J. On-chip microwave-to-optical quantum coherent converter based on a superconducting resonator coupled to an electro-optic microresonator. Phys. Rev. A 2016, 94, 053815. [Google Scholar] [CrossRef]
  40. Tian, L. Optoelectromechanical transducer: reversible conversion between microwave and optical photons. Ann. Phys. (Berlin) 2015, 527, 1. [Google Scholar] [CrossRef]
  41. Caprara Vivoli, V.; Barnea, T.; Galland, C.; Sangouard, N. Proposal for an optomechanical Bell test. Phys. Rev. Lett. 2016, 116, 070405. [Google Scholar] [CrossRef] [PubMed]
  42. Tsang, M. Cavity quantum electro-optics. Phys. Rev. A 2010, 81, 063837. [Google Scholar] [CrossRef]
  43. Tsang, M. Cavity quantum electro-optics. II. Input-output relations between traveling optical and microwave fields. Phys. Rev. A 2011, 84, 043845. [Google Scholar] [CrossRef]
  44. Chang, Y.; Shi, T.; Liu, Y.X.; Sun, C.P.; Nori, F. Multistability of electromagnetically induced transparency in atom-assisted optomechanical cavities. Phys. Rev. A 2011, 83, 063826. [Google Scholar] [CrossRef]
  45. Jing, H.; Özdemir, Ş.K.; Geng, Z.; Zhang, J.; Lü, X.Y.; Peng, B.; Yang, L.; Nori, F. Optomechanically-Induced Transparency in parity-time-symmetric microresonators. Sci. Rep. 2015, 5, 9663. [Google Scholar] [CrossRef] [PubMed]
  46. Wang, H.; Gu, X.; Liu, Y.X.; Miranowicz, A.; Nori, F. Optomechanical analog of two-color electromagnetically-induced transparency: Photon transmission through an optomechanical device with a two-level system. Phys. Rev. A 2014, 90, 023817. [Google Scholar] [CrossRef]
  47. Liu, Y.L.; Wu, R.; Zhang, J.; Özdemir, Ş.K.; Yang, L.; Nori, F.; Liu, Y.X. Controllable optical response by modifying the gain and loss of a mechanical resonator and cavity mode in an optomechanical system. Phys. Rev. A 2017, 95, 013843. [Google Scholar] [CrossRef]
  48. Huang, S. Quantum state transfer in cavity electro-optic modulators. Phys. Rev. A 2015, 92, 043845. [Google Scholar] [CrossRef]
  49. Larson, J.; Horsdal, M. Photonic Josephson effect, phase transitions, and chaos in optomechanical systems. Phys. Rev. A 2011, 84, 021804. [Google Scholar] [CrossRef]
  50. Metcalfe, M. Applications of cavity optomechanics. Appl. Phys. Rev. 2014, 1, 031105. [Google Scholar] [CrossRef]
  51. Alidoosty, M.; Khorasani, S.; Aram, M.H. Simulation of multipartite cavity quantum electrodynamics. IEEE J. Quant. Electron. 2013, 49, 1066. [Google Scholar] [CrossRef]
  52. Alidoosty, M.; Khorasani, S.; Aram, M.H. Theory and simulation of cavity quantum electro-dynamics in multi-partite quantum complex systems. Appl. Phys. A 2014, 115, 595. [Google Scholar] [CrossRef]
  53. Monifi, F.; Zhang, J.; Özdemir, Ş.K.; Peng, B.; Liu, Y.-X.; Bo, F.; Nori, F.; Yang, L. Optomechanically induced stochastic resonance and chaos transfer between optical fields. Nat. Photon. 2016, 10, 399. [Google Scholar] [CrossRef]
  54. Neill, C.; Roushan, P.; Fang, M.; Chen, Y.; Kolodrubetz, M.; Chen, Z.; Megrant, A.; Barends, R.; Campbell, B.; Chiaro, B.; et al. Ergodic dynamics and thermalization in an isolated quantum system. Nat. Phys. 2016, 12, 1037. [Google Scholar] [CrossRef]
  55. Thompson, J.D.; Zwickl, B.M.; Jayich, A.M.; Marquardt, F.; Girvin, S.M.; Harris, J.G.E. Strong dispersive coupling of a high-finesse cavity to a micromechanical membrane. Nature 2008, 452, 72. [Google Scholar] [CrossRef] [PubMed]
  56. Sankey, J.C.; Yang, C.; Zwickl, B.M.; Jayich, A.M.; Harris, J.G.E. Strong and tunable nonlinear optomechanical coupling in a low-loss system. Nat. Phys. 2010, 6, 707. [Google Scholar] [CrossRef]
  57. Nunnenkamp, A.; Børkje, K.; Harris, J.G.E.; Girvin, S.M. Cooling and squeezing via quadratic optomechanical coupling. Phys. Rev. A 2010, 82, 021806. [Google Scholar] [CrossRef]
  58. Lei, C.U.; Weinstein, A.J.; Suh, J.; Wollman, E.E.; Kronwald, A.; Marquardt, F.; Clerk, A.A.; Schwab, K.C. Quantum nondemolition measurement of a quantum squeezed state beyond the 3 dB limit. Phys. Rev. Lett. 2016, 117, 100801. [Google Scholar] [CrossRef] [PubMed]
  59. Liao, J.Q.; Nori, F. Photon blockade in quadratically coupled optomechanical systems. Phys. Rev. A 2013, 88, 023853. [Google Scholar] [CrossRef]
  60. Liao, J.Q.; Nori, F. Single-photon quadratic optomechanics. Sci. Rep. 2014, 4, 6302. [Google Scholar] [CrossRef] [PubMed]
  61. Kim, E.J.; Johansson, J.R.; Nori, F. Circuit analog of quadratic optomechanics. Phys. Rev. A 2015, 91, 033835. [Google Scholar] [CrossRef]
  62. Purdy, T.P.; Brooks, D.W.; Botter, T.; Brahms, N.; Ma, Z.Y.; Stamper-Kurn, D.M. Tunable cavity optomechanics with ultracold atoms. Phys. Rev. Lett. 2010, 105, 133602. [Google Scholar] [CrossRef] [PubMed]
  63. Shen, Z.; Zhang, Y.-L.; Chen, Y.; Zou, C.-L.; Xiao, Y.-F.; Zou, X.-B.; Sun, F.-W.; Guo, G.-C.; Dong, C.-H. Experimental realization of optomechanically induced non-reciprocity. Nat. Photon. 2016, 10, 657. [Google Scholar] [CrossRef]
  64. Brawley, G.A.; Vanner, M.R.; Larsen, P.E.; Schmid, S.; Boisen, A.; Bowen, W.P. Nonlinear optomechanical measurement of mechanical motion. Nat. Commun. 2016, 7, 10988. [Google Scholar] [CrossRef] [PubMed]
  65. Leijssen, R.; La Gala, G.; Freisem, L.; Muhonen, J.T.; Verhagen, E. Nonlinear cavity optomechanics with nanomechanical thermal fluctuations. arxiv, 2016; 1612.08072. [Google Scholar]
  66. Branginskii, V.B.; Manukn, A.B. Ponderomotive effects of electromagnetic radiation. Sov. Phys. JETP 1967, 25, 653. [Google Scholar]
  67. Cheung, H.K.; Law, C.K. Nonadiabatic optomechanical Hamiltonian of a moving dielectric membrane in a cavity. Phys. Rev. A 2011, 84, 023812. [Google Scholar] [CrossRef]
  68. Castaños, L.O.; Weder, R. Equations of a moving mirror and the electromagnetic field. Phys. Scr. 2015, 90, 068011. [Google Scholar] [CrossRef]
  69. Wilson, C.M.; Johansson, G.; Pourkabirian, A.; Simoen, M.; Johansson, J.R.; Duty, T.; Nori, F.; Delsing, P. Observation of the dynamical Casimir effect in a superconducting circuit. Nature 2011, 479, 376. [Google Scholar] [CrossRef] [PubMed]
  70. You, J.Q.; Nori, F. Atomic physics and quantum optics using superconducting circuits. Nature 2011, 474, 589. [Google Scholar] [CrossRef] [PubMed]
  71. Schleich, W. Quantum Optics in Phase Space; Wiley-VCH: Berlin, Germany, 2001. [Google Scholar]
  72. Khorasani, S. A proof for Poisson bracket in non-commutative algebra of quantum mechanics. Electron. J. Theor. Phys. 2016, 13, 57. [Google Scholar]
  73. Khorasani, S. Applied Quantum Mechanics; Delarang: Tehran, Germany, 2010. [Google Scholar]
  74. Hetherington, J.H.; Willard, F.D.C. Two-, three-, and four-atom exchange effects in bcc 3He. Phys. Rev. Lett. 1975, 35, 1442. [Google Scholar] [CrossRef]
  75. Roger, M.; Hetherington, J.H.; Delrieu, J.M. Magnetism in solid 3He. Rev. Mod. Phys. 1983, 55, 1. [Google Scholar] [CrossRef]
  76. Agarwal, G.S.; Jha, S.S. Theory of optomechanical interactions in superfluid He. Phys. Rev. A 2014, 90, 023812. [Google Scholar] [CrossRef]
  77. Suh, J.; Weinstein, A.J.; Schwab, K.C. Optomechanical effects of two-level systems in a back-action evading measurement of micro-mechanical motion. Appl. Phys. Lett. 2013, 103, 052604. [Google Scholar] [CrossRef]
  78. Roelli, P.; Galland, C.; Piro, N.; Kippenberg, T.J. Molecular cavity optomechanics as a theory of plasmon-enhanced Raman scattering. Nat. Nanotechnol. 2016, 11, 164. [Google Scholar] [CrossRef] [PubMed]
  79. Bagci, T.; Simonsen, A.; Schmid, S.; Villanueva, L.G.; Zeuthen, E.; Appel, J.; Taylor, J.M.; Sørensen, A.; Usami, K.; Schliesser, A.; et al. Optical detection of radio waves through a nanomechanical transducer. Nature 2014, 507, 81. [Google Scholar] [CrossRef] [PubMed]
  80. Lannebere, S.; Silveirinha, M.G. Wave instabilities and unidirectional light flow in a cavity with rotating walls. Phys. Rev. A 2016, 94, 033810. [Google Scholar] [CrossRef]
  81. Abbott, B.P.; Abbott, R.; Abbott, T.D.; Abernathy, M.R.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R.X.; et al. Observation of Gravitational Waves from a Binary Black Hole Merger. Phys. Rev. Lett. 2016, 116, 061102. [Google Scholar] [CrossRef] [PubMed]
  82. Paraïso, T.K.; Kalaee, M.; Zang, L.; Pfeifer, H.; Marquardt, F.; Painter, O. Position-squared coupling in a tunable photonic crystal optomechanical cavity. Phys. Rev. X 2015, 5, 041024. [Google Scholar] [CrossRef]
  83. Kalaee, M.; Paraiso, T.K.; Pfeifer, H.; Painter, O. Design of a quasi-2D photonic crystal optomechanical cavity with tunable, large x2-coupling. Opt. Express 2016, 24, 21308. [Google Scholar] [CrossRef] [PubMed]
  84. Metelmann, A.; Clerk, A.A. Nonreciprocal photon transmission and amplification via reservoir Engineering. Phys. Rev. X 2015, 5, 021025. [Google Scholar] [CrossRef]
  85. Gütig, R.; Eberlein, C. Dynamical Casimir effect with Dirichlet and Neumann boundary conditions. J. Phys. A 1998, 31, 6819. [Google Scholar] [CrossRef]
  86. Khorasani, S. Higher-order interactions in quantum optomechanics: Analytical solution of nonlinearity. arxiv, 2017; 1702.04982. [Google Scholar]
  87. Khorasani, S.; Rashidian, B. Guided light propagation in dielectric slab waveguide with conducting interfaces. J. Opt. A 2001, 3, 380. [Google Scholar] [CrossRef]
  88. Khorasani, S.; Rashidian, B. Modified transfer matrix method for conducting interfaces. J. Opt. A 2002, 4, 251. [Google Scholar] [CrossRef]
  89. Karimi, F.; Khorasani, S. Optical modulation by conducting interfaces. IEEE J. Quant. Electron. 2013, 49, 607. [Google Scholar] [CrossRef]
  90. Gardiner, C.W.; Zoller, P. Quantum Noise; Springer: Berlin, Germany, 2004. [Google Scholar]

Share and Cite

MDPI and ACS Style

Khorasani, S. Higher‐Order Interactions in Quantum Optomechanics: Revisiting Theoretical Foundations. Appl. Sci. 2017, 7, 656. https://doi.org/10.3390/app7070656

AMA Style

Khorasani S. Higher‐Order Interactions in Quantum Optomechanics: Revisiting Theoretical Foundations. Applied Sciences. 2017; 7(7):656. https://doi.org/10.3390/app7070656

Chicago/Turabian Style

Khorasani, Sina. 2017. "Higher‐Order Interactions in Quantum Optomechanics: Revisiting Theoretical Foundations" Applied Sciences 7, no. 7: 656. https://doi.org/10.3390/app7070656

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop