Optomechanical Circuits for Nanomechanical Continuous Variable Quantum State Processing

Optomechanical circuits for nanomechanical continuous variable quantum state processing

Michael Schmidt Max Ludwig Friedrich-Alexander-Universität Erlangen-Nürnberg, Staudtstr. 7, D-91058 Erlangen, Germany Florian Marquardt Friedrich-Alexander-Universität Erlangen-Nürnberg, Staudtstr. 7, D-91058 Erlangen, Germany Max Planck Institute for the Science of Light, Günther-Scharowsky-Straße 1/Bau 24, D-91058 Erlangen, Germany

Abstract

We propose and analyze a nanomechanical architecture where light is used to perform linear quantum operations on a set of many vibrational modes. Suitable amplitude modulation of a single laser beam is shown to generate squeezing, entanglement, and state-transfer between modes that are selected according to their mechanical oscillation frequency. Current optomechanical devices based on photonic crystals, as well as other systems with sufficient control over multiple mechanical modes, may provide a platform for realizing this scheme.

pacs:

42.50.Wk, 07.10.Cm, 03.67.Bg, 03.67.Lx

1 Introduction

The field of cavity optomechanics studies the interaction between light and mechanical motion, with promising prospects in fundamental tests of quantum physics, ultrasensitive detection, and applications in quantum information processing (see [1, 2, 3] for reviews). One particularly promising platform consists of "optomechanical crystals", in which strongly localized optical and vibrational modes are implemented in a photonic crystal structure [4]. So far, several interesting possibilities have been pointed out that would make use of multi-mode setups designed on this basis. For example, suitably engineered setups may coherently convert phonons to photons [5] and collective nonlinear dynamics might be observed in optomechanical arrays [6]. Moreover, optomechanical systems in general have been demonstrated to furnish the basic ingredients for writing quantum information from the light field into the long-lived mechanical modes [7, 8, 9]. The recent success in ground state laser-cooling [10, 11] has now opened the door to coherent quantum dynamics in optomechanical systems.

In this paper, we propose a general scheme for continuous-variable quantum state processing [12] utilizing the vibrational modes of such structures. We will show how entanglement and state transfer operations can be applied selectively to pairs of modes, by suitable intensity modulation of a single incoming laser beam. We will discuss limitations for entanglement generation and transfer fidelity, and show how to pick suitable designs to address these challenges.

2 Model

We will first restrict our attention to a single optical mode coupled to many mechanical modes, such that the following standard optomechanical Hamiltonian describes the photon field , the phonons of different localized vibrational modes ( ), their mutual coupling, and the laser drive:

Schematic setup figure illustrating an optomechanical  crystal with localized vibrational modes, coupled by a common optical  mode. The optical mode is driven by an amplitude-modulated laser beam  to engineer frequency-selective entanglement, squeezing and state-transfer  operations of vibrations (see main text).
Figure 1: Schematic setup figure illustrating an optomechanical crystal with localized vibrational modes, coupled by a common optical mode. The optical mode is driven by an amplitude-modulated laser beam to engineer frequency-selective entanglement, squeezing and state-transfer operations of vibrations (see main text).

We are working in a frame rotating at the laser frequency with detuning . Here, denotes the coupling to the drive, which is proportional to the laser amplitude. We omitted to explicitly write down the coupling to the photon and phonon baths, with damping rates and , respectively, although these will be included in our treatment. The bare (single-photon) coupling constants depend on the overlap between the optical and mechanical mode functions and are on the order of , where is the mechanical zero-point amplitude. In photonic crystals the effective cavity length reaches down to wavelength dimensions (see Figure1 for the illustration of a setup).

After application of the standard procedure of splitting off the coherent optical amplitude induced by the laser, , and omitting terms quadratic in (valid for strong drive) to the Hamiltonian, we recover the linearized optomechanical coupling,

Here the dressed couplings can be tuned via the laser intensity, i.e. the circulating photon number: ( without loss of generality). We can now eliminate the driven cavity field (noting that is in the ground state) by second-order perturbation theory: At large detuning the energy scales for phonons and photons separate, and we retain a fully coherent, light-induced mechanical coupling between all the vibrational modes,

where is the mechanical displacement in units of . For this to be valid we have to fulfill , which prevents unwanted transitions. Eq. (3) may be viewed as a "collective optical spring" effect, coupling all the mechanical displacements. The couplings can be changed in-situ and in a time-dependent manner via the laser intensity or the detuning. In the numerical simulations we take and assume all modes to have equal masses. This feature will be crucial for our approach described below. Note that if multiple optical modes are driven, the corresponding coupling constants will add.

In general, the couplings in Eq. (3) will induce quantum state transfer between mutually resonant mechanical modes, and entanglement at low temperatures (usually with the help of optomechanical laser cooling). The generation of entanglement between two mechanical modes based on the bilinear interaction, Eq. (3), has been analyzed in [13, 14, 15, 16, 17, 18, 19] or in [20, 21, 22, 19] for entanglement between the light field and the phononic mode. However, these schemes are not easily scalable to many mechanical modes, mainly because it is not possible to address the modes that are to be entangled. Moreover, these schemes produce an entangled steady-state that is sensitive to thermal fluctuations. We will describe a different scheme that allows mode selective entanglement and is particularly suited for muti-mode setups. Its robustness against thermal influences is enhanced, since it employs parametric instabilities for entanglement generation.

3 General scheme

In contrast to the schemes mentioned above, we have in mind a multi-mode situation for continuous variable quantum information processing and are interested in an efficient approach to selectively couple arbitrary pairs of modes, both for entanglement and state transfer. There are several desiderata to address for a suitable optomechanical architecture of that style: one should be able to (i) switch couplings in time; (ii) easily select pairs for operations; (iii)  get by with only one laser (or a limited number); (iv) achieve large enough operation speeds to beat decoherence; (v) scale to a reasonably large number of modes.

3.1 Frequency selective operations

Static couplings as in Eq. (3) could be used for selective pairwise operations if one were able to shift locally the mechanical frequency to bring the two respective modes into resonance. In principle, this is achievable via the optical spring effect, but would require local addressing with independent laser beams. This could prove challenging in a micron-scale photonic crystal, severely hampering scalability.

Instead, we propose to employ frequency-selective operations, by modulating the laser intensity (and thus ) in time. Entanglement generation by parametric driving has been analyzed recently in various contexts, including superconducting circuits [23], trapped ions [24] , general studies of entanglement in harmonic oscillators [25, 26, 27], optomechanical state transfer and entanglement between the motion of a trapped atom and a mechanical oscillator [28] and entanglement between mechanical and radiation modes [29]. Parametric driving can also lead to mechanical squeezing in optomechanical systems [30].

Let us now consider two modes with coupling . Note that for the purposes of our discussion we set for sake of simplicity. The results mentioned below, however, remain valid for the general case of unequal couplings, which is also used for the numerical simulations. The time-dependent coupling is achieved by modulating the laser intensity at a frequency of the order of the mechanical frequencies. For the circulating photon number follows adiabatically , and we have . The resulting time-dependent light-induced mechanical coupling can be broken down into several contributions, whose relative importance will be determined by the drive frequency . The static terms, , shift the oscillator frequencies by , and give rise to an off-resonant coupling that is ineffective for , but gains influence for . In a realistic setup this interaction might be enhanced due to intrinsically present phonon tunneling between distinct vibrational modes, which could easily be included into the analysis, since it is of the same structure. Moreover the static terms contain single mode squeezing terms, that are always off-resonant and of negligible influence . For the oscillating terms

there are three important cases. A mechanical beam-splitter (state-transfer) interaction is selected for a laser drive modulation frequency . In the interaction picture with respect to and , the resonant part of the full Hamiltonian reads

In contrast, for , we obtain from equation (4) a two-mode squeezing (non-degenerate parametric amplifier) Hamiltonian,

which can lead to efficient entanglement between the modes. Finally, selects the squeezing interaction for a given mode,

out of equation (4). These laser-tunable, frequency-selective mechanical interactions are the basic ingredients for the architecture that we will develop and analyze here. Furthermore, combinations of these Hamiltonians can be constructed, when the laser intensity is modulated with multiple frequencies. Note that one has to keep in mind that the modulation generates also a time dependent radiation pressure force, , that leads to a coherent driving of each of the mechanical modes for . For the parameters we choose here, these processes are off-resonant, though.

Schematic diagram of the resonance regions  for squeezing and entanglement operations between vibrational modes  (
Figure 2: Schematic diagram of the resonance regions for squeezing and entanglement operations between vibrational modes ( , and ), induced by harmonically modulating the laser intensity at frequency . For too weak laser intensity (small ), dissipation prevents parametric resonances (dashed lines indicate resonance regions without dissipation). With increasing light-induced coupling dissipation is overcome and regions of parametric resonance for entangling and squeezing operations emerge. By adjusting the modulation frequency , these operations can be addressed selectively. When is further increased, different regions start to overlap and selectivity is lost. At large J unwanted swapping processes occur, leading to an off-resonant coupling between the modes.

3.2 Limitations

Selective entanglement between two  out of three vibrational modes, produced by an amplitude-modulated  laser beam driving a common optical mode, as sketched in Figure
Figure 3: Selective entanglement between two out of three vibrational modes, produced by an amplitude-modulated laser beam driving a common optical mode, as sketched in Figure 2. All modes are coupled to a thermal bath and assumed to be cooled to their ground state initially. (a,b) Entanglement measure (the logarithmic negativity) between modes 1 and 3, as a function of the driving strength and the modulation frequency , at time . [Contour lines: entanglement between 1&2 or 2&3 larger than 0.5, blue arrows at , red line: resonance position as derived in rotating wave approximation]. In (a) the vibrational modes are spaced densely , leading to overlapping resonance regions (as in Figure2). (b) Larger frequency spacing increases the selectivity significantly [ ]. (c) Dependence on driving strength, i.e. cut along dashed red lines in (a,b), see main text. (d) Dependence on temperature and mechanical quality factor . (e) Same as (d) but starting from thermal equilibrium. [Parameters: , (a,b,c), (b,d,e), (d,e), (a-e)]

We now address the constraining factors for the operation fidelity, both for the two-mode and the multi-mode case. At higher drive powers (as needed for fast operations), the frequency-time uncertainty implies that the different processes need not to be resonant exactly, with an allowable spread . The parametric instabilities occur for . Once these intervals start to overlap, process selectivity is lost and the fidelity suffers. At low operation speeds quantum dissipation and thermal fluctuations will limit the fidelity. This is the essential problem faced by a multi-mode setup, and we will discuss possible remedies further below. The schematic situation for the example of three modes is illustrated in Figure2.

In order to analyze decoherence and dissipation, we employ a Lindblad master equation to evolve the joint state of the mechanical modes. The evolution of any expectation value can be derived from the master equation and is governed by:

Here describes the vibrational modes and already contains the effective interaction (3), with modulated time-dependent couplings. are the damping rates of the vibrational modes, and their equilibrium occupations at the bulk temperature. Note that we effectively added the light-induced decoherence rate to the intrinsic rate [31]. is suppressed by a factor and can be arbitrarily small for larger detuning (at the expense of higher photon number to keep the same ). In the numerical simulation we use and as independent parameters. Note, however, that the sign of the detuning affects the sign of The dissipative term in the first line describes damping (spontaneous and induced emission), and the second line refers to absorption of thermal fluctuations. The relaxation superoperators are defined by (in contrast to the equation for ). For the quadratic Hamiltonian studied here, the equations for the correlators remain closed and are sufficient to describe the Gaussian states produced in the evolution.

We evaluate the logarithmic negativity

as a measure of entanglement for any two given modes ( and ), where is the state of these two modes, and the partial transpose acts on only. For Gaussian states, can be calculated via the symplectic eigenvalues of the position-momentum covariance matrix [32].

In Figure3 we show the simulation results for entangling two out of three vibrational modes. The entanglement saturates at later times, while the phonon number grows exponentially. We plot the results at the fixed time , since there the logarithmic negativity has already saturated. One clearly sees the features predicted above (Figure 2), i.e. the unwanted overlap between entanglement processes at higher driving strengths (Figure3a,b). Increasing the vibrational frequency spacing suppresses these unwanted effects (Figure3b). In Figure3c the threshold for entanglement generation at finite temperature is evident, as is the loss of entanglement at large . Finally, Figure3d,e shows the dependence on temperature and mechanical quality factor indicating that this scheme should be feasible for realistic experimental parameters (see below).

3.3 Larger arrays

Having evenly spaced mechanical frequencies is impossible, because the state transfers between adjacent modes would all be addressed at the same modulation frequency. Any simple layout that offers selectivity and avoids resonance overlap seems to require a frequency interval that grows exponentially with the number of modes. Hence, another approach is needed for large .

The scheme (Figure4) that solves this challenge involves an auxiliary mode at , removed in frequency from the array of evenly spaced "memory" modes in . All operations will take place between a selected memory mode and the auxiliary mode. Then, the state transfer resonances are in the band , and entanglement is addressed within . To make this work, one needs to fulfill the mild constraint where the upper limit prevents unintended driving of the modes caused by the modulated radiation pressure force (see above). Starting with an arbitrary multimode state, state transfer between two memory modes is performed in three steps (swapping , and ), as is entanglement (swap , entangle and swap ). Note that this overhead does not grow with the number of memory modes. Figure5 shows the transfer of a squeezed state from the auxiliary mode to a memory mode.

Several such arrays could be connected in a 2D scheme by linking their auxiliaries with an optical mode (Figure4c). The spectrum of the auxiliaries can be chosen as in the introductory example (Figure 2), which is allowed by the above constraint and enables selective operations between the auxiliaries. State transfer between memory modes of distinct arrays is possible in five steps (swap , swap swap , swap , swap ) as well as entanglement (swap , swap , entangle , swap , swap ). Note that the transfer from the memory modes to the auxiliaries can be done in parallel, hence the scheme can effectively be completed in three steps and there is no time delay compared to the single array with auxiliary. Moreover one might employ more sophisticated transfer schemes [33, 34, 35], to improve the transfer fidelities. Two blocks can be connected by introducing a higher order auxiliary that couples to both auxiliary arrays. State transfer can then be performed between auxiliary modes from different blocks. In principle many blocks can be connected, when their auxiliary modes are stringed together in a chain via higher order auxiliaries.

(a) Layout for an optomechanical array with  an auxiliary vibrational mode, circumventing the problem of frequency  crowding (see main text). (b) Corresponding mechanical frequency spectrum.  The laser modulation frequency has to lie within certain intervals  to select entanglement or state-transfer operations between the
Figure 4: (a) Layout for an optomechanical array with an auxiliary vibrational mode, circumventing the problem of frequency crowding (see main text). (b) Corresponding mechanical frequency spectrum. The laser modulation frequency has to lie within certain intervals to select entanglement or state-transfer operations between the "memory modes" and the auxiliary mode (arrows). (c) Extension to 2D block consisting of three arrays with operations "around the corner".

4 Implementation

Regarding the experimental implementation, in principle any optomechanical system with long-lived mechanical modes can be used. One promising platform are "optomechanical crystals", as introduced by Painter et al. [4] that feature vibrational defect cavities in the GHz regime with experimentally accessible [36]. These would be very well suited for the scheme presented here, due to their design flexibility, particularly of two dimensional structures, and the all-integrated approach, as well as the very large optomechanical coupling strength. Given the currently achieved coupling strength [11, 36] , a detuning of and around cavity photons (reached in recent experiments), we can estimate the induced coupling to approach the damping rate, . This corresponds to the threshold for coherent operations, provided one were to cool down the bath to . This is in principle possible (at ), but will likely run into practical difficulties due to the re-heating of the structure via spurious photon absorption or other effects. At finite bath temperatures corresponding to a thermal occupation , the light intensity must be increased by a factor towards . Initially, the vibrational ground state would be prepared via laser-cooling, as demonstrated in [11].

 Transfer of a squeezed state (
Figure 5: Transfer of a squeezed state ( ) from the auxiliary to the second memory mode (all in their ground state due to prior side-band cooling). (a) Dependence of the transfer fidelity on driving strength (negative J due to red detuning ) and pulse time (swap pulse ideally at ). For long pulse times dissipation hampers the transfer, for large driving strengths resonances overlap, medium times are optimal. (b) Wigner density of the initial squeezed state in the auxiliary mode, and (c) after the transfer. Plotted Wigner density maximizes the transfer fidelity for fixed . (d) Maximization of transfer fidelity over parameter ranges used in (a). [Parameters: , , , , , (b,c)]

In these devices, localized vibrational and optical modes can be produced at engineered defects in a periodic array of holes cut into a free-standing substrate. Adjacent optical and vibrational modes are coupled via tunneling. The typical photon tunnel coupling for modes spaced apart by several lattice constants is [6] in the range of several . Thus, hybridized optical modes will form, one of which can be selected via the laser driving frequency while the others remain idle. The vibrational modes' frequencies can be different or equal, in which case delocalized hybridized mechanical modes are produced. A 'snowflake' crystal made of connected triangles (honeycomb lattice) supports wave guides (line defects) and localized defect modes with optomechanical interaction [5]. Placing point defects (heavier triangles/thicker bridges) in the structure, a tight binding analysis indicates that the mechanical frequency spectrum (Figure4b) can be generated.

Utilizing the phononic modes of an optomechanical system for processing continuous variable quantum information has a number of desirable aspects in comparison to other systems like optical modes or cold atomic vapors. First, the phononic modes can be integrated on a chip and the devices are thus naturally scalable. The Gaussian operations discussed above can be applied easily and the decoherence times are already reasonable, although admittedly worse than for cold atomic vapors.

We briefly mention another option for improving the fidelity: Optimal control techniques [25] could be employed to numerically optimize the pulse shape .

Finally, an essential ingredient will be read-out. We have pointed out [37] how to produce a quantum-non-demolition read-out of the mechanical quadratures in an optomechanical setup. A laser beam (detuning ) is amplitude-modulated at the mechanical frequency of one of the modes. The reflected light carries information only about one quadrature . Its phase is selected by the phase of the amplitude-modulation, while the measurement back-action perturbs solely the other quadrature. Different modes can be read out simultaneously, and the covariance matrix may thus be obtained in repeated experimental runs. Taking measurement statistics for continuously varied quadrature phases would also allow to do full quantum-state tomography, and thereby ultimately process tomography. Alternatively, short pulses may be used for readout (and manipulation) [38].

5 Conclusions

The scheme described here would enable coherent scalable nanomechanical state processing in optomechanical arrays. It can form the basis for generating arbitrary entangled mechanical Gaussian multi-mode states like continuous variable cluster states [39]. An interesting application would be to investigate the decoherence of such states due to the correlated quantum noise acting on the nanomechanical modes. Moreover, recent experiments have shown in principle how arbitrary states can be written from the light field into the mechanics [7, 8, 9]. These could then be manipulated by the interactions described here. Alternatively, for very strong coupling , non-Gaussian mechanical states [40, 41, 42] could be produced, and the induced nonlinear interactions (see e.g. [43, 44]) could potentially open the door to universal quantum computation with continuous variables [12] in these systems.

We acknowledge the ITN Cavity Quantum Optomechanics, an ERC Starting Grant, the DFG Emmy-Noether program and DARPA ORCHID for funding.

References

References

  • [1] T. J. Kippenberg and K. J. Vahala. Cavity optomechanics: Back-action at the mesoscale. Science, 321(5893):1172–1176, 2008.
  • [2] Ivan Favero and Khaled Karrai. Optomechanics of deformable optical cavities. Nature Photonics, 3:201, 2009.
  • [3] Florian Marquardt and Steve Girvin. Optomechanics. Physics, 2:40, 2009.
  • [4] Matt Eichenfield, Jasper Chan, Ryan M. Camacho, Kerry J. Vahala, and Oskar Painter. Optomechanical crystals. Nature, 462:78–82, 2009.
  • [5] Amir H Safavi-Naeini and Oskar Painter. Proposal for an optomechanical traveling wave phonon-photon translator. New J. Phys., 13:013017, 2011.
  • [6] Georg Heinrich, Max Ludwig, Jiang Qian, Björn Kubala, and Florian Marquardt. Collective dynamics in optomechanical arrays. Phys. Rev. Lett., 107:043603, 2011.
  • [7] Amir H. Safavi-Naeini, Thiago P. Mayer Alegre, Jasper Chan, Matt Eichenfield, Martin Winger, Qiang Lin, Jeffrey T. Hill, Darrick Chang, and Oskar Painter. Electromagnetically induced transparency and slow light with optomechanics. Nature, 472:69, 2011.
  • [8] Victor Fiore, Yong Yang, Mark C. Kuzyk, Russell Barbour, Lin Tian, and Hailin Wang. Storing optical information as a mechanical excitation in a silica optomechanical resonator. Phys. Rev. Lett., 107:133601, 2011.
  • [9] E. Verhagen, S. Deléglise, S. Weis, A. Schliesser, and T. J. Kippenberg. Quantum-coherent coupling of a mechanical oscillator to an optical cavity modes. Nature, 482:63–67, 2012.
  • [10] J. D. Teufel, T. Donner, Dali Li, J. W. Harlow, M. S. Allman, K. Cicak, A. J. Sirois, J. D. Whittaker, K. W. Lehnert, and R. W. Simmonds. Sideband cooling of micromechanical motian to the quantum ground state. Nature, 475:359–363, 2011.
  • [11] Jasper Chan, T. P. Mayer Alegre, Amir H. Safavi-Naeini, Jeff T. Hill, Alex Krause, Simon Groeblacher, Markus Aspelmeyer, and Oskar Painter. Laser cooling of a nanomechanical oscillator into its quantum ground state. Nature, 478:89, 2011.
  • [12] Samuel L. Braunstein and Peter van Loock. Quantum information with continuous variables. Rev. Mod. Phys., 77(2):513–577, 2005.
  • [13] M. Pinard, A. Dantan, D. Vitali, O. Arcizet, T. Briant, and A. Heidmann. Entangling movable mirrors in a double-cavity system. EPL, 72(5):747, 2005.
  • [14] Helge Müller-Ebhardt, Henning Rehbein, Roman Schnabel, Karsten Danzmann, and Yanbei Chen. Entanglement of macroscopic test masses and the standard quantum limit in laser interferometry. Phys. Rev. Lett., 100(1):013601, 2008.
  • [15] M. Bhattacharya and P. Meystre. Multiple membrane cavity optomechanics. Phys. Rev. A, 78(4):041801, 2008.
  • [16] Michael J. Hartmann and Martin B. Plenio. Steady state entanglement in the mechanical vibrations of two dielectric membranes. Phys. Rev. Lett., 101(20):200503, 2008.
  • [17] K. Hammerer, M. Wallquist, C. Genes, M. Ludwig, F. Marquardt, P. Treutlein, P. Zoller, J. Ye, and H. J. Kimble. Strong coupling of a mechanical oscillator and a single atom. Phys. Rev. Lett., 103(6):063005, 2009.
  • [18] Max Ludwig, K. Hammerer, and Florian Marquardt. Entanglement of mechanical oscillators coupled to a nonequilibrium environments. Phys. Rev. A, 82(1):012333, 2010.
  • [19] Uzma Akram and G. J. Milburn. Photon phonon entanglement in coupled optomechanical arrays. arXiv:1109.0790, 2011.
  • [20] D. Vitali, S. Gigan, A. Ferreira, H. R. Böhm, P. Tombesi, A. Guerreiro, V. Vedral, A. Zeilinger, and M. Aspelmeyer. Optomechanical entanglement between a movable mirror and a cavity field. Phys. Rev. Lett., 98(3):030405, 2007.
  • [21] M. Paternostro, D. Vitali, S. Gigan, M. S. Kim, C. Brukner, J. Eisert, and M. Aspelmeyer. Creating and probing multipartite macroscopic entanglement with light. Phys. Rev. Lett., 99(25):250401, 2007.
  • [22] M. Abdi, Sh. Barzanjeh, P. Tombesi, and D. Vitali. Effect of phase noise on the generation of stationary entanglement in cavity optomechanics. Phys. Rev. A, 84:032325, 2011.
  • [23] L Tian, M S Allman, and R W Simmonds. Parametric coupling between macroscopic quantum resonators. New J. Phys., 10(11):115001, 2008.
  • [24] Alessio Serafini, Alex Retzker, and Martin B Plenio. Manipulating the quantum information of the radial modes of trapped ions: linear phononics, entanglement generation, quantum state transmission and non-locality tests. New J. Phys., 11(2):023007, 2009.
  • [25] Fernando Galve and Eric Lutz. Energy cost and optimal entanglement production in harmonic chains. Phys. Rev. A, 79(3):032327, 2009.
  • [26] V. M. Bastidas, J. H. Reina, C. Emary, and T. Brandes. Entanglement and parametric resonance in driven quantum systems. Phys. Rev. A, 81(1):012316, 2010.
  • [27] Fernando Galve, Leonardo A. Pachón, and David Zueco. Bringing entanglement to the high temperature limit. Phys. Rev. Lett., 105(18):180501, 2010.
  • [28] M. Wallquist, K. Hammerer, P. Zoller, C. Genes, M. Ludwig, F. Marquardt, P. Treutlein, J. Ye, and H. J. Kimble. Single-atom cavity qed and optomicromechanics. Phys. Rev. A, 81(2):023816, 2010.
  • [29] A Mari and J Eisert. Opto- and electro-mechanical entanglement improved by modulation. New Journal of Physics, 14(7):075014, 2012.
  • [30] A. Mari and J. Eisert. Gently modulating optomechanical systems. Phys. Rev. Lett., 103(21):213603, 2009.
  • [31] F. Marquardt, J. P. Chen, A. A. Clerk, and S. M. Girvin. Quantum theory of cavity-assisted sideband cooling of mechanical motion. Phys. Rev. Lett., 99(9):093902, 2007.
  • [32] G. Vidal and R. F. Werner. Computable measure of entanglement. Phys. Rev. A, 65(3):032314, 2002.
  • [33] K. Bergmann, H. Theuer, and B. W. Shore. Coherent population transfer among quantum states of atoms and molecules. Rev. Mod. Phys., 70(3):1003–1025, 1998.
  • [34] Ying-Dan Wang and Aashish A. Clerk. Using interference for high fidelity quantum state transfer in optomechanics. Phys. Rev. Lett., 108:153603, 2012.
  • [35] Lin Tian. Adiabatic state conversion and pulse transmission in optomechanical systems. Phys. Rev. Lett., 108:153604, 2012.
  • [36] Jasper Chan, Amir H. Safavi-Naeini, Jeff T. Hill, Sean Meenehan, and Oskar Painter. Optimized optomechanical crystal cavity with acoustic radiation shield. Applied Physics Letters, 101(8):081115, 2012.
  • [37] A. A. Clerk, F. Marquardt, and K. Jacobs. Back-action evasion and squeezing of a mechanical resonator using a cavity detector. New J. Phys., 10(9):095010, 2008.
  • [38] M. R. Vanner, I. Pikovski, G. D. Cole, M. S. Kim, C. Brukner, K. Hammerer, G. J. Milburn, and M. Aspelmeyer. Pulsed quantum optomechanics. Proc Natl Acad Sci USA, 108:16182, 2011.
  • [39] Peter van Loock, Christian Weedbrook, and Mile Gu. Building gaussian cluster states by linear optics. Phys. Rev. A, 76:032321, 2007.
  • [40] Max Ludwig, Björn Kubala, and Florian Marquardt. The optomechanical instability in the quantum regime. New Journal of Physics, 10(9):095013+, 2008.
  • [41] A. Nunnenkamp, K. Børkje, and S. M. Girvin. Single-photon optomechanics. Phys. Rev. Lett., 107:063602, 2011.
  • [42] Jiang Qian, A. A. Clerk, K. Hammerer, and Florian Marquardt. Quantum signatures of the optomechanical instability. arXiv:1112.6200, 2011.
  • [43] Max Ludwig, Amir H. Safavi-Naeini, Oskar Painter, and Florian Marquardt. Optomechanical photon detection and enhanced dispersive phonon readout. arXiv:1202.0532, 2012.
  • [44] K. Stannigel, P. Komar, S. J. M. Habraken, S. D. Bennett, M. D. Lukin, P. Zoller, and P. Rabl. Optomechanical quantum information processing with photons and phonons. Phys. Rev. Lett., 109:013603, 2012.

whitacreestion.blogspot.com

Source: https://www.arxiv-vanity.com/papers/1202.3659/

0 Response to "Optomechanical Circuits for Nanomechanical Continuous Variable Quantum State Processing"

Post a Comment

Iklan Atas Artikel

Iklan Tengah Artikel 1

Iklan Tengah Artikel 2

Iklan Bawah Artikel