<<

Exploring corrections to the Optomechanical Hamiltonian

Kamila Sala and Tommaso Tufarelli∗ Centre for the Mathematics and Theoretical of Quantum Non-Equilibrium Systems, School of Mathematical Sciences, The University of Nottingham, University Park, Nottingham NG7 2RD, United Kingdom

We compare two approaches to refine the “linear model” of cavity , in order to describe effects that are beyond first order in the coupling constant. We compare corrections derived from (I) a widely used phenomenological Hamiltonian that conserves the number and (II) a two-mode trunca- tion of C. K. Law’s microscopic model, which we take as the “true” Hamiltonian of the system. While these approaches agree at first order, the latter model does not conserve the photon number, resulting in challenging computations from second order onwards. Our numerics suggest that the phenomenological Hamiltonian sig- nificantly improves the linear model, yet it does not fully capture all second-order corrections arising from the C. K. Law model. We conclude that, even when the mechanical frequency is much lower than the cavity one, photon number conservation must be eventually given up to model cavity optomechanics with high accuracy.

I. INTRODUCTION ω(x) ' ω(0) + x ω0(0) (equivalently, this may be seen as a first order expansion in the coupling constant). We shall call linear model the resulting Hamiltonian, arguably the most widely Cavity quantum optomechanics is a rapidly developing re- used modelling tool in SC optomechanics. search field exploring the interaction of quantised light with Despite the undeniable success of the linear model as a theo- macroscopic mirrors, membranes and levitated nano-objects retical device, the need is arising to go beyond this approx- through radiation pressure [1]. The field has recently wit- imation. For example, it was noted by Brunelli et al. [6] nessed impressive experimental progress, including demon- that the linearized optomechanical Hamiltonian is unbounded strations of near-ground-state cooling of mechanical oscilla- from below, with (unphysical) negative energies cropping up tors and normal-mode splitting due to -photon inter- at very high photon numbers [22]. While this pathology may action [2]. Among its many applications, optomechanics is a be seen as somewhat mathematical (see Appendix A), it high- promising platform to demonstrate nonclassical behaviour in lights that the model must be eventually refined, particularly massive objects [6, 7], and may provide novel avenues for the in view of future experiments aiming to achieve ever larger coherent manipulation of quantum light [17, 21]. Recently it coupling strengths. Even in systems that are far from the SC, was also suggested that optomechanics may allow for ultra- the quest for ultra-precise measurements (e.g. those pertain- high precision measurements, possibly probing Planck-scale ing Planck scale physics) [13], or for the detection of dynami- corrections to the canonical commutation relations [12, 13]. cal Casimir effects [14], demand a more accurate Hamiltonian The simplest optomechanical setup features a single optical description of quantum optomechanics. cavity mode, whose frequency ω(x) depends parametrically Armed with these motivations, in this paper we explore and on the coordinate x (the ‘position’) of a mechanical oscillator compare two different starting points that may be used to go [23]. Perhaps the most notable example is given by a Fabry- beyond the linear model: (I) a widely used phenomenologi- Perot resonator with one movable mirror — see Fig. 1. cal Hamiltonian, which conserves the cavity photon number Most experiments to date have explored the weak coupling (phenomenological approach); (II) a two-mode truncation of regime, in which radiation pressure effects are only visible C. K. Law’s microscopic Hamiltonian for an optomechani- upon strong driving of the cavity, and the system can be de- cal Fabry-Perot cavity [23] (microscopic approach). We shall scribed as a pair of coupled harmonic oscillators [1]. Sev- take this truncated Law Hamiltonian as the benchmark (or eral experimental platforms, however, are now approaching “true model”) against which the various generalizations of the the single-photon strong coupling regime [2–4, 11, 15, 16] linear model should be judged. In this paper we focus on

arXiv:1711.06688v1 [quant-ph] 17 Nov 2017 (strong coupling, or SC, for brevity), in which the anhar- the realistic situation where the bare cavity frequency is much monic nature of the radiation pressure interaction must be larger than the mechanical one. We find that approaches I and taken into account. Strong coupling occurs when the the cou- II agree at first order in the coupling, but at higher orders the pling rate of a single photon is greater than the typical loss microscopic model yields ‘counter-rotating’ terms that vio- rates of the setup. Such regime features clear departures from late photon number conservation [14] and make computations classical behaviour [9, 11, 17], and facilitates the production challenging. We resort to numerical diagonalization of the of nonclassical states in both light and mechanics [10, 20]. Hamiltonian, truncated in Fock space, to deal with such situ- A further appealing feature of strong coupling optomechan- ations (and also to tackle the exact phenomenological Hamil- ics is that the Hamiltonian can be diagonalized analytically tonian). The numerical examples we explored suggest that upon linearising the cavity frequency around the origin, as per the phenomenological approach does a good job in improving the linear model in the typical parameter regimes of optome- chanics experiments, yet it does not fully capture the second- order corrections arising from the microscopic treatment. We ∗Electronic address: [email protected] conclude that photon number conservation must be eventually 2

terms of G. In typical experimental realizations, the hierarchy ω0  Ω  g0 holds between the model parameters. Hamiltonian (1) can be diagonalized exactly [10, 11]. The properties of the corresponding eigensystem are well known, and can be obtained from Section III B by setting β = 0.

III. PHENOMENOLOGICAL APPROACH

A first possibility in extending the linear model, often adopted in the literature, is to consider a phenomenological Hamilto- nian similar to Eq. (1), but where the cavity frequency has a generic position dependence ω(x). Crucially, no change is assumed in the commutation rules for mirror and field oper- ators, an assumption that drastically simplifies computations FIG. 1: Schematic of an optomechanical setup consisting of a Fabry- but may not be justified microscopically [23] — we shall elab- Perot resonator with a movable mirror. The length of the cavity, orate on this point in Section IV. hence its frequency ω(x), is modulated by the mirror coordinate x. For the purposes of illustration, from now on we consider The movable mirror is subject to a trapping potential (depicted as a spring), which is usually harmonic to a good approximation. an optomechanical setup consisting of a Fabry-Perot resonator with one movable mirror (see Fig.1), which amounts to setting ω(x) = ω0/(1 + x/L) with L the bare cavity length. However, given up to model cavity optomechanics with high accuracy, the reader should keep in mind that the same formalism is even in parameter regimes where dynamical Casimir physics applicable when ω(x) is a generic (positive) function. The would be usually discounted. phenomenological Hamiltonian is The paper is organised as follows. In Sec. II we review the ! 2 † 1 pˆ 1 2 2 linear model of optomechanics. In Sec. III we show how this Hphen = ω(x ˆ) aˆ aˆ + + + Ω xˆ . (2) 2 2 2 model can be refined phenomenologically, while in Sec. IV we discuss model corrections that are microscopically mo- If we now pick L = ω0/G = (ω0/g0)xzpf and take a first order tivated. Sec. V features numerical examples comparing the expansion in G (or equivalently g0), Eq. (1) is recovered. The various approaches. We draw our conclusions in Sec. VI, and natural next step is then to retain higher order corrections in G. in Appendix A we discuss in more detail the negative-energies In what follows, we discuss two inequivalent approaches that pathology of the linear model which was outlined in [6]. employ simple mathematics to approxi- mate Eq. (2) beyond first order. Both methods are in principle amenable to analytical treatment. II. LINEAR MODEL: BRIEF REVIEW

We begin by recalling the model that is commonly taken as A. Approach 1: photon - controlled harmonic oscillator a starting point in modelling optomechanics. To lighten the notation we take ~ = 1, and consider a mechanical mode of Our first approach involves the recalculation of the mechan- unit effective mass. The linear model Hamiltonian reads ical equilibrium position and spring constant as functions of the photon number n. Indeed, exploiting the conservation law ! 2 † † 1 pˆ 1 2 2 [H, aˆ aˆ] = 0, we may decompose Hlin = (ω0 − Gxˆ) aˆ aˆ + + + Ω xˆ , (1) 2 2 2 ∞ " # X pˆ2 H = |nihn| ⊗ + V (x ˆ) (3) wherea ˆ is the annihilation operator for the cavity field mode, phen 2 n n=0 xˆ andp ˆ are respectively the canonical position and momen- tum operators of the movable mirror, Ω is the bare mechan- featuring the effective n− dependent mechanical potential ical frequency, ω0 is the bare cavity frequency, and the cou- ! 1 1 2 2 pling constant G is known as frequency pull parameter. The Vn(x ˆ) ≡ ω(x ˆ) n + + Ω xˆ . (4) canonical commutators read [ˆa, aˆ†] = 1, [x ˆ, pˆ] = i, while all 2 2 field operators commute with all mirror operators. In passing The idea is now to perform a quadratic expansion of Vn(x ˆ) we note that the optomechanical interaction strength is often around its minimum: quantified via the vacuum optomechanical√ coupling strength 1 00 2 g0 = Gxzpf, where xzpf = 1/ 2Ω is the (bare) Vn(x ˆ) ' Vn(x ¯n) + Vn (x ¯n) (x ˆ − x¯n) , (5) uncertainty of the mechanical oscillator. Loosely speaking, 2 the dimensionless ratio g0/Ω quantifies the influence a single where the equilibrium positionx ¯n is implicitly determined by photon can have on the mirror motion. That being said, many 0 of our expressions will have a simpler form when written in Vn(x ¯n) = 0. (6) 3

Note that the above procedure is well defined, since it is easily verified that Vn has a unique global minimum at fixed n, and 00 that for all n, Vn (x ¯n) > 0 (see below). The outlined procedure returns the effective Hamiltonian ∞ ! X pˆ2 1 H ≡ |nihn| ⊗ + Ω2(x ˆ−x¯ )2 + V (x ¯ ) , (7) nˆ 2 2 n n n n n=0 | {z } Hn 2 00 Ωn ≡ Vn (x ¯n), (8) where we have chosen the symbol Hnˆ to signify that the pho- FIG. 2: In the two analytically solvable models of Section III, the op- † ton number operatorn ˆ = aˆ aˆ controls which harmonic os- tomechanical eigenstates |Ψn,mi possess a qualitatively similar struc- ture. As the photon number is increased going from (a) to (b), the cillator Hamiltonian Hn is acting on the mirror. For each average mechanical position is pushed further away from the fixed n, the eigenvalues and eigenvectors of Hn may be found via straightforward harmonic oscillator techniques. For example mirror, while the position uncertainty drecreases (i.e., the mechani- cal wavefunctions get squeezed). Note that the latter squeezing effect the energy eigenvalues, labelled by the cavity photon number is not present in the linear model. n and a mechanical ‘phonon’ number m (both semipositive integers), read ! φ x hx|φ i 1 where the mechanical wavefunctions n,m( ) = n,m are En,m = Vn(x ¯n) + Ωn m + . (9) expressible in terms of Hermite functions centered at x = x¯n 2 (details not shown). The following eigenstate properties are easily verified: These are all positive, and thus may be used to calculate the partition function without the need for artificial cutoffs [6]. A hxˆinm ≡ hΨn,m| xˆ |Ψn,mi = x¯n, (13) fully analytical treatment of Hnˆ requires the explicit inversion m + 1/2 of Eq. (6), which may be recast as ∆x2 ≡ hΨ |(x ˆ − hxˆi )2|Ψ i = . (14) nm n,m nm n,m 2Ω ! n 1 ω /L Ω2 x¯ = n + 0 . (10) n 2 In light of the above discussion we gain the following intuition 2 (1 + x¯n/L) about the structure of the Hamiltonian eigenstates in Eq. (12): This equation forx ¯n is in principle solvable (it is equivalent to the mechanical wavefunctions φn,m(x) are pushed further away a third-degree polynomial equation) but leads to very cumber- from the fixed mirror as the photon number n is increased, some expressions. A number of observations, however, can while at the same time they become more and more squeezed be made without resorting to the explicit solution. First, it is in position. Such behaviour is schematised in Fig. 2 evident that it must bex ¯n > 0, and a little more thought reveals thatx ¯n → ∞ when n → ∞. Second, notice that Eq. (10) de- scribes the intersection between a strictly increasing function B. Approach 2: standard quadratic expansion and a decreasing one, such that for any n there is a unique real solution forx ¯n, as anticipated. Third, implicit differentiation Our second approach consists in a more traditional second- in n shows thatx ¯n is a strictly increasing function of the photon order expansion of Eq. 2. Specifically, we expand number (as intution about radiation pressure would suggest). Finally, combining Eqs. (8) and (10) the photon-dependent 1 ω(x) ' ω − Gx + β2 x2, (15) mechanical frequencies may be expressed as 0 2 " # 2x ¯ /L 2 2 2 2 2 n where β = 2ω0/L = 2G /ω0. For brevity, we shall refer to Ωn = Ω 1 + , (11) (1 + x¯n/L) the resulting Hamiltonian as the quadratic model: ! ! which is a strictly increasing function ofx ¯n, hence n. It is an 1 1 pˆ2 1 H = ω − Gxˆ + β2 xˆ2 aˆ†aˆ + + + Ω2 xˆ2. (16) interestingly mathematical√ fact that we obtain the elegant lim- quad 0 2 2 2 2 iting value limn→∞ Ωn = 3Ω, which does not depend on the cavity parameters ω0 and L. However, this result is unlikely While the ‘quadratic coupling’ has been studied before in op- to have any physical significance: the above model may even- tomechanics [18], the focus has often been on the striking tually become unreliable when too many occupy the differences between a purely linear (β = 0) and a purely cavity (for example, the associated limitx ¯n → ∞ means that quadratic (G = 0) interaction. Here, on the other hand, our the average cavity length becomes infinite!). The eigenstates primary goal is to improve the accuracy of our Hamiltonian corresponding to En,m take the form model: we want to explore the corrections to the linear model due to the higher order β term. This is in a similar spirit to |Ψn,mi ≡ |ni|φn,mi, (12) Ref. [13], where it was argued that the inclusion of higher 4 order terms is necessary to investigate Planck-scale modifica- physics than Hamiltonian (2). In the next section, Eq. (23) will tions to the canonical commutator. Taking Eq. (16) as a start- play the role of benchmark, or “true model”, against which our ing point, the relevant calculations for the quadratic model various approximations shall be compared. In the present sec- follow closely those of Sec. III A, mutatis mutandis. What tion, we instead discuss the corrections to the linear model that is more, we can here obtain compact analytical expressions. arise in this framework, highlighting the issues that were not Specifically, we may recast the Hamiltonian in the form (7), present in the phenomenological approach. Note that Hamil- with parameters tonian (25) does not yield photon number conservation in gen- ! eral. More precisely, the concept of photon itself relies on 2 2 1 2 identifying a reference frequency, for which a natural choice Ωn = Ω + n + β , (17) 2 is ω0 ≡ ω(0). We then choose the following definition for the G(n + 1/2) cavity annihilation operator: x¯ = , (18) n 2 r Ωn ω i aˆ = 0 Qˆ + √ Πˆ . (24) so that the resulting energy eigenvalues read 2 2ω0 ! ! 1 1 G2(n + 1/2)2 Sincea ˆ commutes with all mirror operators, it allows us to de- E n ω m − , n,m = + 0 + Ωn + 2 (19) fine a Fock space for the cavity field that does not depend on 2 2 2Ωn the mirror coordinatex ˆ. We found this approach convenient to valid for all semipositive integers n, m. Also here we shall dis- set-up numerical calculations based on the truncation of Fock cuss the limit n → ∞ as a matter of mathematical curiosity. spaces. In passing we note that this was not the path taken in Note that the quadratic model predicts a finite average me- the seminal paper by C. K. Law [23], where insteadx ˆ− depen- chanical displacement even in the limit of large n, while the dent annihilation operators and Fock spaces were introduced to describe the cavity field. Hamiltonian (23) may be recast as mechanical√ frequencies are divergent with asymptotic scaling Ωn ∝ n, in turn implying that the mechanical wavefunctions ! 2 † 1 λ(x ˆ)  2 †2 pˆ 1 2 2 become infinitely squeezed in position: Hmic = ν(x ˆ) aˆ aˆ + + aˆ +aˆ + + Ω xˆ , (25) 2 2 2 2

G 2 2 2 2 lim hxˆinm = lim x¯n = , (20) where ν(x ˆ) = (ω(x ˆ) + ω0)/2ω0, λ(x ˆ) = (ω(x ˆ) − ω0)/2ω0 n→∞ n→∞ β2 and the cavity photon numbera ˆ†aˆ is evidently not conserved lim Ωn = +∞, (21) 2 †2 n→∞ due to the appearance of counter-rotating terms (i.e.a ˆ + aˆ ). Let us now discuss how to approximate Hamiltonian (23) at lim ∆xnm = 0. (22) n→∞ first and second order in the coupling G. A straightforward first-order expansion of Eq. (25) yields Gxˆ   IV. MICROSCOPIC APPROACH H(1) = H − aˆ2 +aˆ†2 ' H , (26) mic lin 2 lin In this section we take as a starting point the microscopic where the linear model (1) has been recovered through Hamiltonian derived by C. K. Law through canonical quanti- a rotating-wave approximation, justifiable in the parameter zation of a Fabry-Perot cavity with a movable mirror [23]. For regime ω0  Ω  g0 that we are considering here. More the sake of tractability we shall assume that only two degrees in detail, a back-of-the-envelope calculation shows that the of freedom, one optical and one mechanical, are sufficient to counter-rotating terms only provide a contribution at order  2 2  model our optomechanical system. Under this assumption, O G xˆ /ω0 — for example this can be proved by time- the Law Hamiltonian takes the simplified form averaging the Hamiltonian on timescales  1/ω0 [24]. In the parameter regimes of interest we can thus reach the reas- ˆ 2 2 Π 1 2 2 pˆ 1 2 2 Hmic = + ω (x ˆ)Qˆ + + Ω xˆ , (23) suring conclusion that the microscopic and phenomenological 2 2 2 2 approaches yield the same linear model. We should however point out that there can be situations where ω and Ω have where Qˆ, Πˆ are canonical quadrature operators for the cavity 0 comparable magnitude, and in that case the (ˆa2 + aˆ†2) term field mode and ω(x ˆ) retains the same form as before. The cannot be neglected even at first order in G [14]. elementary canonical commutator for the cavity field is now It is at second order in G that non-negligible discrepancies [Qˆ, Πˆ ] = i, and again all field operators commute with the arise between the microscopic and phenomenological ap- mechanical variablesx ˆ andp ˆ. Note that Hamiltonian (23) im- proaches. Following similar lines of reasoning as above, we plies d pˆ ∝ Qˆ 2, i.e., the instantaneous radiation pressure force dt find that the second-order microscopic Hamiltonian is is now proportional to the intensity of a single field quadrature ! (rather than to the field energy). This makes sense: since the β2 xˆ2 1 Gxˆ   H(2) = H + aˆ†aˆ + − aˆ2 +aˆ†2 , (27) (transverse) electric field E must vanish on the mirror, the ra- mic quad 4 2 2 diation pressure force will be proportional to the squared mag- 2 2 2 2 netic field |B| alone (evaluated at the mirror surface). Hence, where we recall β = 2ω0/L = 2G /ω0, and we note two within the two-mode approximation, we argue that Eq. (23) additional terms that were not present in the phenomenolog- should be a more reliable description of radiation pressure ical model and cannot be neglected at the considered order. 5

Due to the counter-rotating terms, Hamiltonian (27) presents (iii) Mechanical uncertainty, ∆xn,0 the same computational challenges as the original model in (2) Eq. (23) — albeit its numerical solution appears to be slightly Note that the eigensystems of Hamiltonians Hlin, Hphen, Hnˆ are faster in the examples we explored. The relevance of Eq. (27) fully characterized by these quantities. In contrast, n is not a to our discussion is more conceptual than practical: it high- (2) good quantum number for Hmic and Hmic, since these Hamil- lights that the phenomenological approach of Section III, tonians do not commute withn ˆ. Yet, in the parameter regimes when taken beyond first order in G, may begin to miss some here considered we were able define an effective photon num- of the finer details of optomechanical physics. In particular, ber n as the integer closest ton ¯ = haˆ†aˆi, since |n − n¯| remained the appearance of counter-rotating terms provides a qualita- small in all the reported examples. In turn, this allowed us to tive departure from the phenomenological model. identify the lowest eigenvalue at fixed n, En,0, as well as the Luckily, the examples to follow indicate that the analytical associated eigenstate |Ψn,0i. From these we were able to de- models derived via the phenomenological approach may still fine, and calculate, quantities (i)-(iii) for the models Hmic and provide significant improvements to the linear model in the (2) Hmic. The top half of Figure 3 shows that all the approaches regime ω0  Ω  g0. considered here provide a significant improvement to the lin- (2) ear model, but at this stage there is no clear winner: while Hmic is better at predicting energy eigenvalues, the phenomenologi- V. NUMERICAL EXAMPLES cal Hamiltonian Hphen can predict mechanical equilibrium po- sitions and (most) variances more accurately. In passing, it So far we have introduced several Hamiltonians that aim to is interesting to note that Hnˆ can be seen as a sort of hybrid improve the linear model. It is now interesting to compare model emulating properties of both Hquad (equilibrium posi- these refined models in a concrete example: first by examin- tions) and Hphen (mechanical variances). ing a few observables that characterize their low-energy eigen- systems, then by looking at their predictions for the time evo- lution of mirror position and field amplitude. Since we are B. Dynamics looking at small deviations from the predictions of Hlin, it is practically impossible to compare by eye the plots of absolute To compare dynamical predictions we take the initial state quantities. Instead, for any quantity A calculated via a given approximate model, we shall plot the absolute error |A − A |, mic |Ψ(0)i = |αia|0ib, (28) where Amic is computed by solving Hmic numerically. Note (2) that numerical solutions are adopted also for H and H . −|α|2/2 P n phen mic where |αia = e n(α /n!)|nia is a coherent state and For all these cases we wrote a simple Mathematica script to |0ib is the bare ground state of the mechanics. It is apparent diagonalize the Hamiltonian in a truncated space and hence that the evolution of initial state (28) is going to involve sev- calculate all quantities of interest. To confirm the solidity of eral photon number sectors even with the number-conserving these results we performed the same calculations in Python, Hamiltonians. Moreover, according to the linear model, the which produced visually identical plots. For our purposes evolution of this state can induce interesting nonclassical fea- truncations of nmax = 20 photons and mmax = 30 tures in the system, such as Wigner-Function negativity and were sufficient to obtain convergence, but these numbers rely opto-mechanical entanglement [10]. Therefore, we think that heavily on the chosen model parameters, and on the number the initial state (28) embodies an interesting test-bed for our of reliable eigenvalues one is seeking. various approximate Hamiltonians: a reliable description of In the displayed examples we set Ω = 100g0, inspired by the subsequent dynamics must capture well the evolution of state-of-the-art experiments featuring g0/Ω . 0.01 [2–4], self-correlations in the cavity field together with entanglement while the bare cavity frequency is ω0 = 100Ω. Qualitatively between the two subsystems. In our examples we picked 3 similar results were obtained by increasing ω0 to 10 Ω. With α = 2, and looked at the accuracy of the various models in our current numerical means we found it difficult to obtain re- predicting the time-dependent averages hxˆ(t)i and haˆ(t)i: the liable results when increasing ω0 further: the handling of such relevant results are displayed in the bottom half of Figure 3. cases required prohibitively high numerical precisions. It may (2) From these we infer that Hmic has an edge on the other mod- be useful to overcome this limitation in the future, in order els, as it may be expected from its construction as the ‘correct’ to address systems where many orders of magnitude separate second-order expansion of Hmic. As a corollary, we argue that mechanical and optical frequencies. the inclusion of counter-rotating terms will eventually become necessary in trying to model optomechanical system with ever higher accuracy. On the other hand, the good news is that once A. Low-energy eigensystems again the phenomenological approach provides a significant improvement to the linear model (except for a few isolated To compare eigensystems, we focus on the following quanti- points in time, where the linear model yields better estimates ties that depend on the photon number n: for hxˆ(t)i). Furthermore, in the chosen parameter regimes we do not observe significant differences between Hphen, Hnˆ and (i) Lowest energy eigenvalue, En,0 Hquad. Under these circumstances, our choice would naturally (ii) Average mechanical displacement, hxˆin,0 fall on Hquad due to its simple analytical solution. 6

Error in n,0 Error inE n,0 Error inΔx n,0 1.2× 10 -7 □△▽ -9 ▽△ -10 △ -7 -5 x 4.× 10 x -6 3.× 10 10 x 5.×10 △ x 7.×10 -6 1.× 10 x x ▽ -10 x x □△▽ x 2.5× 10 x x -8 x -9 x ▽△ x △ 8.× 10 x 3.× 10 x x x x x x x ▽△ -10 x x □△▽ ○□△▽x ○ □▽△x ○ □▽△x ○ □▽△○ □▽△○ □▽△○ □▽△○ □▽△○ □▽△○ □▽△ 0 2.× 10 x x ○ -8 0 ○□△▽x ○□△▽ ○□△▽ ○□△▽ ○□△▽ ○□△▽ ○□△▽ ○□△▽ ○□△▽ ○□△▽ ▽△ ○△□□□□□□□□□□▽○△▽○△▽○△▽○▽△○▽△○▽△○▽△○▽△○▽△ 0 6.× 10 -9 0 5 ○ -10 0 5 △ 0 5 □△▽ 2.× 10 ▽△ ○ 1.5× 10 ○ -8 ○ 4.× 10 □△▽ ▽△ ○ -10 △ ○ -9 ○ 1.× 10 ○ -8 □△▽ 1.× 10 ▽△ ○ △ □▽ 2.× 10 ○ -11 ○ □▽ □▽ □△▽ ▽△ ○ 5.× 10 ○□▽△ □▽ □▽ □△▽ □△▽ ○ □▽ ○□▽△ 0 ○□△▽ ○ ○ ○ ○ ○ ○ ○ ○ ○ 0 ○□△▽ □ □ □ □ □ □ □ □ □ 0 ○△□▽ ○□△▽ ○△ 0 2 4 6 8 2 4 6 8 10 0 2 4 6 8 photon number photon number photon number

Error in Error in 0.000020 x xx xx xxxxx xxx x 0.00005 xxx xxxxx x xxx xxx 0.000015 x x xx xx x 0.00004 x xx x x xxxxxxxxxxxxx xx x x x xx xx x xx xxxx x x 0.00003 xx 0.000010 x x x xxxxxxx xx x x x x xx xxxxxxx x x x x x xx x 0.00002 xxxx xx xxxx x x x x x x x x△□△□△□△□xx xx △□△□△□△□△□ △□△□△□△□△□ △□△□△□△□△□ -6 xx x x △□x△□ △□□△x xx △□ △□□△ △□ △□□△ △□ △□△□ 5.× 10 x x x x △□x □△x xx △□ □△ △□ □△ △□ □△ x x x x x x x △□x □△xxxxx △□ □△ △□ □△ △□ □△ x x xx x x xx x 0.00001 △□x □△ △□ □△ △□ □△ △□ □△ x▽△□△▽□△▽□△△▽▽□□△▽□△▽□△▽□ x x x▽△□▽△□▽△□△△▽▽□□△▽□△▽□△▽□ x ▽△□▽△□▽△□▽△□△▽□△▽□△▽□△▽□ ▽△□▽△□▽△□▽△□△▽□△▽□△▽□△▽□ △□x □△ △□ □△ △□ □△ △□ □△ x▽△□x▽△□ △▽□△▽□ x ▽△□x▽△□x △▽□△▽□ x ▽△□▽△□x △▽□△▽□ ▽△□▽△□ △▽□△▽□ △□x □△ △□ □△ △□ □△ △□ □△ ▽△□x▽△□x▽△□ △▽□△▽□△▽□xx ▽△□▽△□x △▽□△▽□△▽□ ▽△□▽△□ x △▽□△▽□△▽□ ▽△□▽△□ △▽□△▽□△▽□ x△□x □△ △□ □△ △□ □△ △□ □△ 0.000000 ○○○○○○□△▽x▽△□x▽△□x▽△□x○○○○○○○○○○○○○○○○○△▽□○○○○○○○○△▽□△▽□▽□△x▽△□▽△□▽△□▽△□○○○○x○○○○○○○○○○○○○△▽□○○○○○○○○○△▽□△▽□▽□△x▽□△▽△□▽△□▽△□○○○○○○○○○○○○○○○○△▽□○○○○○○○○○△▽□△▽□▽□△x▽□△▽△□▽△□▽△□○○○○○○○○○○○○x ○○○○△▽□○○○△▽□△▽□▽□△x 0.00000 ○○○○○○○○○○○○○○○○○○○○○□△x□△x△□ ○○○○○○○○○○○○○○○○○○○○○□△□△△□△□△□ ○○○○○○○○○○○○○○○○○○○○○□△□△△□△□△□ ○○○○○○○○○○○○○○○○○○○○○□△△□△□△□△□ ○○○○○○○○○○○○○○○○○□△△□△□ 0 5 10 15 20 25 0 5 10 15 20 25 Ωt Ωt

FIG. 3: Errors of the various Hamiltonians in predicting low-energy eigensystems and dynamics for a two-mode optomechanical system. The microscopic model Hmic is taken as benchmark (or “true model”). For these plots we consider a Fabry-Perot optomechanical setup with 2 2 parameters ω0 = 10 Ω, Ω = 10 g0. The dynamics of haˆi and hxˆi are calculated for the initial state (28) with α = 2.

VI. CONCLUSIONS AND OUTLOOK high precision modelling, a multi-mode treatment (including several degrees of freedom for both light and mechanics) may eventually become necessary. Second, we have left system- We have proposed and compared several Hamiltonians that go environment interactions out of our discussion. Due to the beyond the linear model of optomechanics, featuring correc- presence of three very different energy scales (ω0, Ω, g0), we tions that are either phenomenologically or microscopically anticipate that it will be challenging to develop an accurate derived. Using a two-mode truncation of Law’s Hamiltonian open-system model for optomechanics that goes beyond the [23] as a benchmark, we were able to confirm that the phe- existing phenomenological treatments. Yet, identifying a re- nomenological approach, common in the literature, is able to liable Hamiltonian is a crucial first step towards the rigorous provide meaningful corrections to the linear model. A partic- modelling of an open quantum system [26]. Third, we have ularly reassuring observation is that, in the regime ωΩg0 assumed the bare mechanical potential to be harmonic. De- studied here, these corrections are well captured by the ana- pending on the implementation, anharmonicity in the mirror lytically solvable model Hquad, i.e., a straightforward second- potential [27] may provide new corrections that can be qual- order expansion of the phenomenological Hamiltonian (2). itatively different from those studied here. Finally, one may However, we also pointed out that the phenomenological ap- also explore relativistic corrections to the Law Hamiltonian proach does not fully capture all second order corrections aris- [28]. The theoretical challenge will ramp up significantly in ing from a microscopic treatment of radiation pressure. In trying to address any of these limitations of our work, yet it particular we have outlined an important effect, the break- will be a worthwhile pursuit in the quest to model ever more down of photon number conservation, which can provide sophisticated optomechanics experiments. non-negligible corrections even in parameter regimes where dynamical-Casimir effects are usually discounted. Our work may benefit future experiments probing the single- photon strong coupling regime or tackling high-precision op- Acknowledgments tomechnical measurements. More importantly, it opens a discussion that may uncover new avenues for the theoreti- We acknowledge fruitful discussions with G. Adesso, F. Ar- cal modelling of optomechanics, beyond the domain of ap- mata, A. Armour, S. Bose, L. A. C. Correa, A. Ferraro, M. plicability of the linear model. There are several important G. Genoni, O. Hess, M. S. Kim, S. P. Kumar, G. J. Milburn, questions that our work leaves open and that should be ad- F. Mintert, A. Nunnenkamp, M. Paternostro, M. Plenio, A. dressed in future studies. First, our analysis is confined to a Serafini, H. Ulbricht and especially Daniele Dorigoni. T.T. two-mode scenario. In the quest for strong coupling and/or acknowledges financial support from the Foundational Ques- 7

where both the frequency dependence of ω and the mechani- cal potential V can be left generic at this stage. We will only assume that V is a regular function, so that it cannot encode hard-wall boundary conditions (more on this point below). It is easy to show that, if there is a region for x where ω(x) < 0, the Hamiltonian cannot be lower-bounded. Indeed, let us sup- pose that ω(x) is not everywhere positive, and that there is a ground state with finite energy E0. For any quantum state it would then follow E0 ≤ hHi. In particular let us consider a product state |ni|ψi ≡ |nia ⊗ |ψib, where |nia indicates a fock state of the cavity field, while |ψib is a generic mechanical wavefunction. Taking the expectation value of H we obtain FIG. 4: An illustration of the pathology carried by the linear model. While a well-defined optomechanical model features ω(x) > 0 for ! ! 1 pˆ2 all x (blue continuous curve), the linearization of ω(x) (red-dashed E ≤ n + hψ|ω(x ˆ)|ψi + hψ| + V(x ˆ) |ψi. (A2) curve) produces a non-physical region (R) where the cavity fre- 0 2 2 quency is negative. It then possible to construct states of arbitrarily low energy by concentrating the mechanical probability density (or- R ange dotted line) in the region , and subsequently populating the If the cavity frequency can assume negative values in a cer- cavity with a large number of photons. tain region R ≡ {x | ω(x) < 0}, it is always possible to ob- tain hψ|ω(x ˆ)|ψi < 0, for example by choosing ψ(x) ≡ hx|ψi R tions Institute (Grant No. FQXi-RFP-1601), and from the as a smooth function with support in (See Fig. 4). At the University of Nottingham via a Nottingham Research Fellow- same time, the second term on the right-hand side (RHS) of ship. K. S. acknowledges financial support from the Univer- Eq. (A2) is a finite constant if the potential does not display sity of Nottingham through a Dr Margaret Jackson Summer singularities. It follows that the expression on the RHS tends −∞ → ∞ Bursary Award. to when n , so that E0 cannot be finite. In hindsight this result may appear obvious, as the very con- cept of a single-mode cavity implicitly assumes ω(x) > 0. Appendix A: Negative energies in the linear model Yet, the common linear approximation ω(x) ' ω0 − Gx fails to satisfy this requirement. Note that in the above discussion In this section we shed further light on the pathologies of we did not need to specify the functional form of V, hence the linear model of optomechanics, which to the best of our we can conclude that the outlined pathology cannot be cured knowledge were first outlined by [6]. We first take one step by adding anharmonic terms to the mechanical confining po- back and consider a generic optomechanical Hamiltonian of tential. The only way to retain a linearised ω(x), while also the form ensuring that H in Eq. (A1) is lower bounded, is to impose ! hard-wall boundary conditions on the oscillator, so that all 1 pˆ2 H = ω(x ˆ) aˆ†aˆ + + + V(x ˆ), (A1) mechanical wavefunctions are forced to vanish in the region 2 2 R.

[1] W. P. Bowen, G. J. Milburn, Quantum optomechanics, CRC 55, 3042 (1997). Press (2015). [11] A. Nunnenkamp, K. Borkje, S. M. Girvin, Phys. Rev. Lett. 107, [2] M. Aspelmeyer, T. J. Kippenberg and F. Marquardt, Rev. Mod. 063602 (2011). Phys. 86, 1391 (2014). [12] I. Pikovski, M. R. Vanner M R, M. Aspelmeyer, M. S. and Cˆ [3] F. Marquardt, Nature Physics 4, 513 (2008). Brukner, Nat. Phys. 8, 39339 (2012). [4] T. J. Kippenberg and K. J. Vahala, Science 321, 1172 (2008). [13] S. P. Kumar and M. B. Plenio, arXiv:1708.05659. [5] A. Cleland, Nature Physics 5, 458 (2009). [14] V. Macr`ı, A. Ridolfo, O. Di Stefano, A. F. Kockum, F. Nori, S. [6] M. Paternostro, Phys. Rev. Lett. 106, 183601 (2011). Savasta, arXiv:1706.04134. [7] M. R. Vanner, I. Pikovski, G. D. Cole, M. S. Kim, C. Brukner, [15] A.J. Rimberg, M.P. Blencowe, A.D. Armour, P.D. Nation, New K. Hammerer, G. J. Milburn, M. Aspelmeyer, Proc. Natl. Acad. J. Phys. 16, 055008 (2014). Sci. USA 108, 16182 (2011). [16] R. Leijssen, G. La Gala, L. Freisem, J. T. Muhonen, E. Verha- [8] S. Groblacher,¨ K. Hammerer, M. R Vanner, M. Aspelmeyer, gen, arXiv:1612.08072. Nature 460, 724 (2009). [17] S. Aldana, C. Bruder, and A. Nunnenkamp, Phys. Rev. A 88, [9] M. Ludwig, B. Kubala, F. Marquardt, New J. Phys. 10, 095013 043826 (2013). (2008). [18] A.M. Jayich et al. New J. Phys. 10, 095008 (2008); J. C. [10] S. Bose, K. Jacobs, and P. L. Knight, Phys. Rev. A 56, 4175 Sankey et al. Nature Physics 6, 707 (2010); A. Nunnenkamp, (1997); S. Mancini, V. I. Man’ko, and P. Tombesi, Phys. Rev. A K. Borkje, J. G. E. Harris, S. M. Girvin, Phys. Rev. A 82, 8

021806(R) (2010); J-Q. Liao and F. Nori, Scientific Reports 4, (2007). 6302 (2014). [25] T. Holstein and H. Primakoff, Phys. Rev. 58, 1098 (1940); K. [19] P. Rabl, Phys. Rev. Lett. 107, 063601 (2011). Hammerer, A. Sørensen, and E. Polzik, Rev. Mod. Phys. 82, [20] L. Latmiral, F. Mintert, arXiv:1705.10334. 1041 (2010). [21] C. Fabre et al., Phys. Rev. A 49, 1337 (1994), S. Mancini and [26] H. P. Breuer and F. Petruccione, The theory of open quantum P. Tombesi, Phys. Rev. A 49, 4055 (1994); systems, Oxford University Press (2002). [22] M. Brunelli, A. Xuereb, A. Ferraro, G. De Chiara, N. Kiesel, [27] L. Latmiral, F. Armata, M. G. Genoni, I. Pikovski, M. S. Kim, M. Paternostro, New J. Phys. 17, 035016 (2015), Phys. Rev. A 93, 052306 (2016). [23] C. K. Law, Phys. Rev. A 51, 2537 (1995). [28] S.Khorasani, Applied Sciences, 7, 656 (2017). [24] D.F. James, J. Jerke, Canadian Journal of Physics 85, 625