26.12.2014 Views

Slowing and stopping light using an optomechanical crystal array

Slowing and stopping light using an optomechanical crystal array

Slowing and stopping light using an optomechanical crystal array

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Home Search Collections Journals About Contact us My IOPscience<br />

<strong>Slowing</strong> <strong><strong>an</strong>d</strong> <strong>stopping</strong> <strong>light</strong> <strong>using</strong> <strong>an</strong> optomech<strong>an</strong>ical <strong>crystal</strong> <strong>array</strong><br />

This article has been downloaded from IOPscience. Please scroll down to see the full text article.<br />

2011 New J. Phys. 13 023003<br />

(http://iopscience.iop.org/1367-2630/13/2/023003)<br />

View the table of contents for this issue, or go to the journal homepage for more<br />

Download details:<br />

IP Address: 131.215.237.47<br />

The article was downloaded on 02/02/2011 at 05:07<br />

Please note that terms <strong><strong>an</strong>d</strong> conditions apply.


New Journal of Physics<br />

The open–access journal for physics<br />

<strong>Slowing</strong> <strong><strong>an</strong>d</strong> <strong>stopping</strong> <strong>light</strong> <strong>using</strong> <strong>an</strong> optomech<strong>an</strong>ical<br />

<strong>crystal</strong> <strong>array</strong><br />

D E Ch<strong>an</strong>g 1,4 , A H Safavi-Naeini 2,4 , M Hafezi 3 <strong><strong>an</strong>d</strong> O Painter 2,5<br />

1 Institute for Qu<strong>an</strong>tum Information <strong><strong>an</strong>d</strong> Center for the Physics of Information,<br />

California Institute of Technology, Pasadena, CA 91125, USA<br />

2 Thomas J Watson, Sr, Laboratory of Applied Physics, California Institute of<br />

Technology, Pasadena, CA 91125, USA<br />

3 Joint Qu<strong>an</strong>tum Institute <strong><strong>an</strong>d</strong> Department of Physics, University of Maryl<strong><strong>an</strong>d</strong>,<br />

College Park, MD 20742, USA<br />

E-mail: opainter@caltech.edu<br />

New Journal of Physics 13 (2011) 023003 (26pp)<br />

Received 24 November 2010<br />

Published 1 February 2011<br />

Online at http://www.njp.org/<br />

doi:10.1088/1367-2630/13/2/023003<br />

Abstract. One of the major adv<strong>an</strong>ces needed to realize all-optical information<br />

processing of <strong>light</strong> is the ability to delay or coherently store <strong><strong>an</strong>d</strong> retrieve optical<br />

information in a rapidly tunable m<strong>an</strong>ner. In the classical domain, this optical<br />

buffering is expected to be a key ingredient of m<strong>an</strong>aging the flow of information<br />

over complex optical networks. Such a system also has profound implications for<br />

qu<strong>an</strong>tum information processing, serving as a long-term memory that c<strong>an</strong> store<br />

the full qu<strong>an</strong>tum information contained in <strong>an</strong> optical pulse. Here, we suggest a<br />

novel approach to <strong>light</strong> storage involving <strong>an</strong> optical waveguide coupled to <strong>an</strong><br />

optomech<strong>an</strong>ical <strong>crystal</strong> <strong>array</strong>, where <strong>light</strong> in the waveguide c<strong>an</strong> be dynamically<br />

<strong><strong>an</strong>d</strong> coherently tr<strong>an</strong>sferred into long-lived mech<strong>an</strong>ical vibrations of the <strong>array</strong>.<br />

Under realistic conditions, this system is capable of achieving large b<strong><strong>an</strong>d</strong>widths<br />

<strong><strong>an</strong>d</strong> storage/delay times in a compact, on-chip platform.<br />

4 These authors contributed equally to this work.<br />

5 Author to whom <strong>an</strong>y correspondence should be addressed.<br />

New Journal of Physics 13 (2011) 023003<br />

1367-2630/11/023003+26$33.00 © IOP Publishing Ltd <strong><strong>an</strong>d</strong> Deutsche Physikalische Gesellschaft


2<br />

Contents<br />

1. Introduction 2<br />

2. Description of the system: <strong>an</strong> optomech<strong>an</strong>ical <strong>crystal</strong> (OMC) <strong>array</strong> 3<br />

3. <strong>Slowing</strong> <strong><strong>an</strong>d</strong> <strong>stopping</strong> <strong>light</strong> 5<br />

3.1. Static regime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

3.2. Storage of optical pulse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

3.3. Imperfections in storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

4. OMC design 10<br />

5. Outlook 13<br />

Acknowledgments 13<br />

Appendix A. Equations of motion for <strong>an</strong> OMC <strong>array</strong> 13<br />

Appendix B. Tr<strong>an</strong>sfer matrix <strong>an</strong>alysis of propagation 15<br />

Appendix C. Optical noise power 16<br />

Appendix D. The b<strong><strong>an</strong>d</strong> structure <strong>an</strong>alysis 17<br />

Appendix E. Implementation in <strong>an</strong> OMC 20<br />

References 24<br />

1. Introduction<br />

Light is a natural c<strong><strong>an</strong>d</strong>idate for tr<strong>an</strong>smitting information across large networks owing to its high<br />

speed <strong><strong>an</strong>d</strong> low propagation losses. A major obstacle to building more adv<strong>an</strong>ced optical networks<br />

is the lack of <strong>an</strong> all-optically controlled device that c<strong>an</strong> robustly delay or store optical wave<br />

packets over a tunable amount of time. In the classical domain, such a device would enable alloptical<br />

buffering <strong><strong>an</strong>d</strong> switching, bypassing the need to convert <strong>an</strong> optical pulse to <strong>an</strong> electronic<br />

signal. In the qu<strong>an</strong>tum realm, such a device could serve as a memory to store the full qu<strong>an</strong>tum<br />

information contained in a <strong>light</strong> pulse until it c<strong>an</strong> be passed to a processing node at some<br />

later time.<br />

A number of schemes to coherently delay <strong><strong>an</strong>d</strong> store optical information are being<br />

actively explored. These r<strong>an</strong>ge from tunable coupled resonator optical waveguide (CROW)<br />

structures [1, 2], where the propagation of <strong>light</strong> is dynamically altered by modulating the<br />

refractive index of the system, to electromagnetically induced tr<strong>an</strong>sparency (EIT) in atomic<br />

media [3, 4], where the optical pulse is reversibly mapped into internal atomic degrees<br />

of freedom. While these schemes have been demonstrated in a number of remarkable<br />

experiments [5]–[8], they remain difficult to implement in a practical setting. Here, we present<br />

a novel approach to store or stop <strong>an</strong> optical pulse propagating through a waveguide, wherein<br />

coupling between the waveguide <strong><strong>an</strong>d</strong> a nearby n<strong>an</strong>omech<strong>an</strong>ical resonator <strong>array</strong> enables one to<br />

map the optical field into long-lived mech<strong>an</strong>ical excitations. This process is completely qu<strong>an</strong>tum<br />

coherent <strong><strong>an</strong>d</strong> allows the delay <strong><strong>an</strong>d</strong> release of pulses to be rapidly <strong><strong>an</strong>d</strong> all-optically tuned. Our<br />

scheme combines m<strong>an</strong>y of the best attributes of previously proposed approaches, in that it<br />

simult<strong>an</strong>eously allows for large b<strong><strong>an</strong>d</strong>widths of operation, on-chip integration, relatively long<br />

delay/storage times <strong><strong>an</strong>d</strong> ease of external control. Beyond <strong>light</strong> storage, this work opens up the<br />

intriguing possibility of a platform for qu<strong>an</strong>tum or classical all-optical information processing<br />

<strong>using</strong> mech<strong>an</strong>ical systems.<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


3<br />

2. Description of the system: <strong>an</strong> optomech<strong>an</strong>ical <strong>crystal</strong> (OMC) <strong>array</strong><br />

An optomech<strong>an</strong>ical <strong>crystal</strong> (OMC) [9] is a periodic structure that comprises both a photonic [10]<br />

<strong><strong>an</strong>d</strong> a phononic [11] <strong>crystal</strong>. The ability to engineer optical <strong><strong>an</strong>d</strong> mech<strong>an</strong>ical properties in the<br />

same structure should enable unprecedented control over <strong>light</strong>–matter interactions. Pl<strong>an</strong>ar twodimensional<br />

(2D) photonic <strong>crystal</strong>s, formed from patterned thin dielectric films on the surface of<br />

a microchip, have been succesfully employed as n<strong>an</strong>oscale optical circuits capable of efficiently<br />

routing, diffracting <strong><strong>an</strong>d</strong> trapping <strong>light</strong>. Fabrication techniques for such 2D photonic <strong>crystal</strong>s have<br />

matured signific<strong>an</strong>tly over the last decade, with experiments on <strong>an</strong> Si chip [12] demonstrating<br />

excellent optical tr<strong>an</strong>smission through long (N > 100) linear <strong>array</strong>s of coupled photonic <strong>crystal</strong><br />

cavities. In a similar Si chip platform, it has recently been shown that suitably designed photonic<br />

<strong>crystal</strong> cavities also contain localized acoustic reson<strong>an</strong>ces that are strongly coupled to the<br />

optical field via radiation pressure [9]. These pl<strong>an</strong>ar OMCs are thus a natural c<strong><strong>an</strong>d</strong>idate for<br />

the implementation of our proposed slow-<strong>light</strong> scheme.<br />

In the following, we consider <strong>an</strong> OMC containing a periodic <strong>array</strong> of such defect cavities<br />

(see figures 1(a) <strong><strong>an</strong>d</strong> (b)). Each element of the <strong>array</strong> contains two optical cavity modes (denoted<br />

1, 2) <strong><strong>an</strong>d</strong> a co-localized mech<strong>an</strong>ical reson<strong>an</strong>ce. The Hamiltoni<strong>an</strong> describing the dynamics of a<br />

single element is of the form<br />

˜H om = ¯hω 1 â † 1â1 + ¯hω 2 â † 2â2 + ¯hω m ˆb † ˆb + ¯hh( ˆb + ˆb † )(â † 1â2 + â † 2â1). (1)<br />

Here, ω 1,2 are the reson<strong>an</strong>ce frequencies of the two optical modes, ω m is the mech<strong>an</strong>ical<br />

reson<strong>an</strong>ce frequency <strong><strong>an</strong>d</strong> â 1 , â 2 , ˆb are <strong>an</strong>nihilation operators for these modes. The<br />

optomech<strong>an</strong>ical interaction cross-couples the cavity modes 1 <strong><strong>an</strong>d</strong> 2 with a strength characterized<br />

by h <strong><strong>an</strong>d</strong> that depends linearly on the mech<strong>an</strong>ical displacement ˆx ∝ ( ˆb + ˆb † ). While we formally<br />

treat â 1 , â 2 , ˆb as qu<strong>an</strong>tum mech<strong>an</strong>ical operators, for the most part it also suffices to treat these<br />

terms as dimensionless classical qu<strong>an</strong>tities describing the positive-frequency components of the<br />

optical fields <strong><strong>an</strong>d</strong> mech<strong>an</strong>ical position. In addition to the optomech<strong>an</strong>ical interaction described<br />

by equation (1), cavity modes 1 are coupled to a common two-way waveguide (described<br />

below). Each element is decoupled from the others except through the waveguide.<br />

The design considerations necessary for achieving such a system are discussed in detail in<br />

the section ‘Optomech<strong>an</strong>ical <strong>crystal</strong> design’. For now, we take as typical parameters ω 1 /2π =<br />

200 THz, ω m /2π = 10 GHz, h/2π = 0.35 MHz <strong><strong>an</strong>d</strong> mech<strong>an</strong>ical <strong><strong>an</strong>d</strong> (unloaded) optical quality<br />

factors of Q m ≡ ω m /γ m ∼ 10 3 (room temperature)–10 5 (low temperature) <strong><strong>an</strong>d</strong> Q 1 ≡ ω 1 /κ 1,in =<br />

3 × 10 6 , where γ m is the mech<strong>an</strong>ical decay rate <strong><strong>an</strong>d</strong> κ 1,in is the intrinsic optical cavity decay<br />

rate. Similar parameters have been experimentally observed in other OMC systems [9, 13]. In<br />

practice, one c<strong>an</strong> also over-couple cavity mode 1 to the waveguide, with a waveguide-induced<br />

optical decay rate κ ex that is much larger th<strong>an</strong> κ in .<br />

For the purpose of slowing <strong>light</strong>, cavity modes 2 will be reson<strong>an</strong>tly driven by <strong>an</strong> external<br />

laser, so that to good approximation â 2 ≈ α 2 (t)e −iω2t c<strong>an</strong> be replaced by its me<strong>an</strong>-field value. We<br />

furthermore consider the case where the frequencies are tuned such that ω 1 = ω 2 + ω m . Keeping<br />

only the reson<strong>an</strong>t terms in the optomech<strong>an</strong>ical interaction, we arrive at a simplified Hamiltoni<strong>an</strong><br />

for a single <strong>array</strong> element (see figure 1(b)),<br />

H om = ¯hω 1 â † 1â1 + ¯hω m ˆb † ˆb + ¯h m (t)(â † 1 ˆb e −i(ω 1−ω m )t + h.c.). (2)<br />

Here, we have defined <strong>an</strong> effective optomech<strong>an</strong>ical driving amplitude m (t) = hα 2 (t) <strong><strong>an</strong>d</strong><br />

assume that α 2 (t) is real. Mode 2 thus serves as a ‘tuning’ cavity that mediates population<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


4<br />

(a)<br />

(b)<br />

a R (z)<br />

optical<br />

waveguide<br />

a R (z+d)<br />

a R (z)<br />

optical<br />

waveguide<br />

a R (z+d)<br />

a L (z)<br />

in<br />

ex<br />

a 1<br />

h<br />

b<br />

m<br />

a 2<br />

ex<br />

in<br />

a L (z+d)<br />

optical<br />

cavities<br />

mech<strong>an</strong>ical<br />

cavity<br />

a L (z)<br />

in<br />

ex<br />

a<br />

b<br />

m (t)<br />

a L (z+d)<br />

“active”<br />

optical<br />

cavity<br />

mech<strong>an</strong>ical<br />

cavity<br />

m<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

(c)<br />

2<br />

|t| , Ω = 0<br />

2<br />

|r| , Ω = 0<br />

2<br />

|t| , Ω = κ/10<br />

2<br />

|r| , Ω = κ/10<br />

0<br />

-0.1 0 0.1<br />

0<br />

-3 -2 -1 0 1 2 3<br />

δ/κ<br />

1<br />

0.5<br />

(d)<br />

a R (z,t)<br />

|n , n<br />

m 1 ><br />

|n , n +<br />

m 1<br />

1><br />

m (t)<br />

|n + 1, n<br />

m 1><br />

Figure 1. (a) Illustration of a double optical cavity system forming the unit<br />

cell of the optomech<strong>an</strong>ical <strong>array</strong>. A two-way optical waveguide is coupled to<br />

a pair of optical cavity modes a 1 <strong><strong>an</strong>d</strong> a 2 , whose reson<strong>an</strong>ce frequencies differ<br />

by the frequency of the mech<strong>an</strong>ical mode b. Both optical modes leak energy<br />

into the waveguide at a rate κ ex <strong><strong>an</strong>d</strong> have <strong>an</strong> inherent decay rate κ in . The<br />

mech<strong>an</strong>ical resonator optomech<strong>an</strong>ically couples the two optical reson<strong>an</strong>ces with<br />

a cross-coupling rate of h. (b) A simplified system diagram where the classically<br />

driven cavity mode a 2 is effectively eliminated to yield <strong>an</strong> optomech<strong>an</strong>ical<br />

driving amplitude m between the mech<strong>an</strong>ical mode <strong><strong>an</strong>d</strong> the cavity mode a 1 .<br />

(c) Frequency-dependent reflect<strong>an</strong>ce (black curve) <strong><strong>an</strong>d</strong> tr<strong>an</strong>smitt<strong>an</strong>ce (red) of<br />

a single <strong>array</strong> element, in the case of no optomech<strong>an</strong>ical driving amplitude<br />

m = 0 (dotted line) <strong><strong>an</strong>d</strong> <strong>an</strong> amplitude of m = κ ex /10 (solid line). The inherent<br />

cavity decay is chosen to be κ in = 0.1κ ex . (Inset) The optomech<strong>an</strong>ical coupling<br />

creates a tr<strong>an</strong>sparency window of width ∼4 2 m /κ ex for a single element <strong><strong>an</strong>d</strong><br />

enables perfect tr<strong>an</strong>smission on reson<strong>an</strong>ce, δ k = 0. (d) Energy level structure<br />

of the simplified system. The number of photons <strong><strong>an</strong>d</strong> phonons are denoted by<br />

n 1 <strong><strong>an</strong>d</strong> n m , respectively. The optomech<strong>an</strong>ical driving amplitude m couples<br />

states |n m + 1, n 1 〉 ↔ |n m , n 1 + 1〉, while the <strong>light</strong> in the waveguide couples states<br />

|n m , n 1 〉 ↔ |n m , n 1 + 1〉. The two couplings create a set of -type tr<strong>an</strong>sitions<br />

<strong>an</strong>alogous to that in EIT.<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


5<br />

tr<strong>an</strong>sfer (Rabi oscillations) between the ‘active’ cavity mode 1 <strong><strong>an</strong>d</strong> the mech<strong>an</strong>ical resonator<br />

at a controllable rate (t), which is the key mech<strong>an</strong>ism for our stopped-<strong>light</strong> protocol. In<br />

the following <strong>an</strong>alysis, we will focus exclusively on the active cavity mode <strong><strong>an</strong>d</strong> drop the ‘1’<br />

subscript.<br />

A Hamiltoni<strong>an</strong> of the form (2) also describes <strong>an</strong> optomech<strong>an</strong>ical system with a<br />

single optical mode, when the cavity is driven off reson<strong>an</strong>ce at frequency ω 1 − ω m <strong><strong>an</strong>d</strong> â 1<br />

corresponds to the sideb<strong><strong>an</strong>d</strong>s generated at frequencies ±ω m around the classical driving field.<br />

For a single system, this Hamiltoni<strong>an</strong> leads to efficient optical cooling of the mech<strong>an</strong>ical<br />

motion [14, 15], a technique being used to cool n<strong>an</strong>omech<strong>an</strong>ical systems toward their qu<strong>an</strong>tum<br />

ground states [16]–[19]. While the majority of such work focuses on how optical fields<br />

affect the mech<strong>an</strong>ical dynamics, here we show that the optomech<strong>an</strong>ical interaction strongly<br />

modifies optical field propagation to yield the slow/stopped <strong>light</strong> phenomenon. Equation (2)<br />

is quite general <strong><strong>an</strong>d</strong> thus this phenomenon could, in principle, be observed in <strong>an</strong>y <strong>array</strong> of<br />

optomech<strong>an</strong>ical systems coupled to a waveguide. In practice, there are several considerations<br />

that make the 2D OMC ‘ideal’. Firstly, our system exhibits <strong>an</strong> extremely large optomech<strong>an</strong>ical<br />

coupling h <strong><strong>an</strong>d</strong> contains a second optical tuning cavity that c<strong>an</strong> be driven reson<strong>an</strong>tly, which<br />

enables large driving amplitudes m <strong>using</strong> reasonable input power [20]. Using two different<br />

cavities also potentially allows for greater versatility <strong><strong>an</strong>d</strong> addressability of our system. For<br />

inst<strong>an</strong>ce, in our proposed design the photons in cavity 1 are spatially filtered from those in cavity<br />

2 [20]. Secondly, the 2D OMC is <strong>an</strong> easily scalable <strong><strong>an</strong>d</strong> compact platform. Finally, as described<br />

below, the high mech<strong>an</strong>ical frequency of our device compared to typical optomech<strong>an</strong>ical<br />

systems allows for a good bal<strong>an</strong>ce between long storage times <strong><strong>an</strong>d</strong> suppression of noise<br />

processes.<br />

3. <strong>Slowing</strong> <strong><strong>an</strong>d</strong> <strong>stopping</strong> <strong>light</strong><br />

3.1. Static regime<br />

We first <strong>an</strong>alyze propagation in the waveguide when m (t) = m is static during the tr<strong>an</strong>sit<br />

interval of the signal pulse. As shown in the appendix, the evolution equations in a rotating<br />

frame for a single element located at position z j along the waveguide are given by<br />

dâ<br />

dt = −κ 2 â + i m ˆb + i<br />

√<br />

cκex<br />

2 (â R,in(z j ) + â L,in (z j )) + √ cκ in â N (z j ), (3)<br />

dˆb<br />

dt = −γ m<br />

2 ˆb + i m â + ˆF N (t). (4)<br />

Equation (3) is a st<strong><strong>an</strong>d</strong>ard input relation characterizing the coupling of right- (â R,in ) <strong><strong>an</strong>d</strong> leftpropagating<br />

(â L,in ) optical input fields in the waveguide with the cavity mode. Here κ = κ ex + κ in<br />

is the total optical cavity decay rate, â N (z) is qu<strong>an</strong>tum noise associated with the inherent optical<br />

cavity loss, <strong><strong>an</strong>d</strong> for simplicity we have assumed a linear dispersion relation ω k = c|k| in the<br />

waveguide. Equation (4) describes the optically driven mech<strong>an</strong>ical motion, which decays at a<br />

rate γ m <strong><strong>an</strong>d</strong> is subject to thermal noise ˆF N (t). The cavity mode couples to the right-propagating<br />

field through the equation<br />

( 1 ∂<br />

c ∂t + ∂ ) √<br />

κex<br />

â R (z, t) = i<br />

∂z<br />

2c δ(z − z j)â + ik 0 â R , (5)<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


6<br />

where k 0 = ω 1 /c. We solve the above equations to find the reflection <strong><strong>an</strong>d</strong> tr<strong>an</strong>smission<br />

coefficients r, t of a single element for a right-propagating incoming field of frequency ω k (see<br />

the appendix). In the limit where γ m = 0, <strong><strong>an</strong>d</strong> defining δ k ≡ ω k − ω 1 ,<br />

δ k κ ex<br />

r(δ k ) = −<br />

, (6)<br />

δ k (−2iδ k + κ) + 2i 2 m<br />

while t = 1 + r. Example reflect<strong>an</strong>ce <strong><strong>an</strong>d</strong> tr<strong>an</strong>smitt<strong>an</strong>ce curves are plotted in figure 1(c). For<br />

<strong>an</strong>y non-zero m , a single element is perfectly tr<strong>an</strong>smitting on reson<strong>an</strong>ce, whereas for m = 0<br />

reson<strong>an</strong>t tr<strong>an</strong>smission past the cavity is blocked. When m ≠ 0, excitation of the cavity mode is<br />

inhibited through destructive interference between the incoming field <strong><strong>an</strong>d</strong> the optomech<strong>an</strong>ical<br />

coupling. In EIT, a similar effect occurs via interference between two electronic tr<strong>an</strong>sitions. This<br />

<strong>an</strong>alogy is further elucidated by considering the level structure of our optomech<strong>an</strong>ical system<br />

(figure 1(d)), where the interference pathways <strong><strong>an</strong>d</strong> the ‘’-type tr<strong>an</strong>sition reminiscent of EIT<br />

are clearly visible. The interference is accomp<strong>an</strong>ied by a steep phase variation in the tr<strong>an</strong>smitted<br />

field around reson<strong>an</strong>ce, which c<strong>an</strong> result in a slow group velocity. These steep features <strong><strong>an</strong>d</strong><br />

their similarity to EIT in a single optomech<strong>an</strong>ical system have been theoretically [21, 22] <strong><strong>an</strong>d</strong><br />

experimentally studied [23, 24], while interference effects between a single cavity mode <strong><strong>an</strong>d</strong><br />

two mech<strong>an</strong>ical modes have also been observed [25].<br />

From r, t for a single element, the propagation characteristics through <strong>an</strong> infinite <strong>array</strong><br />

(figure 2(a)) c<strong>an</strong> be readily obtained via b<strong><strong>an</strong>d</strong> structure calculations [10]. To maximize the<br />

propagation b<strong><strong>an</strong>d</strong>width of the system, we choose the spacing d between elements such that<br />

k 0 d = (2n + 1)π/2 where n is a non-negative integer. With this choice of phasing, the reflections<br />

from multiple elements destructively interfere under optomech<strong>an</strong>ical driving. Typical b<strong><strong>an</strong>d</strong><br />

structures are illustrated in figures 2(b)–(f). The color coding of the dispersion curves (red for<br />

waveguide, green for optical cavity <strong><strong>an</strong>d</strong> blue for mech<strong>an</strong>ical reson<strong>an</strong>ce) indicates the distribution<br />

of energy or fractional occupation in the various degrees of freedom of the system in steady<br />

state. Far away from the cavity reson<strong>an</strong>ce, the dispersion relation is nearly linear <strong><strong>an</strong>d</strong> simply<br />

reflects the character of the input optical waveguide, while the propagation is strongly modified<br />

near reson<strong>an</strong>ce (ω = ω 1 = ω 2 + ω m ). In the absence of optomech<strong>an</strong>ical coupling ( m = 0), a<br />

tr<strong>an</strong>smission b<strong><strong>an</strong>d</strong> gap of width ∼κ forms around the optical cavity reson<strong>an</strong>ce (reflections from<br />

the bare optical cavity elements constructively interfere). In the presence of optomech<strong>an</strong>ical<br />

driving, the b<strong><strong>an</strong>d</strong> gap splits in two (blue shaded regions) <strong><strong>an</strong>d</strong> a new propagation b<strong><strong>an</strong>d</strong> centered<br />

on the cavity reson<strong>an</strong>ce appears in the middle of the b<strong><strong>an</strong>d</strong> gap. For weak driving ( m κ), the<br />

width of this b<strong><strong>an</strong>d</strong> is ∼4 2 m /κ, whereas for strong driving ( m κ), one recovers the ‘normal<br />

mode splitting’ of width ∼2 m [26]. This relatively flat polaritonic b<strong><strong>an</strong>d</strong> yields the slow-<strong>light</strong><br />

propagation of interest. Indeed, for small m , the steady-state energy in this b<strong><strong>an</strong>d</strong> is almost<br />

completely mech<strong>an</strong>ical in character, indicating the strong mixing <strong><strong>an</strong>d</strong> conversion of energy in<br />

the waveguide to mech<strong>an</strong>ical excitations along the <strong>array</strong>.<br />

It c<strong>an</strong> be shown that the Bloch wavevector near reson<strong>an</strong>ce is given by (see the appendix)<br />

k eff ≈ k 0 + κ exδ k<br />

+ iκ exκ in δk<br />

2<br />

2d 2 m<br />

4d 4 m<br />

+ (2κ3 ex − 3κ exκ 2 in + 12κ ex 2 m )δ3 k<br />

24d 6 m<br />

. (7)<br />

The group velocity on reson<strong>an</strong>ce, v g = (dk eff /dδ k ) −1 | δk =0 = 2d 2 m /κ ex, c<strong>an</strong> be dramatically<br />

slowed by <strong>an</strong> amount that is tunable through the optomech<strong>an</strong>ical coupling strength m . The<br />

quadratic <strong><strong>an</strong>d</strong> cubic terms in k eff characterize pulse absorption <strong><strong>an</strong>d</strong> group velocity dispersion,<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


7<br />

(a)<br />

d<br />

(b)<br />

0.8<br />

Ω m<br />

= 0 Ω m<br />

= κ/4 Ω m<br />

= κ/2 Ω m<br />

= κ<br />

0.7<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

a j-1<br />

a j<br />

a j+1<br />

ex<br />

in<br />

m (t)<br />

0.3<br />

b j-1 b j+1<br />

0.2<br />

0 0.5 1<br />

m<br />

κ = ω/4<br />

Ω = κ/10<br />

ω (π c / a)<br />

0.6<br />

0.5<br />

0.4<br />

0 0.5 1 0 0.5 1 0 0.5 1<br />

K (π/d)<br />

(c) (d) (e) (f )<br />

0.60<br />

0.58<br />

0.56<br />

0.54<br />

2<br />

4Ω<br />

~ m<br />

κ<br />

ω (π c / a)<br />

0.6<br />

0.5<br />

0.4<br />

0.52<br />

0.50<br />

0.48<br />

~ κ<br />

0.3<br />

0.46<br />

0.2<br />

0.44<br />

0.1<br />

0.42<br />

0<br />

0 0.5 1<br />

K (π/d)<br />

0 0.5 1<br />

fraction of excitations<br />

0.400<br />

0 0.2 0.4 0.6 0.8 1<br />

K (π/d)<br />

0 0.2 0.4 0.6 0.8 1<br />

fraction of excitations<br />

Figure 2. (a) Illustration of <strong>an</strong> OMC <strong>array</strong>. A two-way optical waveguide is<br />

coupled to a periodic <strong>array</strong> of optomech<strong>an</strong>ical elements spaced by a dist<strong>an</strong>ce d.<br />

The optical cavity modes a j of each element leak energy into the waveguide<br />

at a rate κ ex <strong><strong>an</strong>d</strong> have <strong>an</strong> inherent decay rate κ in . The mech<strong>an</strong>ical resonator of<br />

each element has frequency ω m <strong><strong>an</strong>d</strong> is optomech<strong>an</strong>ically coupled to the cavity<br />

mode through a tuning cavity (shown in figure 1) with strength m . (b) The b<strong><strong>an</strong>d</strong><br />

structure of the system, for a r<strong>an</strong>ge of driving strengths between m = 0 <strong><strong>an</strong>d</strong><br />

m = κ. The blue shaded regions indicate b<strong><strong>an</strong>d</strong> gaps, while the color of the b<strong><strong>an</strong>d</strong>s<br />

elucidates the fractional occupation (red for energy in the optical waveguide,<br />

green for the optical cavity <strong><strong>an</strong>d</strong> blue for mech<strong>an</strong>ical excitations). The dynamic<br />

compression of the b<strong><strong>an</strong>d</strong>width is clearly visible as m → 0. (c) B<strong><strong>an</strong>d</strong> structure<br />

for the case m = κ/10 is shown in greater detail. (d) The fractional occupation<br />

for each b<strong><strong>an</strong>d</strong> in (c) is plotted separately. It c<strong>an</strong> be seen that the polaritonic<br />

slow-<strong>light</strong> b<strong><strong>an</strong>d</strong> is mostly mech<strong>an</strong>ical in nature, with a small mixing with the<br />

waveguide modes <strong><strong>an</strong>d</strong> negligible mixing with the optical cavity mode. Zoom-ins<br />

of p<strong>an</strong>els (c) <strong><strong>an</strong>d</strong> (d) are shown in (e) <strong><strong>an</strong>d</strong> (f).<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


8<br />

respectively. In the relev<strong>an</strong>t regime where κ ex ≫ κ in , κ ex m , these effects are negligible<br />

within a b<strong><strong>an</strong>d</strong>width ω ∼ min( 2√ 2<br />

√ 2 m<br />

Nκex κ in<br />

, 2(6π)1/3 2 m<br />

κ ex<br />

). The second term is the b<strong><strong>an</strong>d</strong>width over<br />

N 1/3<br />

which certain frequency components of the pulse acquire a π-phase shift relative to others,<br />

leading to pulse distortion. This yields a b<strong><strong>an</strong>d</strong>width–delay product of<br />

ωτ delay ∼ min( √ 2Nκ ex /κ in , (6π N 2 ) 1/3 ) (8)<br />

for static m <strong><strong>an</strong>d</strong> negligible mech<strong>an</strong>ical losses. When intrinsic optical cavity losses are<br />

negligible, <strong><strong>an</strong>d</strong> if one is not concerned with pulse distortion, <strong>light</strong> c<strong>an</strong> propagate over the full<br />

b<strong><strong>an</strong>d</strong>width ∼4 2 m<br />

/κ of the slow-<strong>light</strong> polariton b<strong><strong>an</strong>d</strong> <strong><strong>an</strong>d</strong> the b<strong><strong>an</strong>d</strong>width–delay product increases<br />

to ωτ delay ∼ N (see the appendix). On the other h<strong><strong>an</strong>d</strong>, we note that if we had operated in a<br />

regime where k 0 d = πn, constructive interference in reflection would limit the b<strong><strong>an</strong>d</strong>width–delay<br />

product to ωτ delay ∼ 1, independent of system size.<br />

In the static regime, the b<strong><strong>an</strong>d</strong>width–delay product obtained here is <strong>an</strong>alogous to CROW<br />

systems [2]. In the case of EIT, a static b<strong><strong>an</strong>d</strong>width–delay product of ωτ delay ∼ √ OD results,<br />

where OD is the optical depth of the atomic medium. This product is limited by photon<br />

absorption<br />

√<br />

<strong><strong>an</strong>d</strong> rescattering into other directions, <strong><strong>an</strong>d</strong> is <strong>an</strong>alogous to our result ωτ delay ∼<br />

Nκex /κ in in the case of large intrinsic cavity linewidth. On the other h<strong><strong>an</strong>d</strong>, when κ in is<br />

negligible, photons are never lost <strong><strong>an</strong>d</strong> reflections c<strong>an</strong> be suppressed by interference. This yields<br />

<strong>an</strong> improved scaling ωτ delay ∼ N 2/3 or ∼N, depending on whether one is concerned with<br />

group velocity dispersion. In atomic media, the weak atom–photon coupling makes achieving<br />

OD > 100 very challenging [27]. In contrast, in our system as few as N ∼ 10 elements would<br />

be equivalently dense.<br />

3.2. Storage of optical pulse<br />

We now show that the group velocity v g (t) = 2d 2 m (t)/κ ex c<strong>an</strong> in fact be adiabatically ch<strong>an</strong>ged<br />

once a pulse is completely localized inside the system, leading to distortion-less propagation at<br />

a dynamically tunable speed. In particular, by tuning v g (t) → 0, the pulse c<strong>an</strong> be completely<br />

stopped <strong><strong>an</strong>d</strong> stored.<br />

This phenomenon c<strong>an</strong> be understood in terms of the static b<strong><strong>an</strong>d</strong> structure of the system<br />

(figure 2) <strong><strong>an</strong>d</strong> a ‘dynamic compression’ of the pulse b<strong><strong>an</strong>d</strong>width. The same physics applies for<br />

CROW structures [1, 28], <strong><strong>an</strong>d</strong> the argument is re-summarized here. First, under const<strong>an</strong>t m , <strong>an</strong><br />

optical pulse within the b<strong><strong>an</strong>d</strong>width of the polariton b<strong><strong>an</strong>d</strong> completely enters the medium. Once<br />

the pulse is inside, we consider the effect of a gradual reduction in m (t). Decomposing the<br />

pulse into Bloch wavevector components, it is clear that each Bloch wavevector is conserved<br />

under arbitrary ch<strong>an</strong>ges of m , as it is fixed by the system periodicity. Furthermore, tr<strong>an</strong>sitions<br />

to other b<strong><strong>an</strong>d</strong>s are negligible provided that the energy levels are varied adiabatically compared<br />

to the size of the gap, which tr<strong>an</strong>slates into <strong>an</strong> adiabatic condition |(d/dt)( 2 m /κ)| κ2 . Then,<br />

conservation of the Bloch wavevector implies that the b<strong><strong>an</strong>d</strong>width of the pulse is dynamically<br />

compressed, <strong><strong>an</strong>d</strong> the reduction in slope of the polariton b<strong><strong>an</strong>d</strong> (figure 2) causes the pulse to<br />

propagate at <strong>an</strong> inst<strong>an</strong>t<strong>an</strong>eous group velocity v g (t) without <strong>an</strong>y distortion. In the limit that<br />

m → 0, the polaritonic b<strong><strong>an</strong>d</strong> becomes flat <strong><strong>an</strong>d</strong> completely mech<strong>an</strong>ical in character, indicating<br />

that the pulse has been reversibly <strong><strong>an</strong>d</strong> coherently mapped onto stationary mech<strong>an</strong>ical excitations<br />

within the <strong>array</strong>. We note that since m is itself set by the tuning cavities, its rate of ch<strong>an</strong>ge<br />

c<strong>an</strong>not exceed the optical linewidth <strong><strong>an</strong>d</strong> thus the adiabaticity condition is always satisfied in the<br />

weak driving regime.<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


9<br />

The maximum storage time is set by the mech<strong>an</strong>ical decay rate, ∼1/γ m . For realistic system<br />

parameters, ω m /2π = 10 GHz <strong><strong>an</strong>d</strong> Q m = 10 5 , this yields a storage time of ∼10 µs. In CROW<br />

structures, <strong>light</strong> is stored as circulating fields in optical n<strong>an</strong>o-cavities, where state-of-the-art<br />

quality factors of Q ∼ 10 6 limit the storage time to ∼1 ns. The key feature of our system is that<br />

we effectively ‘down-convert’ the high-frequency optical fields to low-frequency mech<strong>an</strong>ical<br />

excitations, which naturally decay over much longer time scales. While storage times of ∼10 ms<br />

are possible <strong>using</strong> atomic media [29], their b<strong><strong>an</strong>d</strong>widths so far have been limited to


10<br />

heating is removed through the optical ch<strong>an</strong>nel. For high temperatures, the thermal noise scales<br />

inversely with ω m , <strong><strong>an</strong>d</strong> the use of high-frequency mech<strong>an</strong>ical oscillators ensures that the noise<br />

remains easily tolerable even at room temperature.<br />

Thermal noise in the high-frequency oscillator c<strong>an</strong> essentially be eliminated in cryogenic<br />

environments, which then enables faithful storage of single photons. Intuitively, a single-photon<br />

pulse c<strong>an</strong> be stored for a period only as long as the mech<strong>an</strong>ical decay time ∼γm<br />

−1 <strong><strong>an</strong>d</strong> as long<br />

as a noise-induced mech<strong>an</strong>ical excitation is unlikely to be generated over a region covering the<br />

pulse length <strong><strong>an</strong>d</strong> over the tr<strong>an</strong>sit time τ delay . The latter condition is equivalent to the statement<br />

that the power P ph ∼ ¯hω 1 ω in the single-photon pulse exceeds P noise . While we have focused<br />

on the static regime so far, when thermal heating is negligible, realizing P ph /P noise 1 in the<br />

static case in fact ensures that the inequality holds even when Ɣ opt (t) is time varying. Physically,<br />

the rate of Stokes scattering scales linearly with Ɣ opt while the group velocity scales inversely,<br />

<strong><strong>an</strong>d</strong> thus the probability of a noise excitation being added on top of the single-photon pulse is<br />

fixed over a given tr<strong>an</strong>sit length.<br />

In a realistic setting, the optomech<strong>an</strong>ical driving amplitude m itself will be coupled<br />

to the bath temperature, as absorption of the pump photons in the tuning cavities leads to<br />

material heating. To underst<strong><strong>an</strong>d</strong> the limitations as a qu<strong>an</strong>tum memory, we have numerically<br />

optimized the static b<strong><strong>an</strong>d</strong>width–delay product ωτ delay for a train of single-photon pulses,<br />

subject to the constraints ω < min(2 √ 2 2 m /√ Nκ ex κ in , 2(6π) 1/3 2 m /(κ exN 1/3 )), P ph /P noise ><br />

1 <strong><strong>an</strong>d</strong> γ m τ delay < 1. As a realistic model for the bath temperature, we take T b = T 0 + χα2 2 =<br />

T 0 + χ( m /h) 2 , where T 0 is the base temperature <strong><strong>an</strong>d</strong> χ ∼ 2 µK is a temperature coefficient<br />

that describes heating due to pump absorption (see the appendix). Using T 0 = 100 mK <strong><strong>an</strong>d</strong><br />

Q m = 10 5 , we find (ωτ delay ) max ∼ 110, which is achieved for parameter values N ∼ 275,<br />

κ ex /2π ∼ 1.1 GHz <strong><strong>an</strong>d</strong> m /2π ∼ 130 MHz.<br />

4. OMC design<br />

A schematic diagram showing a few periods of our proposed 2D OMC slow-<strong>light</strong> structure is<br />

given in figure 3. The structure is built around a ‘snowflake’ <strong>crystal</strong> pattern of etched holes into a<br />

silicon slab [20]. This pattern, when implemented with a physical lattice const<strong>an</strong>t of a = 400 nm,<br />

snowflake radius r = 168 nm <strong><strong>an</strong>d</strong> snowflake width w = 60 nm (see figure 3(a)), provides a<br />

simult<strong>an</strong>eous phononic b<strong><strong>an</strong>d</strong> gap from 8.6 to 12.6 GHz <strong><strong>an</strong>d</strong> a photonic pseudo-b<strong><strong>an</strong>d</strong> gap from<br />

180 to 230 THz (see the appendix). Owing to its unique b<strong><strong>an</strong>d</strong> gap properties, the snowflake<br />

patterning c<strong>an</strong> be used to form waveguides <strong><strong>an</strong>d</strong> reson<strong>an</strong>t cavities for both acoustic <strong><strong>an</strong>d</strong> optical<br />

waves simply by removing regions of the pattern. For inst<strong>an</strong>ce, a single point defect, formed<br />

by removing two adjacent holes (a so-called ‘L2’ defect), yields the co-localized phononic<br />

<strong><strong>an</strong>d</strong> photonic reson<strong>an</strong>ces shown in figures 3(b) <strong><strong>an</strong>d</strong> (c), respectively. The radiation pressure,<br />

or optomech<strong>an</strong>ical coupling between the two reson<strong>an</strong>ces, c<strong>an</strong> be qu<strong>an</strong>tified by a coupling rate,<br />

g, which corresponds to the frequency shift in the optical reson<strong>an</strong>ce line introduced by a single<br />

phonon of the mech<strong>an</strong>ical reson<strong>an</strong>ce. Numerical finite-element method (FEM) simulations of<br />

the L2 defect indicate that the mech<strong>an</strong>ical reson<strong>an</strong>ce occurs at ω m /2π = 11.2 GHz, with a<br />

coupling rate of g/2π = 489 kHz to the optical mode at frequency ω o /2π = 199 THz (freespace<br />

optical wavelength of λ 0 ≈ 1500 nm).<br />

To form the double-cavity system described in the slow-<strong>light</strong> scheme above, a pair of L2<br />

cavities are placed in the near field of each other as shown in the dashed box region of figure 3(a).<br />

Modes of the two degenerate L2 cavities mix, forming supermodes of the double-cavity system<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


11<br />

(a)<br />

240<br />

230<br />

(d)<br />

199.000<br />

(e) (f )<br />

σ = +1<br />

y<br />

σ = -1<br />

y<br />

220<br />

198.990<br />

(b)<br />

ν (THz)<br />

210<br />

200<br />

198.980<br />

198.970<br />

ν m<br />

(c)<br />

190<br />

180<br />

0 0.5<br />

1 0<br />

K (π/d)<br />

0.5 1<br />

y<br />

x<br />

d<br />

r 170<br />

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5<br />

w<br />

a<br />

K (π/a)<br />

Figure 3. (a) Top view of the proposed OMC <strong>array</strong>, with the superlattice unit<br />

cell of length d high<strong>light</strong>ed in the center. The unit cell contains two coupled<br />

L2 defect cavities (shaded in gray) with two side-coupled linear defect optical<br />

waveguides (shaded in red). The different envelope functions pertain to the odd<br />

<strong><strong>an</strong>d</strong> even optical cavity supermodes (green solid <strong><strong>an</strong>d</strong> dashed lines, respectively),<br />

the odd mech<strong>an</strong>ical supermode (blue solid line) <strong><strong>an</strong>d</strong> the odd <strong><strong>an</strong>d</strong> even optical<br />

waveguide modes (red solid <strong><strong>an</strong>d</strong> dashed lines). The displacement field amplitude<br />

|Q(r)| of the mech<strong>an</strong>ical mode <strong><strong>an</strong>d</strong> in-pl<strong>an</strong>e electric field amplitude |E(r)| of<br />

the optical mode are shown in (b) <strong><strong>an</strong>d</strong> (c), respectively, for a single L2 defect<br />

cavity. (d) B<strong><strong>an</strong>d</strong>structure of the linear-defect waveguide (gray) <strong><strong>an</strong>d</strong> the zone<br />

folded superlattice of the entire coupled-resonator system (red). The cavity mode<br />

(green) crosses the superlattice b<strong><strong>an</strong>d</strong> at mid-zone, <strong><strong>an</strong>d</strong> the waveguide–cavity<br />

interaction is shown in more detail in insets (e) <strong><strong>an</strong>d</strong> (a) for the even (σ y = +1)<br />

<strong><strong>an</strong>d</strong> odd (σ y = −1) supermodes, respectively.<br />

that are split in frequency. The frequency splitting between modes c<strong>an</strong> be tuned via the number<br />

of snowflake periods between the cavities. As described in more detail in the appendix, it is the<br />

optomech<strong>an</strong>ical cross-coupling of the odd (E − ) <strong><strong>an</strong>d</strong> even (E + ) optical supermodes mediated by<br />

the motion of the odd parity mech<strong>an</strong>ical supermode (Q − ) of the double cavity that drives the<br />

slow-<strong>light</strong> behavior of the system. Since Q − is a displacement field that is <strong>an</strong>tisymmetric about<br />

the two cavities, there is no optomech<strong>an</strong>ical self-coupling between the optical supermodes <strong><strong>an</strong>d</strong><br />

this mech<strong>an</strong>ical mode. On the other h<strong><strong>an</strong>d</strong>, the cross-coupling between the two different parity<br />

optical supermodes is large <strong><strong>an</strong>d</strong> given by h = g/ √ 2 = 2π(346 kHz). By letting â 1 , â 2 <strong><strong>an</strong>d</strong> ˆb be<br />

the <strong>an</strong>nihilation operators for the modes E − , E + <strong><strong>an</strong>d</strong> Q − , we obtain the system Hamiltoni<strong>an</strong> of<br />

equation (1).<br />

The different spatial symmetries of the optical cavity supermodes also allow them to be<br />

addressed independently. To achieve this, we create a pair of linear defects in the snowflake<br />

lattice, as shown in figure 3(a), each acting as a single-mode optical waveguide at the desired<br />

frequency of roughly 200 THz (see figure 3(d)). Sending <strong>light</strong> down both waveguides, with the<br />

individual waveguide modes either in or out of phase with each other, will then excite the even<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


12<br />

Required Input Power (mW)<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

-5 -4 -3 -2 -1 0 1 2 3 4<br />

0<br />

5<br />

(ω − ω )/κ<br />

2 ex<br />

Figure 4. The input power (red line) required to achieve the system parameters<br />

used in the text, i.e. m /2π = 130 MHz with h/2π = 0.346 MHz, <strong><strong>an</strong>d</strong> the<br />

attenuation per unit cell α (solid black line) are shown as a function of detuning<br />

of the pump beam from the pump cavity frequency. The dotted line is the<br />

approximate expression derived for the attenuation, α ≈ κ ex κ in /4δk 2 . The gray<br />

region indicates the b<strong><strong>an</strong>d</strong> gap in which the pump cavities c<strong>an</strong>not be excited from<br />

the waveguide. The trade-off between small pump input powers <strong><strong>an</strong>d</strong> low pump<br />

attenuation factors is readily apparent in this plot.<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

α<br />

or odd supermode of the double cavity, respectively. The waveguide width <strong><strong>an</strong>d</strong> proximity to the<br />

L2 cavities c<strong>an</strong> be used to tune the cavity loading (see the appendix), which for the structure<br />

in figure 3(a) results in the desired κ ex /2π = 2.4 GHz. It should be noted that these line-defect<br />

waveguides do not guide phonons at the frequency of Q − <strong><strong>an</strong>d</strong> thus no additional phonon leakage<br />

is induced in the localized mech<strong>an</strong>ical reson<strong>an</strong>ce.<br />

The full slow-<strong>light</strong> waveguide system consists of a periodic <strong>array</strong> of the double-cavity,<br />

double-waveguide structure. The numerically computed b<strong><strong>an</strong>d</strong> diagram, for spacing d = 15a<br />

periods of the snowflake lattice between cavity elements (the superlattice period), is shown in<br />

figure 3(d). This choice of superlattice period results in the folded superlattice b<strong><strong>an</strong>d</strong> intersecting<br />

the E − (â 1 ) cavity frequency ω 1 at roughly mid-zone, corresponding to the desired inter-cavity<br />

phase shift of kd = π/2. A zoom-in of the b<strong><strong>an</strong>d</strong>structure near the optical cavity reson<strong>an</strong>ces<br />

is shown in figures 3(e) <strong><strong>an</strong>d</strong> (f). In figure 3(e), the even parity supermode b<strong><strong>an</strong>d</strong>structure is<br />

plotted (i.e. assuming that the even supermode of the double waveguide is excited), whereas<br />

in figure 3(f) it is the odd parity supermode b<strong><strong>an</strong>d</strong>structure.<br />

A subtlety in the optical pumping of the periodically <strong>array</strong>ed waveguide system is that for<br />

the E + (â 2 ) optical cavity reson<strong>an</strong>ce at ω 2 , there exists a tr<strong>an</strong>smission b<strong><strong>an</strong>d</strong> gap. To populate<br />

cavity â 2 , then, <strong><strong>an</strong>d</strong> to create the polaritonic b<strong><strong>an</strong>d</strong> at ω 1 , the pump beam must be s<strong>light</strong>ly offreson<strong>an</strong>t<br />

from ω 2 , but still at ω 1 − ω m . We achieve this by choosing a double-cavity separation<br />

(14 periods) resulting in a cavity mode splitting ((ω 1 − ω 2 )/2π = 9.7 GHz) s<strong>light</strong>ly smaller th<strong>an</strong><br />

the mech<strong>an</strong>ical frequency (ω m /2π = 11.2 GHz), as shown in figures 3(e) <strong><strong>an</strong>d</strong> (f). By ch<strong>an</strong>ging<br />

the detuning between ω 1 − ω 2 <strong><strong>an</strong>d</strong> ω m , a trade-off c<strong>an</strong> be made between the attenuation of the<br />

pump beam per unit cell, α ≡ exp(−Im{K }d), <strong><strong>an</strong>d</strong> total required input power, shown in figure 4.<br />

In appendix D.3, we show that the total attenuation per unit cell is given by α ≈ κ ex κ in /4δ 2 k .<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


13<br />

Interestingly, by <strong>using</strong> higher input powers such that α → 0, it is in principle possible to<br />

eliminate completely effects due to absorption in the <strong>array</strong>, which may lead to inhomogeneous<br />

pump photon occupations.<br />

5. Outlook<br />

The possibility of <strong>using</strong> optomech<strong>an</strong>ical systems to facilitate major tasks in classical optical<br />

networks has been suggested in several recent proposals [20, 25, 35]. The present work not<br />

only extends these prospects but proposes a fundamentally new direction where optomech<strong>an</strong>ical<br />

systems c<strong>an</strong> be used to control <strong><strong>an</strong>d</strong> m<strong>an</strong>ipulate <strong>light</strong> at a qu<strong>an</strong>tum mech<strong>an</strong>ical level. Such<br />

efforts would closely mirror the m<strong>an</strong>y proposals to perform similar tasks <strong>using</strong> EIT <strong><strong>an</strong>d</strong> atomic<br />

ensembles [4]. At the same time, the optomech<strong>an</strong>ical <strong>array</strong> has a number of novel features<br />

compared to atoms, in that each element c<strong>an</strong> be deterministically positioned, addressed <strong><strong>an</strong>d</strong><br />

m<strong>an</strong>ipulated, <strong><strong>an</strong>d</strong> a single element is already optically dense. Furthermore, the ability to freely<br />

convert between phonons <strong><strong>an</strong>d</strong> photons enables new possibilities for m<strong>an</strong>ipulating <strong>light</strong> through<br />

the m<strong>an</strong>ipulation of sound. Taken together, this raises the possibility that mech<strong>an</strong>ical systems<br />

c<strong>an</strong> provide a novel, highly configurable on-chip platform for realizing qu<strong>an</strong>tum optics <strong><strong>an</strong>d</strong><br />

‘atomic’ physics.<br />

Acknowledgments<br />

This work was supported by the DARPA/MTO ORCHID program through a gr<strong>an</strong>t from AFOSR.<br />

DC acknowledges support from the NSF <strong><strong>an</strong>d</strong> the Gordon <strong><strong>an</strong>d</strong> Betty Moore Foundation through<br />

Caltech’s Center for the Physics of Information. ASN acknowledges support from NSERC. MH<br />

acknowledges support from the US Army Research Office MURI award W911NF0910406.<br />

Appendix A. Equations of motion for <strong>an</strong> OMC <strong>array</strong><br />

Here we derive the equations of motion for <strong>an</strong> <strong>array</strong> of optomech<strong>an</strong>ical systems coupled to a<br />

two-way waveguide. Because each element in the <strong>array</strong> couples independently to the waveguide,<br />

it suffices here to only consider a single element, from which the result for <strong>an</strong> arbitrary number<br />

of elements is easily generalized.<br />

We model the interaction between the active cavity mode 1 <strong><strong>an</strong>d</strong> the waveguide with the<br />

following Hamiltoni<strong>an</strong>,<br />

H cav-wg =<br />

∫ ∞<br />

−∞<br />

dk ¯h(ck − ω 1 )â † R,kâR,k −<br />

−¯hg √ 2π<br />

∫ ∞<br />

−∞<br />

∫ ∞<br />

−∞<br />

dk ¯h(ck + ω 1 )â † L,kâL,k<br />

dz δ(z − z j )(â † 1 (â R(z) + â L (z)) + h.c.).<br />

(A.1)<br />

Here, â R,k , â L,k are <strong>an</strong>nihilation operators for left- <strong><strong>an</strong>d</strong> right-going waveguide modes of<br />

wavevector k, <strong><strong>an</strong>d</strong> ω 1 is the frequency of the cavity mode. For convenience, we have defined all<br />

optical energies relative to ω 1 , <strong><strong>an</strong>d</strong> assumed that the waveguide has a linear dispersion relation<br />

ω(k) = c|k|. The last term on the right describes a point-like coupling between the cavity (at<br />

position z j ) <strong><strong>an</strong>d</strong> the left- <strong><strong>an</strong>d</strong> right-going waveguide modes, with a strength g. The operator<br />

â R (z) physically describes the <strong>an</strong>nihilation of a right-going photon at position z <strong><strong>an</strong>d</strong> is related to<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


14<br />

∫<br />

the wavevector <strong>an</strong>nihilation operators by â R (z) = √ 1 ∞ dk 2π −∞ eikz â R,k (with a similar definition<br />

for â L (z)). Equation (A.1) resembles a st<strong><strong>an</strong>d</strong>ard Hamiltoni<strong>an</strong> used to formulate qu<strong>an</strong>tum cavity<br />

input–output relations [36], properly generalized to the case when the cavity accepts <strong>an</strong> input<br />

from either direction. Note that we make the approximation that the left- <strong><strong>an</strong>d</strong> right-going<br />

waves c<strong>an</strong> be treated as separate qu<strong>an</strong>tum fields, with modes in each direction running from<br />

−∞ < k < ∞. This allows both the left- <strong><strong>an</strong>d</strong> right-going fields to separately satisfy c<strong>an</strong>onical<br />

field commutation relations, [â R (z), â R (z ′ )] = [â L (z), â L (z ′ )] = δ(z − z ′ ), while commuting<br />

with each other. Thus each field contains some unphysical modes (e.g. wavevector components<br />

k < 0 for the right-going field), but the approximation remains valid as long as there is no<br />

process in the system evolution that allows for the population of such modes.<br />

From the Hamiltoni<strong>an</strong> above, one finds the following Heisenberg equation of motion for<br />

the right-going field,<br />

( 1 ∂<br />

c ∂t + ∂ ) √<br />

2πig<br />

â R (z) = δ(z − z j )â 1 + ik 0 â R ,<br />

(A.2)<br />

∂z<br />

c<br />

where k 0 = ω 1 /c. A similar equation holds for â L . The coupling of the cavity mode to a<br />

continuum of waveguide modes leads to irreversible decay of the cavity at a rate κ ex . Below,<br />

we will show that κ ex is related to the parameters in the Hamiltoni<strong>an</strong> by κ ex = 4πg 2 /c. With this<br />

identification, one recovers equation (5) in the main text.<br />

The Heisenberg equation of motion for the cavity mode is given by<br />

d<br />

dt â1 = ig √ 2π(â R (z j ) + â L (z j )).<br />

(A.3)<br />

To cast this equation into a more useful form, we first integrate the field equation (A.2) across<br />

the discontinuity at z j ,<br />

√<br />

2πig<br />

â R (z + j ) = â R(z − j ) + â 1 ,<br />

c<br />

√<br />

2πig<br />

â L (z − j ) = â L(z + j ) + â 1 .<br />

c<br />

(A.4)<br />

(A.5)<br />

We c<strong>an</strong> define â L,in (z j ) = â L (z + j ) <strong><strong>an</strong>d</strong> â R,in(z j ) = â R (z − j ) as the input fields to the cavity. It then<br />

follows that<br />

d<br />

dt â1 = ig √ 2π ( ) 2πg 2<br />

â R,in + â L,in − â 1 ,<br />

(A.6)<br />

c<br />

<strong><strong>an</strong>d</strong> thus we indeed see that the waveguide induces a cavity decay rate κ ex /2 = 2πg 2 /c. In<br />

the case where the cavity has <strong>an</strong> additional intrinsic decay rate κ in , a similar derivation holds<br />

to connect the intrinsic decay with some corresponding noise input field â in . From these<br />

considerations, <strong><strong>an</strong>d</strong> including the optomech<strong>an</strong>ical coupling, one arrives at equation (3) in the<br />

main text,<br />

dâ 1<br />

= − κ √<br />

dt 2 â + i ˆb<br />

cκex<br />

m + i<br />

(âR,in (z j ) + â L,in (z j ) ) + √ cκ in â N (z j ). (A.7)<br />

2<br />

Finally, we consider the equation of motion for the mech<strong>an</strong>ical mode given by equation (4)<br />

in the main text,<br />

dˆb<br />

dt = −γ m<br />

2 ˆb + i m â 1 + ˆF N (t). (A.8)<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


15<br />

The zero-me<strong>an</strong> noise operator ˆF N must accomp<strong>an</strong>y the decay term in the mech<strong>an</strong>ical evolution<br />

in order to preserve c<strong>an</strong>onical commutation relations of ˆb, ˆb † at all times. In the case where<br />

the decay causes the mech<strong>an</strong>ical motion to return to thermal equilibrium with some reservoir at<br />

temperature T b , the noise operator has a two-time correlation function given by 〈F N (t)F † N (t ′ )〉 =<br />

γ m (¯n + 1)δ(t − t ′ ) [37], where ¯n = (e¯hω m/k B T b<br />

− 1) −1 is the Bose occupation number at the<br />

mech<strong>an</strong>ical frequency ω m .<br />

Appendix B. Tr<strong>an</strong>sfer matrix <strong>an</strong>alysis of propagation<br />

First, we derive the reflection <strong><strong>an</strong>d</strong> tr<strong>an</strong>smission coefficients for a single element in the case of<br />

const<strong>an</strong>t optomech<strong>an</strong>ical driving amplitude m . Given the linearity of the system, it suffices to<br />

treat equations (A.2), (A.7) <strong><strong>an</strong>d</strong> (A.8) as classical equations for this purpose, <strong><strong>an</strong>d</strong> furthermore to<br />

set the noise terms ˆF N = 0, â in = 0. For concreteness, we will consider the case of <strong>an</strong> incident<br />

right-going cw field in the waveguide. Upon interaction with the optomech<strong>an</strong>ical system at z j ,<br />

the total right-going field c<strong>an</strong> be written in the form<br />

a R (z) = e ikz−iδ kt ( (−z + z j ) + t(δ k )(z − z j ) ) ,<br />

(B.1)<br />

while the left-going field is given by a L (z) = e −ikz−iδkt r(δ k )(−z + z j ). Here, (z) is the unit<br />

step function, δ k = ck − ω 1 is the detuning of the input field from the cavity reson<strong>an</strong>ce, <strong><strong>an</strong>d</strong> r, t<br />

are the reflection <strong><strong>an</strong>d</strong> tr<strong>an</strong>smission coefficients for the system. At the same time, we look for<br />

solutions of the cavity field <strong><strong>an</strong>d</strong> mech<strong>an</strong>ical mode of the form a 1 = Ae −iδkt <strong><strong>an</strong>d</strong> b = Be −iδkt . The<br />

coefficients r, t, A, B c<strong>an</strong> be obtained by substituting this <strong>an</strong>satz into equations (A.2), (A.7) <strong><strong>an</strong>d</strong><br />

(A.8). This yields the reflection coefficient given by equation (6) in the main text, while the<br />

tr<strong>an</strong>smission coefficient is related by t = 1 + r.<br />

To calculate propagation through <strong>an</strong> <strong>array</strong> of N elements, it is convenient to introduce a<br />

tr<strong>an</strong>sfer matrix formalism. Specifically, the fields immediately to the right of the optomech<strong>an</strong>ical<br />

element (at z = z + j ) c<strong>an</strong> be related to those immediately to the left (at z = z− j ) in terms of a<br />

tr<strong>an</strong>sfer matrix M om ,<br />

(<br />

aR (z + j )<br />

) (<br />

aR (z − j<br />

a L (z + j ) = M )<br />

)<br />

om<br />

a L (z − j ) , (B.2)<br />

where<br />

M om = 1 t<br />

( )<br />

t 2 − r 2 r<br />

. (B.3)<br />

−r 1<br />

On the other h<strong><strong>an</strong>d</strong>, free propagation in the waveguide is characterized by the matrix M f ,<br />

( ) ( )<br />

aR (z + d) aR (z)<br />

= M f , (B.4)<br />

a L (z + d) a L (z)<br />

where<br />

M f =<br />

( )<br />

e<br />

ikd<br />

0<br />

0 e −ikd . (B.5)<br />

The tr<strong>an</strong>sfer matrix for <strong>an</strong> entire system c<strong>an</strong> then be obtained by successively multiplying<br />

the tr<strong>an</strong>sfer matrices for a single element <strong><strong>an</strong>d</strong> for free propagation together. In particular, the<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


16<br />

tr<strong>an</strong>sfer matrix for a single ‘block’, defined as interaction with a single optomech<strong>an</strong>ical element<br />

followed by free propagation over a dist<strong>an</strong>ce d to the next optomech<strong>an</strong>ical element, is given by<br />

M block = M f M om , <strong><strong>an</strong>d</strong> the propagation over N blocks is simply characterized by M N = Mblock N .<br />

Before studying the propagation through the entire <strong>array</strong>, we first focus on the propagation<br />

past two blocks, M 2 = Mblock 2 . Because we w<strong>an</strong>t our device to be highly tr<strong>an</strong>smitting when the<br />

optomech<strong>an</strong>ical coupling is turned on, we choose the spacing d between consecutive blocks to<br />

be such that k 0 d = π (2n + 1), where n is <strong>an</strong> integer. Physically, this spaces consecutive elements<br />

2<br />

by <strong>an</strong> odd multiple of λ/4, where λ is the reson<strong>an</strong>t wavelength, such that the reflections from<br />

consecutive elements tend to destructively interfere with each other. This c<strong>an</strong> be confirmed by<br />

examining the resulting reflection coefficient for the two-block system,<br />

r 2 ≡ M 2(1, 2)<br />

M 2 (2, 2) = −κ2 ex δ2 k<br />

2 4 m<br />

+ O(δ 3 k<br />

), (B.6)<br />

where M 2 (i, j) denotes matrix elements of M 2 . Note that the reflection coefficient is now<br />

suppressed as a quadratic in the detuning, whereas for a single element r ≈ iκ ex δ k /2 2 m<br />

is linear.<br />

In the above equation, we have made the simplifying approximation that kd ≈ k 0 d, since in<br />

realistic systems the dispersion from free propagation will be negligible compared to that arising<br />

from interaction with <strong>an</strong> optomech<strong>an</strong>ical element.<br />

Now we c<strong>an</strong> consider tr<strong>an</strong>smission past N/2 pairs of two blocks (i.e. N elements in total).<br />

Because the reflection r 2 is quadratic in the detuning, its effect on the total tr<strong>an</strong>smission is<br />

only of the order of δk 4 (because the lowest-order contribution is <strong>an</strong> event where the field is<br />

reflected twice before passing through the system). Thus, up to O(δk 3 ), the total tr<strong>an</strong>smission<br />

coefficient t N is just given by t N ≈ t N/2<br />

2<br />

, where t 2 = 1/M 2 (2, 2) is the tr<strong>an</strong>smission coefficient<br />

for a two-block system. It is convenient to write t N = e ik eff Nd in terms of <strong>an</strong> effective wavevector<br />

k eff , which leads to equation (7) in the main text. Performing a similar <strong>an</strong>alysis for the case<br />

where k 0 d = nπ, where reflections from consecutive elements interfere constructively, one<br />

finds that the b<strong><strong>an</strong>d</strong>width–delay product for the system ωτ delay ∼ 1 does not improve with the<br />

system size.<br />

Appendix C. Optical noise power<br />

In this section, we derive the optical cooling equation given by equation (9) in the main text.<br />

We begin by considering the system Hamiltoni<strong>an</strong> for a single element (equation (1) of the main<br />

text), in the case where the tuning mode 2 is driven on reson<strong>an</strong>ce <strong><strong>an</strong>d</strong> c<strong>an</strong> be approximated by a<br />

classical field, â 2 →α 2 e −iω 2t ,<br />

˜H om = ¯hω 1 â † 1â1 + ¯hω m ˆb † ˆb + ¯hh(ˆb + ˆb † )(â † 1 α 2e −iω 2t + α ∗ 2 eiω 2tâ<br />

1 ).<br />

(C.1)<br />

Defining a detuning δ L = ω 2 − ω 1 indicating the frequency difference between the two cavity<br />

modes, we c<strong>an</strong> rewrite equation (C.1) in a rotating frame,<br />

˜H om = −¯hδ L â † 1â1 + ¯hω m ˆb † ˆb + ¯h m (ˆb + ˆb † )(â 1 + â † 1 ),<br />

(C.2)<br />

where m = hα 2 (we have redefined the phases such that m is real). In the weak driving limit<br />

( m κ), the cavity dynamics c<strong>an</strong> be formally eliminated to arrive at effective optically induced<br />

cooling equations for the mech<strong>an</strong>ical motion [14, 15]. In particular, the optomech<strong>an</strong>ical coupling<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


17<br />

terms â † 1 ˆb + h.c. <strong><strong>an</strong>d</strong> â † 1 ˆb † + h.c. induce <strong>an</strong>ti-Stokes <strong><strong>an</strong>d</strong> Stokes scattering, respectively. These<br />

processes yield respective optically induced cooling (Ɣ − ) <strong><strong>an</strong>d</strong> heating (Ɣ + ) rates,<br />

Ɣ ∓ =<br />

κ 2 m<br />

(δ L ±ω m ) 2 + (κ/2) 2 . (C.3)<br />

In the case where δ L = −ω m , the cooling process is reson<strong>an</strong>tly enh<strong>an</strong>ced by the cavity, yielding<br />

a cooling rate Ɣ opt ≡ Ɣ − (δ L = −ω m ) = 4 2 m<br />

/κ as given in the main text. Also in this case, the<br />

κ<br />

optical heating rate is given by Ɣ + = Ɣ 2<br />

opt . This leads to the net cooling dynamics given<br />

κ 2 +16ωm<br />

2<br />

by equation (9) in the main text,<br />

dE m<br />

dt<br />

κ 2<br />

= −γ m (E m − ¯hω m ¯n th ) − Ɣ opt E m + Ɣ opt (E<br />

κ 2 + 16ωm<br />

2 m + ¯hω m ). (C.4)<br />

Because the optical cooling process removes phonons from the mech<strong>an</strong>ical system via optical<br />

photons that leak out of the cavity, one c<strong>an</strong> identify (ω 1 /ω m )Ɣ opt E m as the amount of optical<br />

power that is being leaked by the cavity in the <strong>an</strong>ti-Stokes sideb<strong><strong>an</strong>d</strong> during the cooling process.<br />

κ<br />

Similarly, the cavity leaks <strong>an</strong> amount of power (ω 1 /ω m )Ɣ 2<br />

opt (E<br />

κ 2 +16ωm 2 m + ¯hω m ) in the Stokes<br />

sideb<strong><strong>an</strong>d</strong>. We have ignored this contribution in equation (10) in the main text, because its<br />

large frequency separation (2ω m ) from the signal allows it to be filtered out, but otherwise it<br />

approximately contributes <strong>an</strong> extra factor of 2 to the last term in equation (10). Finally, we<br />

remark that the expression for P noise given by equation (10) represents <strong>an</strong> upper bound in that it<br />

does not account for the possibility that the output spectrum from a single element may exceed<br />

the tr<strong>an</strong>sparency b<strong><strong>an</strong>d</strong>width, which could cause some <strong>light</strong> to be absorbed within the system<br />

after multiple reflections <strong><strong>an</strong>d</strong> not make it to the end of the waveguide.<br />

Appendix D. The b<strong><strong>an</strong>d</strong> structure <strong>an</strong>alysis<br />

D.1. Derivation of dispersion relation<br />

For simplicity, here we work only with the classical equations so that the intrinsic noise terms in<br />

the Heisenberg–L<strong>an</strong>gevin equations c<strong>an</strong> be ignored. We begin by tr<strong>an</strong>sforming equations (A.7)<br />

<strong><strong>an</strong>d</strong> (A.8) to the Fourier domain,<br />

0 =<br />

(<br />

0 =<br />

(<br />

iδ k − κ )<br />

a + i m b + i<br />

2<br />

iδ k − γ m<br />

2<br />

)<br />

b + i m a.<br />

√<br />

cκex<br />

2 (a R,in(z j ) + a L,in (z j )), (D.1)<br />

(D.2)<br />

To simplify the notation, we define operators at the boundaries of the unit cells (immediately to<br />

the left of <strong>an</strong> optomech<strong>an</strong>ical element) given by c j = −i √ ca R (z − j ), d j = −i √ ca L (z − j ). It is also<br />

convenient to re write the tr<strong>an</strong>sfer matrix M om in the form<br />

( )<br />

1 − β −β<br />

M om =<br />

, (D.3)<br />

β 1 + β<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


18<br />

with the parameter β(δ k ) given by<br />

β(δ k ) =<br />

−iκ ex δ k<br />

−iκ in δ k + 2( 2 m − δ2 k ).<br />

(D.4)<br />

The tr<strong>an</strong>sfer matrix M block describing propagation to the next unit cell c<strong>an</strong> subsequently be<br />

diagonalized, M block = SDS −1 , with the diagonal matrix D given by<br />

( )<br />

e<br />

iK d<br />

0<br />

D =<br />

0 e −iK d . (D.5)<br />

Physically, this diagonalization corresponds to finding the Bloch wavevectors K (δ k ) of the<br />

periodic system. The dispersion relation for the system c<strong>an</strong> be readily obtained through the<br />

equation<br />

cos(K (δ k )d) = cos(kd) − iβ(δ k ) sin(kd).<br />

(D.6)<br />

Writing k in terms of δ k , we arrive at kd = ω 1 d/c + δ k d/c. As described previously, the desirable<br />

operation regime of the system is such that the phase imparted in free propagation should be<br />

ω 1 d/c = (2n + 1)π/2. For concreteness, we set here ω 1 d/c = π/2, satisfying this condition.<br />

For the frequencies δ k of interest, which easily satisfy the condition |δ k | ≪ d/c <strong><strong>an</strong>d</strong> ignoring the<br />

intrinsic loss κ in , the simple approximate dispersion formula<br />

κ exδ k<br />

cos(K (δ k )d) = −<br />

2( 2 m − δ2 k ) (D.7)<br />

c<strong>an</strong> be found. This dispersion relation yields two b<strong><strong>an</strong>d</strong> gaps, which extend from ±κ ex /2 <strong><strong>an</strong>d</strong><br />

±2 2 m /κ, in the weakly coupled EIT regime ( m κ ex ). We therefore have three br<strong>an</strong>ches in<br />

the b<strong><strong>an</strong>d</strong> structure, with the narrow central br<strong>an</strong>ch having a width of 4 2 m /κ ex. This br<strong>an</strong>ch has<br />

<strong>an</strong> optically tunable width <strong><strong>an</strong>d</strong> yields the slow-<strong>light</strong> propagation.<br />

The dispersive <strong><strong>an</strong>d</strong> lossy properties of the <strong>array</strong> c<strong>an</strong> also be found by <strong>an</strong>alyzing<br />

equation (D.6) perturbatively. Exp<strong><strong>an</strong>d</strong>ing equation (D.6) as a power series in δ k , we find<br />

K (δ k ) = k 0 + κ exδ k<br />

+ iκ exκ in δk<br />

2<br />

2d 2 m<br />

4d 4 m<br />

which agrees with equation (7) in the main text.<br />

+ (2κ3 ex − 3κ exκ 2 in + 12κ ex 2 m )δ3 k<br />

24d 6 m<br />

+ O(δ 4 k<br />

), (D.8)<br />

D.2. Fractional occupation calculation<br />

In our system, the Bloch functions are hybrid waves arising from the mixing of optical<br />

waveguide, optical cavity <strong><strong>an</strong>d</strong> mech<strong>an</strong>ical cavity excitations. It is therefore of interest to<br />

calculate the hybrid or polaritonic properties of these waves, by studying the energy distribution<br />

of each Bloch mode.<br />

The number of photons n WG in the waveguide c<strong>an</strong> be found by taking the sum of the left<strong><strong>an</strong>d</strong><br />

right-moving photons in a section of the device. Over one unit cell, one obtains<br />

n WG = ( |c j | 2 + |d j | 2) d<br />

c .<br />

(D.9)<br />

The relation between this value <strong><strong>an</strong>d</strong> the amplitude of the hybrid Bloch wave may be found by<br />

considering the symmetry tr<strong>an</strong>sformation used to diagonalize the unit-cell tr<strong>an</strong>smission matrix.<br />

Defining C j to be the amplitude of the Bloch mode of interest, one finds that c j = s 11 C j <strong><strong>an</strong>d</strong><br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


19<br />

d j = s 21 C j , while from the properties of the symmetry matrix S, |C j | 2 = |c j | 2 + |d j | 2 . From here<br />

we c<strong>an</strong> deduce the number of excitations in the waveguide n WG , the optical cavity n o <strong><strong>an</strong>d</strong> the<br />

mech<strong>an</strong>ical cavity n m for a given Bloch wave amplitude,<br />

n WG = d c |C j| 2 ,<br />

(D.10)<br />

n o = |a| 2<br />

(D.11)<br />

= 2|β(δ k)| 2<br />

κ ex<br />

|c j + d j | 2 (D.12)<br />

= 2|β(δ k)| 2<br />

κ ex<br />

|s 11 + s 21 | 2 |C j | 2 , (D.13)<br />

n m = |b| 2<br />

= | m| 2<br />

δ 2 k + γ 2<br />

i /4|a|2 .<br />

(D.14)<br />

(D.15)<br />

We then define the fractional occupation in the mech<strong>an</strong>ical mode by n m /(n WG + n 0 + n m )<br />

(with <strong>an</strong>alogous definitions for the other components). These relations were used to plot the<br />

fractional occupation <strong><strong>an</strong>d</strong> colored b<strong><strong>an</strong>d</strong> diagrams shown in the main text.<br />

D.3. Pump input power<br />

The technicalities associated with pumping the OMC <strong>array</strong> system are subtly distinct from those<br />

in its atomic system <strong>an</strong>alogue. Due to the periodic nature of the structure <strong><strong>an</strong>d</strong> its strongly coupled<br />

property (κ ex ≫ κ in ), a b<strong><strong>an</strong>d</strong> gap in the waveguide will arise at the frequency of the ‘tuning’ or<br />

pump cavities. This prevents these cavities from being reson<strong>an</strong>tly pumped via <strong>light</strong> propagating<br />

in the waveguide. This problem may be circumvented by making the pump frequency offreson<strong>an</strong>t<br />

from ω 2 (but still at ω 1 − ω m ), <strong><strong>an</strong>d</strong> by, for example ch<strong>an</strong>ging the splitting ω 1 − ω 2<br />

to be less th<strong>an</strong> the mech<strong>an</strong>ical frequency. Interestingly, the periodic nature of the system also<br />

allows one to suppress attenuation of the pump beam through the waveguide (which would<br />

cause inhomogeneous driving of different elements along the <strong>array</strong>). Here we calculate the<br />

waveguide input powers required to drive the pump cavities. We then find the effect of the<br />

cavity dissipation rate κ in on the beam intensity, to provide estimates for the power drop-off as a<br />

function of dist<strong>an</strong>ce propagated in the system. Finally, we note that there is a trade-off between<br />

required pump intensity <strong><strong>an</strong>d</strong> pump power drop-off. In other words, by tuning the pump beam to<br />

a frequency closer to the pump cavity reson<strong>an</strong>ce, the required input power is reduced, but the<br />

attenuation per unit cell exp(−Im{K }d) is increased.<br />

To calculate the waveguide input power, we start by considering the net photon flux in the<br />

right-moving direction,<br />

R = |c j | 2 − |d j | 2 .<br />

(D.16)<br />

Using the properties of the Bloch tr<strong>an</strong>sformation matrix S <strong><strong>an</strong>d</strong> equation (D.13), this expression<br />

may be written in terms of the number of photons in the optical cavity of interest (n o, j ),<br />

R = κ ex |s 11 | 2 − |s 21 | 2<br />

n<br />

2β1<br />

2 |s 11 + s 21 | 2 o, j , (D.17)<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


20<br />

where<br />

β 1 (δ k ) =<br />

−iκ ex/2<br />

−iκ in /2 − δ k<br />

.<br />

(D.18)<br />

The required input power P in = ¯hω o R for the system parameters studied in the main text is<br />

shown in figure 4.<br />

To find the attenuation per unit cell exp(−α), with α = Im{K (δ k )}d, of this pump beam,<br />

we use a perturbative approach similar to that used to find the polaritonic b<strong><strong>an</strong>d</strong> dispersion. By<br />

exp<strong><strong>an</strong>d</strong>ing the Bloch vector as K (δ k ) = k (0) + k (−1) /δ k + k −2 /δk 2 . . . <strong><strong>an</strong>d</strong> <strong>using</strong> equation (D.6),<br />

we find k (−1) = κ ex /2d <strong><strong>an</strong>d</strong> k (−2) = iκ ex κ in /4d, implying that α ≈ κ ex κ in /4δk 2 . This approximate<br />

expression is shown along with the exact calculated values for the attenuation in figure 4.<br />

Appendix E. Implementation in <strong>an</strong> OMC<br />

For the theoretical demonstration of our slow-<strong>light</strong> scheme, we confine ourselves to a simplified<br />

model of <strong>an</strong> OMC where only the 2D Maxwell equations for TE waves <strong><strong>an</strong>d</strong> the equation of<br />

elasticity for in-pl<strong>an</strong>e deformations of a thick slab are taken into account. These equations<br />

approximate fairly accurately the qualitative characteristics of in-pl<strong>an</strong>e optical <strong><strong>an</strong>d</strong> mech<strong>an</strong>ical<br />

waves in thin slabs, <strong><strong>an</strong>d</strong> become exact as the slab thickness is increased. In this way, m<strong>an</strong>y of the<br />

intricacies of the design of high-Q photonic <strong>crystal</strong> cavities (which are treated elsewhere [34])<br />

may be ignored, <strong><strong>an</strong>d</strong> the basic design principles c<strong>an</strong> be demonstrated in a s<strong>light</strong>ly simplified<br />

system.<br />

The 2DOMC system used here utilizes the ‘snowflake’ design [34], which provides large<br />

simult<strong>an</strong>eous photonic <strong><strong>an</strong>d</strong> phononic b<strong><strong>an</strong>d</strong> gaps in frequency. Here, we choose to use optical<br />

wavelengths in the telecom b<strong><strong>an</strong>d</strong>, i.e. corresponding to a free-space wavelength of λ ≈ 1.5 µm.<br />

For this wavelength, we found that the <strong>crystal</strong> characterized by a lattice const<strong>an</strong>t a = 400 nm,<br />

snowflake radius r = 168 nm <strong><strong>an</strong>d</strong> width w = 60 nm, shown in figure 2, should work well.<br />

E.1. Optical <strong><strong>an</strong>d</strong> mech<strong>an</strong>ical cavities<br />

E.1.1. Single cavity system. We begin our design by foc<strong>using</strong> on the creation of a single<br />

optomech<strong>an</strong>ical cavity on the 2DOMC, with one relev<strong>an</strong>t optical <strong><strong>an</strong>d</strong> mech<strong>an</strong>ical mode. This<br />

cavity is formed by creating a point defect, consisting of two adjacent removed holes (a so-called<br />

‘L2’ cavity). We calculate the optical <strong><strong>an</strong>d</strong> mech<strong>an</strong>ical spectra of this cavity <strong>using</strong> COMSOL, a<br />

commercial FEM package, <strong><strong>an</strong>d</strong> find a discrete set of confined modes. Of these, one optical <strong><strong>an</strong>d</strong><br />

one mech<strong>an</strong>ical mode were chosen, exhibiting the most promising value of the optomech<strong>an</strong>ical<br />

coupling strength g (see below for calculation). These modes are shown in figures 2(b) <strong><strong>an</strong>d</strong> (c),<br />

<strong><strong>an</strong>d</strong> were found to have frequencies ν m = 11.2 GHz <strong><strong>an</strong>d</strong> ν o = 199 THz, respectively.<br />

E.1.2. Double-cavity system. From here, we move on to designing the nearly degenerate<br />

double optical cavity system with large cross-coupling rates. As two separate L2 cavities are<br />

brought close to one <strong>an</strong>other, their interaction causes the formation of even <strong><strong>an</strong>d</strong> odd optical <strong><strong>an</strong>d</strong><br />

mech<strong>an</strong>ical super modes with splittings in the optical <strong><strong>an</strong>d</strong> mech<strong>an</strong>ical frequencies. This splitting<br />

may be tuned by ch<strong>an</strong>ging the spacing between the cavities. We take the even <strong><strong>an</strong>d</strong> odd optical<br />

modes of this two-cavity system as our optical reson<strong>an</strong>ces at ω 2 <strong><strong>an</strong>d</strong> ω 1 .<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


21<br />

E.1.3. Optomech<strong>an</strong>ical coupling rates. The optomech<strong>an</strong>ical coupling arises from a shift in the<br />

optical frequency caused by a mech<strong>an</strong>ical deformation. Our Hamiltoni<strong>an</strong> for the single-cavity<br />

system c<strong>an</strong> then be written as<br />

Ĥ = ¯hω( ˆx)â † â + ¯hω m ˆb † ˆb,<br />

(E.1)<br />

where ˆx = x ZPF (ˆb † + ˆb) is the qu<strong>an</strong>tized displacement of the mech<strong>an</strong>ical mode, <strong><strong>an</strong>d</strong> x ZPF is the<br />

characteristic per-phonon displacement amplitude. The deformation-dependent frequency ω( ˆx)<br />

may be calculated to first order in ˆx <strong>using</strong> a vari<strong>an</strong>t of the Feynm<strong>an</strong>–Hellm<strong>an</strong> perturbation<br />

theory, the Johnson perturbation theory [38], which has been used successfully in the past to<br />

model OMC cavities [9, 39]. The Hamiltoni<strong>an</strong> is then given to first order by<br />

Ĥ = ¯hω o â † â + ¯hω m ˆb † ˆb + ¯hg(ˆb † + ˆb)â † â,<br />

where ω 0 is the optical mode frequency in the absence of deformation <strong><strong>an</strong>d</strong><br />

√<br />

g = ω o<br />

2<br />

(E.2)<br />

∫ (<br />

¯h dl (Q · n) ɛ|E ‖ | 2 − (ɛ −1 )|D ⊥ | 2)<br />

√<br />

2ω m d ∫ ∫ . (E.3)<br />

s<br />

dA ρ|Q|<br />

2<br />

dA ɛ|E| 2<br />

Here, E, D <strong><strong>an</strong>d</strong> Q are the optical mode electric field, optical mode displacement field <strong><strong>an</strong>d</strong><br />

mech<strong>an</strong>ical mode displacement field, respectively, d s is the thickness of the slab <strong><strong>an</strong>d</strong> ɛ(r) is<br />

the dielectric const<strong>an</strong>t.<br />

These concepts c<strong>an</strong> be extended to optically multi-mode systems, represented by the<br />

Hamiltoni<strong>an</strong><br />

Ĥ = ¯h ∑ i<br />

ω o,i â † i â i + ¯hω m ˆb † ˆb + ¯h<br />

2<br />

∑<br />

g i, j (ˆb † + ˆb)â † i â j ,<br />

i, j<br />

(E.4)<br />

where now the cross-coupling rates c<strong>an</strong> be calculated by the following expression,<br />

√ ∫ (<br />

)<br />

g i, j = ω i, j ¯h dl(Q · n) ɛE ‖∗<br />

i · E ‖ j − (ɛ−1 )Di<br />

⊥∗ · D ⊥ j<br />

√<br />

2 2ω m d ∫ ∫ s dA ρ|Q|<br />

2<br />

dA ɛ|E i | ∫ . (E.5)<br />

2 dAɛ|E j | 2<br />

We denote this expression for convenience as g i, j ≡ 〈E i | Q|E j 〉.<br />

For the modes of the L2 cavity shown in figures 2(b) <strong><strong>an</strong>d</strong> (c), the optomech<strong>an</strong>ical coupling<br />

was calculated to be 〈E|Q|E〉/2π = 489 kHz for silicon. When two cavities are brought in the<br />

vicinity of each other, supermodes form as is shown in figure 2(a). We denote the symmetric<br />

(+) <strong><strong>an</strong>d</strong> <strong>an</strong>tisymmetric (−) combinations by E ± <strong><strong>an</strong>d</strong> Q ± . These modes c<strong>an</strong> be written in terms<br />

of the modes localized at cavity 1 <strong><strong>an</strong>d</strong> 2,<br />

E ± = E 1 ± E 2<br />

√<br />

2<br />

<strong><strong>an</strong>d</strong> Q ± = Q 1 ± Q 2<br />

√<br />

2<br />

. (E.6)<br />

By symmetry, the only non-v<strong>an</strong>ishing coupling term involving the Q − (<strong>an</strong>tisymmetric<br />

mech<strong>an</strong>ical) mode is 〈E + |Q − |E − 〉. Assuming that the two cavities are sufficiently separated, we<br />

c<strong>an</strong> approximate 〈E + Q − 〉 ≈ 〈E|Q|E〉/ √ 2. For the supermodes of interest, 〈E1 + |Q − |E − 〉/2π =<br />

h/2π = 346 kHz.<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


22<br />

(a)<br />

14<br />

250<br />

(b) (c)<br />

13<br />

240<br />

W<br />

W<br />

a<br />

r<br />

ν (GHz)<br />

12<br />

11<br />

10<br />

ν m<br />

ν (THz)<br />

230<br />

220<br />

210<br />

200<br />

ν o<br />

(d)<br />

(e)<br />

w<br />

9<br />

8<br />

190<br />

180<br />

y<br />

7<br />

170<br />

x<br />

0 0.5 1<br />

k (π/d)<br />

0 0.5 1<br />

k (π/d)<br />

Figure E.1. (a) The snowflake waveguide consists of the usual snowflake lattice,<br />

with a removed row <strong><strong>an</strong>d</strong> adjusted waveguide width. The guiding region is shaded<br />

red for clarity. (b) The mech<strong>an</strong>ical b<strong><strong>an</strong>d</strong> structure exhibits a full phononic b<strong><strong>an</strong>d</strong><br />

gap at the reson<strong>an</strong>ce frequency of the mech<strong>an</strong>ical mode (ν m ). (c) The optical b<strong><strong>an</strong>d</strong><br />

structure exhibits a single b<strong><strong>an</strong>d</strong> that passes through the reson<strong>an</strong>ce frequency ν 0 of<br />

the optical mode. The waveguide acts as a single-mode optical waveguide with<br />

field pattern H z shown in (d). The H z component of the guided optical mode of<br />

opposite symmetry is shown in (e) for completeness.<br />

E.2. Properties of snowflake <strong>crystal</strong> waveguides<br />

A line defect on <strong>an</strong> OMC acts as a waveguide for <strong>light</strong> [40, 41]. Here, the line defects used<br />

consist of a removed row of holes, with the rows above <strong><strong>an</strong>d</strong> below shifted toward one <strong>an</strong>other<br />

by a dist<strong>an</strong>ce W , such that the dist<strong>an</strong>ce between the centers of the snowflakes across the line<br />

defect is √ 3a − 2W (see figure E.1(a)). The waveguide was designed such that mech<strong>an</strong>ically, it<br />

would have no b<strong><strong>an</strong>d</strong>s reson<strong>an</strong>t with the cavity frequency (see figure E.1(b)) <strong><strong>an</strong>d</strong> would therefore<br />

have no effect on the mech<strong>an</strong>ical Q factors. Optically, it was designed to have a single b<strong><strong>an</strong>d</strong><br />

crossing the cavity frequency (see figure E.1(c)) <strong><strong>an</strong>d</strong> would therefore serve as the single-mode<br />

optical waveguide required by the proposal. The b<strong><strong>an</strong>d</strong> structure of the mech<strong>an</strong>ical waveguide<br />

was calculated <strong>using</strong> COMSOL [42], while for the optical simulations, MPB [43] was used.<br />

E.3. Cavity–waveguide coupling<br />

By bringing the optical waveguide near our cavity, the guided modes of the line defect are<br />

ev<strong>an</strong>escently coupled to the cavity mode, <strong><strong>an</strong>d</strong> a coupling between the two may be induced, as<br />

shown in figure E.2(a). Control over this coupling rate is achieved at a coarse level by ch<strong>an</strong>ging<br />

the dist<strong>an</strong>ce between the cavity <strong><strong>an</strong>d</strong> the waveguide, i.e. the number of unit cells between them.<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


23<br />

(a)<br />

(b)<br />

(c)<br />

κ ex<br />

(GHz)<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.11<br />

0.12 0.13 0.14<br />

W/a<br />

0.15<br />

y<br />

Q m,TED<br />

10 6<br />

10 5<br />

10 1 10 2<br />

x<br />

10 4<br />

T 0<br />

(K)<br />

(d)<br />

ω m<br />

t = 0<br />

ω m<br />

t = π/2<br />

ω m<br />

t = π<br />

Figure E.2. (a) A plot of the magnitude of the time-averaged Poynting vector,<br />

|〈S〉 t | (in arbitrary units), shows the leakage of photons out of the double-cavity<br />

system. This induces a loss rate on the optical modes, which was used to calculate<br />

the extrinsic coupling rate κ ex , plotted in (b) for various values of W/a. (c)<br />

Plot of the mech<strong>an</strong>ical quality factor Q m,TED due to TED, as a function of the<br />

ambient temperature T 0 . (d) The time-harmonic component of the temperature<br />

field T (r) = T − T 0 is plotted at various times during a mech<strong>an</strong>ical oscillation<br />

period.<br />

We found a dist<strong>an</strong>ce of six rows to be sufficient for placing our coupling rate κ ex in a desirable<br />

r<strong>an</strong>ge. At this point, a fine-tuning of the coupling rate may be accomplished by adjustment of the<br />

waveguide width parameter W , described previously. The achievable values of κ ex are plotted<br />

against W in figure E.2(b). For the final design, W = 0.135a is used.<br />

To simulate this coupling rate, we performed finite-element simulations <strong>using</strong> COMSOL<br />

where we placed the waveguide near our cavity, <strong><strong>an</strong>d</strong> placed absorbing boundaries at the<br />

ends of the waveguide away from the cavity. The resulting time-averaged Poynting vector<br />

〈S〉 t = |E × H ∗ |/2 is plotted in figure E.2(a), showing how the power flows out of the system.<br />

E.4. Estimates for thermoelastic damping (TED)<br />

The achievable storage times of our system are determined by the lifetimes of the mech<strong>an</strong>ical<br />

reson<strong>an</strong>ces. Since we use a phononic <strong>crystal</strong>, all clamping losses have been eliminated.<br />

However, other fundamental sources of mech<strong>an</strong>ical dissipation remain, <strong><strong>an</strong>d</strong> here we provide<br />

estimates for one of these, the component due to TED [44, 45].<br />

Using the COMSOL finite-element solver [42], we solved the coupled thermal <strong><strong>an</strong>d</strong><br />

mech<strong>an</strong>ical equations for this system [46]. In these simulations, the ch<strong>an</strong>ge in the thermal<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


24<br />

conductivity <strong><strong>an</strong>d</strong> heat capacity of silicon with temperature were taken into account. The TEDlimited<br />

quality factors, Q m,TED are plotted in figure E.2(d). In these simulations, we see that<br />

for the mode simulated, Q m,TED surpasses 10 6 at bath temperatures of T 0 < 5 K. To illustrate<br />

some representative results of these simulations, we have plotted the ch<strong>an</strong>ge in temperature<br />

field T (r) from the ambient temperature versus the phase of the mech<strong>an</strong>ical oscillation in<br />

figure E.2(d). At ω m t = π/2, there are variations in temperature despite the displacement field<br />

Q being uniformly 0 at this time. This shows that at these frequencies, the temperature does not<br />

follow the displacement adiabatically.<br />

E.5. Estimate for optical pump heating<br />

As mentioned in the main text, in a realistic setting the optomech<strong>an</strong>ical driving amplitude m<br />

itself will be coupled to the bath temperature through absorption of optical pump photons in the<br />

tuning cavities. This optical pump heating of the structure is import<strong>an</strong>t in estimating the practical<br />

limits of the optomech<strong>an</strong>ical system for qu<strong>an</strong>tum applications where thermal noise impacts<br />

system perform<strong>an</strong>ce. As a realistic model for the bath temperature in our proposed silicon OMC<br />

<strong>array</strong>, we take T b = T 0 + χα 2 , where T 0 is the base temperature <strong><strong>an</strong>d</strong> χ is a temperature coefficient<br />

that describes the temperature rise in each cavity per stored cavity photon due to optical<br />

absorption. Our estimate of χ for a thin-film silicon photonic <strong>crystal</strong> structure is as follows. The<br />

absorbed power for |α| 2 photons stored in a cavity is given simply by P loss = ¯hω o κ i |α| 2 , where<br />

ω o is the reson<strong>an</strong>ce frequency <strong><strong>an</strong>d</strong> κ i the optical (intrinsic) linewidth of the cavity. If we assume<br />

that all of this power is being converted to heat, the ch<strong>an</strong>ge in temperature is T = P loss R th ,<br />

where R th is the effective thermal resist<strong>an</strong>ce of the silicon structure. There are a number of<br />

sources in the literature for R th in relev<strong>an</strong>t photonic <strong>crystal</strong> geometries [47, 48]. We choose<br />

here to use the value for a 2D <strong>crystal</strong> system in silicon, R th ≈ 2.7 × 10 4 K W −1 , which yields a<br />

per photon temperature rise of χ ∼ 2 µK assuming <strong>an</strong> intrinsic loss rate of κ i ≈ 4 × 10 9 rad s −1<br />

(Q i ≈ 3 × 10 6 ).<br />

References<br />

[1] Y<strong>an</strong>ik M F, Suh W, W<strong>an</strong>g Z <strong><strong>an</strong>d</strong> F<strong>an</strong> S 2004 Stopping <strong>light</strong> in a waveguide with <strong>an</strong> all-optical <strong>an</strong>alog of<br />

electromagnetically induced tr<strong>an</strong>sparency Phys. Rev. Lett. 93 233903<br />

[2] Scheuer J, Paloczi G T, Poon J K S <strong><strong>an</strong>d</strong> Yariv A 2005 Coupled resonator optical waveguides: toward the<br />

slowing <strong><strong>an</strong>d</strong> storage of <strong>light</strong> Opt. Photonics News 16 36–40<br />

[3] Fleischhauer M <strong><strong>an</strong>d</strong> Lukin M D 2000 Dark-state polaritons in electromagnetically induced tr<strong>an</strong>sparency Phys.<br />

Rev. Lett. 84 5094–7<br />

[4] Fleischhauer M, Imamoglu A <strong><strong>an</strong>d</strong> Mar<strong>an</strong>gos J P 2005 Electromagnetically induced tr<strong>an</strong>sparency: optics in<br />

coherent media Rev. Mod. Phys. 77 633–73<br />

[5] Liu C, Dutton Z, Behroozi C H <strong><strong>an</strong>d</strong> Hau L V 2001 Observation of coherent optical information storage in <strong>an</strong><br />

atomic medium <strong>using</strong> halted <strong>light</strong> pulses Nature 409 490–3<br />

[6] Phillips D F, Fleischhauer A, Mair A, Walsworth R L <strong><strong>an</strong>d</strong> Lukin M D 2001 Storage of <strong>light</strong> in atomic vapor<br />

Phys. Rev. Lett. 86 783–6<br />

[7] Okawachi Y et al 2006 All-optical slow-<strong>light</strong> on a photonic chip Opt. Express 14 2317–22<br />

[8] Xu Q, Dong P <strong><strong>an</strong>d</strong> Lipson M et al 2007 Breaking the delay-b<strong><strong>an</strong>d</strong>width limit in a photonic structure Nat. Phys.<br />

3 406–10<br />

[9] Eichenfield M, Ch<strong>an</strong> J, Camacho R M, Vahala K J <strong><strong>an</strong>d</strong> Painter O 2009 Optomech<strong>an</strong>ical <strong>crystal</strong>s Nature<br />

462 78–82<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


25<br />

[10] Jo<strong>an</strong>nopoulos J D, Johnson S G, Winn J N <strong><strong>an</strong>d</strong> Meade R D 2008 Photonic Crystals: Molding the Flow of<br />

Light 2nd edn (Princeton, NJ: Princeton University Press)<br />

[11] Maldov<strong>an</strong> M <strong><strong>an</strong>d</strong> Thomas E L 2006 Simult<strong>an</strong>eous localization of photons <strong><strong>an</strong>d</strong> phonons in two-dimensional<br />

periodic structures Appl. Phys. Lett. 88 251907<br />

[12] Notomi M, Kuramochi E <strong><strong>an</strong>d</strong> T<strong>an</strong>abe T 2008 Large-scale <strong>array</strong>s of ultrahigh-Q coupled n<strong>an</strong>ocavities Nat.<br />

Photon. 2 741–7<br />

[13] Safavi-Naeini A H, Alegre T P M, Winger M <strong><strong>an</strong>d</strong> Painter O 2010 Optomech<strong>an</strong>ics in <strong>an</strong> ultrahigh-Q slotted<br />

2D photonic <strong>crystal</strong> cavity Appl. Phys. Lett. 97 181106<br />

[14] Wilson-Rae I, Nooshi N, Zwerger W <strong><strong>an</strong>d</strong> Kippenberg T J 2007 Theory of ground state cooling of a mech<strong>an</strong>ical<br />

oscillator <strong>using</strong> dynamical backaction Phys. Rev. Lett. 99 093901<br />

[15] Marquardt F, Chen J P, Clerk A A <strong><strong>an</strong>d</strong> Girvin S M 2007 Qu<strong>an</strong>tum theory of cavity-assisted sideb<strong><strong>an</strong>d</strong> cooling<br />

of mech<strong>an</strong>ical motion Phys. Rev. Lett. 99 093902<br />

[16] Arcizet O, Cohadon P, Bri<strong>an</strong>t T, Pinard M <strong><strong>an</strong>d</strong> Heidm<strong>an</strong>n A 2006 Radiation-pressure cooling <strong><strong>an</strong>d</strong><br />

optomech<strong>an</strong>ical instability of a micromirror Nature 444 71–4<br />

[17] Gig<strong>an</strong> S et al 2006 Self-cooling of a micromirror by radiation pressure Nature 444 67–70<br />

[18] Schliesser A, Del’Haye P, Nooshi N, Vahala K J <strong><strong>an</strong>d</strong> Kippenberg T J 2006 Radiation pressure cooling of a<br />

micromech<strong>an</strong>ical oscillator <strong>using</strong> dynamical backaction Phys. Rev. Lett. 97 243905<br />

[19] Clel<strong><strong>an</strong>d</strong> A 2009 Optomech<strong>an</strong>ics: Photons refrigerating phonons Nat. Phys. 5 458–60<br />

[20] Safavi-Naeini A H <strong><strong>an</strong>d</strong> Painter O 2010 Design of optomech<strong>an</strong>ical cavities <strong><strong>an</strong>d</strong> waveguides on a simult<strong>an</strong>eous<br />

b<strong><strong>an</strong>d</strong>gap phononic–photonic <strong>crystal</strong> slab Optics Express 18(14) 14926–43<br />

[21] Agarwal G S <strong><strong>an</strong>d</strong> Hu<strong>an</strong>g S 2010 Electromagnetically induced tr<strong>an</strong>sparency in mech<strong>an</strong>ical effects of <strong>light</strong><br />

Phys. Rev. A 81 041803(R)<br />

[22] Schliesser A <strong><strong>an</strong>d</strong> Kippenberg T J 2010 Cavity optomech<strong>an</strong>ics with whispering-gallery-mode optical microresonators<br />

arXiv:1003.5922<br />

[23] Weis S et al 2010 Optomech<strong>an</strong>ically induced tr<strong>an</strong>sparency Science 330 1520–23<br />

[24] Safavi-Naeini A H et al 2010 Electromagnetically induced tr<strong>an</strong>sparency <strong><strong>an</strong>d</strong> slow <strong>light</strong> with optomech<strong>an</strong>ics,<br />

in preparation arXiv:1012.1934<br />

[25] Lin Q et al 2010 Coherent mixing of mech<strong>an</strong>ical excitations in n<strong>an</strong>o-optomech<strong>an</strong>ical structures Nat. Photon.<br />

4 236–42<br />

[26] Gröblacher S, Hammerer K, V<strong>an</strong>ner M <strong><strong>an</strong>d</strong> Aspelmeyer M 2009 Observation of strong coupling between a<br />

micromech<strong>an</strong>ical resonator <strong><strong>an</strong>d</strong> <strong>an</strong> optical cavity field Nature 460 724–7<br />

[27] Hong T et al 2009 Realization of coherent optically dense media via buffer-gas cooling Phys. Rev. A<br />

79 013806<br />

[28] Y<strong>an</strong>ik M F <strong><strong>an</strong>d</strong> F<strong>an</strong> S 2005 Stopping <strong><strong>an</strong>d</strong> storing <strong>light</strong> coherently Phys. Rev. A 71 110503<br />

[29] Figueroa E, Vewinger F, Appel J <strong><strong>an</strong>d</strong> Lvovsky A I 2006 Decoherence of electromagnetically induced<br />

tr<strong>an</strong>sparency in atomic vapor Opt. Lett. 31 2625–7<br />

[30] Shen J T <strong><strong>an</strong>d</strong> F<strong>an</strong> S 2005 Coherent photon tr<strong>an</strong>sport from spont<strong>an</strong>eous emission in one-dimensional<br />

waveguides Opt. Lett. 30 2001–3<br />

[31] Ch<strong>an</strong>g D E, Sørensen A S, Hemmer P R <strong><strong>an</strong>d</strong> Lukin M D 2006 Qu<strong>an</strong>tum optics with surface plasmons Phys.<br />

Rev. Lett. 97 053002<br />

[32] Genes C, Vitali D, Tombesi P, Gig<strong>an</strong> S <strong><strong>an</strong>d</strong> Aspelmeyer M 2008 Ground-state cooling of a micromech<strong>an</strong>ical<br />

oscillator: comparing cold damping <strong><strong>an</strong>d</strong> cavity-assisted cooling schemes Phys. Rev. A 77 033804<br />

[33] Genes C, Mari A, Tombesi P <strong><strong>an</strong>d</strong> Vitali D 2008 Robust ent<strong>an</strong>glement of a micromech<strong>an</strong>ical resonator with<br />

output optical fields Phys. Rev. A 78<br />

[34] Safavi-Naeini A H <strong><strong>an</strong>d</strong> Painter O 2011 Proposal for <strong>an</strong> optomech<strong>an</strong>ical travelling wave phonon–photon<br />

tr<strong>an</strong>slator New J. Phys. 13 013017<br />

[35] Rosenberg J, Lin Q <strong><strong>an</strong>d</strong> Painter O 2009 Static <strong><strong>an</strong>d</strong> dynamic wavelength routing via the gradient optical force<br />

Nat. Photonics 3 478–83<br />

[36] Gardiner C W <strong><strong>an</strong>d</strong> Collett M J 1985 Input <strong><strong>an</strong>d</strong> output in damped qu<strong>an</strong>tum systems: qu<strong>an</strong>tum stochastic<br />

differential equations <strong><strong>an</strong>d</strong> the master equation Phys. Rev. A 31 3761–74<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)


26<br />

[37] Meystre P <strong><strong>an</strong>d</strong> Sargent M III 1999 Elements of Qu<strong>an</strong>tum Optics 3rd edn (New York: Springer)<br />

[38] Johnson S G et al 2002 Perturbation theory for Maxwell’s equations with shifting material boundaries Phys.<br />

Rev. E 65 066611<br />

[39] Eichenfield M, Ch<strong>an</strong> J, Safavi-Naeini A H, Vahala K J <strong><strong>an</strong>d</strong> Painter O 2009 Modeling dispersive coupling <strong><strong>an</strong>d</strong><br />

losses of localized optical <strong><strong>an</strong>d</strong> mech<strong>an</strong>ical modes in optomech<strong>an</strong>ical <strong>crystal</strong>s Opt. Express 17 20078–98<br />

[40] Chutin<strong>an</strong> A <strong><strong>an</strong>d</strong> Noda S 2000 Waveguides <strong><strong>an</strong>d</strong> waveguide bends in two-dimensional photonic <strong>crystal</strong> slabs<br />

Phys. Rev. B 82 4488–92<br />

[41] Johnson S G, Villeneuve P R, F<strong>an</strong> S <strong><strong>an</strong>d</strong> Jo<strong>an</strong>nopoulos J D 2000 Linear waveguides in photonic-<strong>crystal</strong> slabs<br />

Phys. Rev. B 62 8212–22<br />

[42] 2009 COMSOL Multphysics 3.5<br />

[43] Johnson S G <strong><strong>an</strong>d</strong> Jo<strong>an</strong>nopoulos J D 2001 Block-iterative frequency-domain methods for Maxwell’s equations<br />

in a pl<strong>an</strong>ewave basis Opt. Express 8 173–90<br />

[44] Lifshitz R <strong><strong>an</strong>d</strong> Rouke M L 2000 Thermoelastic damping in micro- <strong><strong>an</strong>d</strong> n<strong>an</strong>omech<strong>an</strong>ical systems Phys. Rev. B<br />

61 5600–9<br />

[45] Zener C 1938 Internal friction in solids. II General theory of thermoelastic internal friction Phys. Rev. 53 90–9<br />

[46] Duwel A, C<strong><strong>an</strong>d</strong>ler R, Kenny T <strong><strong>an</strong>d</strong> Varghese M 2006 Engineering MEMS resonators with low thermoelastic<br />

damping J. Microelectromech. Syst. 15 1437–45<br />

[47] Barclay P E, Srinivas<strong>an</strong> K <strong><strong>an</strong>d</strong> Painter O 2005 Nonlinear response of silicon photonic <strong>crystal</strong> microcavities<br />

excited via <strong>an</strong> integrated waveguide <strong><strong>an</strong>d</strong> fiber taper Opt. Express 13 801–20<br />

[48] Haret L-D, T<strong>an</strong>abe T, Kuramochi E <strong><strong>an</strong>d</strong> Notomi M 2009 Extremely low power optical bistability in silicon<br />

demonstrated <strong>using</strong> 1D photonic <strong>crystal</strong> n<strong>an</strong>ocavity Opt. Express 17 21108–17<br />

New Journal of Physics 13 (2011) 023003 (http://www.njp.org/)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!