<<

1

Controlling phonons and photons at the wavelength-scale: meets silicon phononics

AMIR H.SAFAVI-NAEINI1,D RIES VAN THOURHOUT2,R OEL BAETS2, AND RAPHAËL VAN LAER1,2

1Department of Applied and Ginzton Laboratory, Stanford University, USA 2Photonics Research Group (INTEC), Department of , Ghent University–imec, Belgium .Center for Nano- and Biophotonics, Ghent University, Belgium

Compiled October 9, 2018

Radio-frequency communication systems have long used bulk- and surface-acoustic-wave de- vices supporting ultrasonic mechanical waves to manipulate and sense signals. These devices have greatly improved our ability to process microwaves by interfacing them to orders-of- magnitude slower and lower loss mechanical fields. In parallel, long-distance communications have been dominated by low-loss infrared optical photons. As electrical signal processing and approaches physical limits imposed by energy dissipation, optical links are now being actively considered for mobile and cloud technologies. Thus there is a strong driver for wavelength-scale mechanical wave or “phononic” circuitry fabricated by scalable semiconduc- tor processes. With the advent of these circuits, new micro- and nanostructures that combine electrical, optical and mechanical elements have emerged. In these devices, such as optomechan- ical waveguides and resonators, optical photons and gigahertz phonons are ideally matched to one another as both have wavelengths on the order of micrometers. The development of phononic circuits has thus emerged as a vibrant field of research pursued for optical signal pro- cessing and sensing applications as well as emerging quantum technologies. In this review, we discuss the key physics and figures of merit underpinning this field. We also summarize the state of the art in nanoscale electro- and optomechanical systems with a focus on scalable plat- forms such as silicon. Finally, we give perspectives on what these new systems may bring and what challenges they face in the coming years. In particular, we believe hybrid electro- and optomechanical devices incorporating highly coherent and compact mechanical elements on a chip have significant untapped potential for electro-optic modulation, quantum microwave-to- optical photon conversion, sensing and microwave signal processing. arXiv:1810.03217v1 [physics.optics] 7 Oct 2018 CONTENTS 3 Photon-phonon interactions5 A Interactions between phonons and optical photons5 1 Introduction2 B Interactions between phonons and microwave photons ...... 7 2 Guiding and confining phonons3 4 State of the art8 A Total internal reflection ...... 3 B Impedance mismatch ...... 3 5 Perspectives9 C Geometric softening ...... 4 A Single-photon nonlinear optics ...... 9 D Phononic bandgaps ...... 4 B Efficient optical modulation ...... 10 E Other confinement mechanisms ...... 4 C Hybrid quantum systems ...... 11 F Material limits ...... 4 D Microwave signal processing ...... 12 2

E General challenges ...... 12 of length scales was used to demonstrate the first 2D- and 3D- confined systems in which both photons and phonons are con- 6 Conclusion 13 fined to the same area or volume [Fig.1]. These measurements have been enabled by advances in low-loss photonic circuits A Theoretical description 21 that couple to material deformations through boundary A Cavities: 3D-confined ...... 21 and photoelastic perturbations. Direct capacitive or piezoelectric A.1 Linear detection of motion ...... 22 coupling to these types of resonances has been harder since the A.2 Back-action on the mechanical mode . . . 22 relatively low speed of sound in solid-state materials means that A.3 Understanding optomechanical coupling gigahertz-frequency phonons have very small volume, leading in a nanophotonic system ...... 23 to miniscule electrically-induced forces at reasonable voltages, B Waveguides: 2D-confined ...... 24 or in other words large motional resistances that are difficult to match to standard microwave circuits [7]. B Single-photon nonlinearity 25

Wavelength-scale Photons Phonons 1. INTRODUCTION con nement Microwave-frequency acoustic or mechanical wave devices have found numerous applications in radio-signal processing and sensing. They already form mature technologies with large mar- 0D kets. The vast majority of these devices are made of piezoelectric materials that are driven by electrical circuits [1–6]. A major technical challenge in such systems is obtaining the suitable 1D l matching conditions for efficient conversion between electrical and mechanical energy. Typically, this entails reducing the effec- tive electrical impedance of the electromechanical component by increasing the capacitance of the driving element. This has 2D generally led to devices with large capacitors that drive mechan- τ τ ical modes with large mode volumes. Here, we describe a recent shift in research towards structures that are only about a wave- length, i.e. roughly one micron at gigahertz frequencies, across 3D in two or more dimensions. Greater confinement of mechanical waves in a device has both advantages and drawbacks depending on the application at hand. In the case of interactions with optical fields, higher Fig. 1. Confining photons and phonons to the wavelength- confinement increases the strength and speed of the interac- scale. Photonic and phononic systems can be classified ac- tion allowing faster switching and lower powers. A smaller cording to the number of dimensions in which they confine system demands less dissipated energy to achieve the same ef- photons and phonons to a wavelength. New structures have fects, simply because it focuses all of the optical and mechanical emerged in which both photons and phonons are confined energy into a smaller volume. High confinement also enables to the wavelength-scale in two or three spatial dimensions. scalable, less costly fabrication with more functionality packed Here we focus on such 2D- or 3D-confined wavelength-scale into a smaller space. Perhaps more importantly, in analogy to systems at gigahertz frequencies. Related reviews on inte- microwave and photonic circuits that become significantly easier grated opto- and electromechanical systems are [8–13]. The to in the single- and few-moded limits, obtaining con- table gives as examples of 2D- and 3D-confined devices a sub- 2 3 trol over the full mode structure of the devices vastly simplifies µm silicon photonic-phononic waveguide [14] and a sub-µm designing and scaling systems to higher complexity. Confin- silicon optomechanical crystal [15]. The depicted 0D- and ing mechanical energy is not without its drawbacks; as we will 1D-confined structures are a long Fabry-Pérot cavity, a vertical- see below, focusing the mechanical energy into a small volume cavity surface-emitting laser [16] for the optical case and a also means that deleterious nonlinear effects manifest at lower thick quartz [17] and a thin aluminum nitride BAW resonator powers, and matching directly to microwave circuits becomes [18] for the mechanical case. significantly more difficult due to vanishing capacitances. We can classify confinement in terms of its dimensionality [Fig.1]. In this review, we primarily consider recent advances in The dimensionality refers to the number of dimensions where gigahertz-frequency phononic devices. These devices have been confinement is on the scale of the wavelength of the excitation demonstrated mainly in the context of photonic circuits, and in bulk. For example, surface acoustic wave (SAW) resonators share many commonalities with integrated photonic structures [1], much like thin-film bulk acoustic wave (BAW) resonators [2], in terms of their design and physics. They also have the poten- have wavelength-scale confinement in only one dimension – per- tial to realize important new functionalities in photonic circuits. pendicular to the chip surface – and are therefore 1D-confined. Despite recent demonstrations of confined phonon devices op- Until a few years ago, wavelength-scale phononic confinement erating at gigahertz frequencies and coupled to optical fields, at gigahertz frequencies beyond 1D remained out of reach. phononic circuits are still in their infancy, and applications be- Intriguingly, both near-infrared optical photons and giga- yond those of interest in integrated photonics remain largely hertz phonons have a wavelength of about one micron. This unexplored. Several attractive aspects of mechanical elements re- results from the five orders of magnitude difference in the speed main unrealized in chip-scale systems, especially in those based of light relative to the speed of sound. The fortuitous matching on non-piezoelectric materials. In this review, we first describe 3 the basic physics underpinning this field with specific attention cal [27] but not optical fields. Still, some structures find a sweet to the mechanical aspects of optomechanical devices. We dis- spot in this trade-off: the principle of total internal reflection is cuss common approaches used to guide and confine mechanical currently exploited to guide phonons in Ge-doped optical fibers waves in nanoscale structures in section2. Next, we describe the [28] and chalcogenide waveguides [9]. key mechanisms behind interactions between phonons and both Since silicon is “slower” than silicon dioxide optically, but optical and microwave photons in section3. These interactions “faster” acoustically, simple index guiding for co-confined optical allow us to efficiently generate and read-out mechanical waves and mechanical fields is not an option in the canonical platform on a chip. Section4 briefly summarizes the state of the art in of silicon photonics, silicon-on-insulator. Below we consider opto- and electromechanical devices. It also describes a few techniques that circumvent this limitation and enable strongly commonly used figures of merit in this field. Finally, we give co-localized optomechanical waves and interactions. our perspectives on the field in section5. In analogy to silicon photonics [19–24], the field may be termed “silicon phononics”. Ω Si While not strictly limited to the material silicon, its goal is to (a) Ω (b) develop a platform whose fabrication is in principle scalable to SiO τ τ 2 many densely integrated mechanical devices. K Γ K

SiO2 Si 2. GUIDING AND CONFINING PHONONS Si Si SiO2

SiO2 Phonons obey broadly similar physics as photons so they can be Si guided and confined by comparable mechanisms, as detailed in Slow materials Impedance mismatch the following subsections.

A. Total internal reflection (c) Ω (d) Ω Ω In a system with continuous translational symmetry, waves incident on a medium totally reflect when they are not phase- Γ X K K matched to any excitations in that medium. This is called total K

Si internal reflection. The waves can be confined inside a slow Si medium sandwiched between two faster media by this mech- Si anism [Fig.2a]. This ensures that at fixed frequency Ω the SiO2 guided wave is not phase-matched to any leaky waves since its Geometric softening Bandgaps wavevector K(Ω) = Ω/vφ – with vφ its phase velocity – exceeds the largest wavevector among waves in the surrounding media at that frequency. In other words, the confined waves must have Fig. 2. Mechanisms for phononic confinement in micro- maximal slowness 1/vφ. This principle applies to both optical and nanostructures. We illustrate the main approaches with and mechanical fields [25, 26]. phononic dispersion diagrams Ω(K) and mark the operat- Still, there are important differences between the optical and ing point in black. (a) A waveguide core whose mechanical mechanical cases. For instance, a bulk material has only two excitations propagate slower than the slowest waves in the sur- transverse optical polarizations while it sustains two transverse rounding materials supports acoustic total internal reflection. mechanical polarizations with speed vt and a longitudinally po- Examples include chalcogenide rib waveguides on silica [9], larized mechanical wave with speed vl. Unlike in the optical silica waveguides cladded by silicon [27] and Ge-doped fibers case, these polarizations generally mix in a complex way at in- [28]. (b) Even when phonons are phase-matched to surface- or terfaces [25]. In addition, a boundary between a material and bulk-excitations, their leakage can be limited by impedance air leads to geometric softening (see next section), a situation mismatch such as in suspended silicon waveguides and disks in which interfaces reduce the speed of certain mechanical po- [14, 29–32] and silica microtoroids [33]. (c) In contrast to op- larizations. This generates slow SAW modes that are absent tical fields, mechanical waves can be trapped by surface per- in the optical case. So achieving mechanical confinement re- turbations that soften the elastic response such as in case of quires care in looking for the slowest waves in the surrounding Rayleigh surface waves [1], silicon fins on silica [34, 35] and structures. These are often surface instead of bulk excitations. all-silicon surface perturbations [36–39]. (d) Finally, phonons Among the bulk excitations, transversely polarized are slower can be trapped to line- or point-defects in periodic structures than longitudinally polarized phonons (vt < vl). with a phononic bandgap such as in silicon optomechanical Conflicting demands often arise when designing waveguides crystals [15, 40], line defects [41] and bullseye disks [42]. Many or cavities to confine photons and phonons in the same region: structures harness a combination of these mechanisms. photons can be confined easily in dense media with a high re- fractive index and thus small speed of light but phonons are naturally trapped in soft and light materials with a small speed B. Impedance mismatch of sound. In particular, the mechanical phase velocities scale as The generally conflicting demands between photonic and p vφ = E/ρ with E the stiffness or Young’s modulus and ρ the phononic confinement (see above) can be reconciled through mass density. For instance, a waveguide core made of silicon impedance mismatch [Fig.2c]. The characteristic acoustic

(refractive index nSi = 3.5) and embedded in silica (nSiO2 = 1.45) impedance of a medium is Zm = ρvφ with ρ the mass den- strongly confines photons by total internal reflection but cannot sity [25]. Interfaces between media with widely different easily trap phonons (for exceptions see next sections). On the impedances Zm, such as between solids and gases, strongly other hand, a waveguide core made of silica (vt = 5500 m/s) reflect phonons. In addition, gases have an acoustic cut-off fre- embedded in silicon (vt = 5843 m/s) can certainly trap mechani- quency – set by the molecular mean-free path – above which 4 they do not support acoustic excitations [43]. At atmospheric bandgap cavities and waveguides, does not lead to a full pressure this frequency is roughly Ωc/(2π) ≈ 0.5 GHz. Above phononic bandgap. Conversely, a rectangular array of cross- this frequency Ωc acoustic leakage and damping because of air shaped holes in a slab as has been used to demonstrate full are typically negligible. The cut-off frequency Ωc can be dras- phononic bandgaps in silicon and other materials, does not sup- tically reduced with vacuum chambers, an approach that has port a photonic bandgap. Nonetheless, both one-dimensional been pursued widely to confine low-frequency phonons [10]. [52] and two-dimensional crystals [40] with simultaneous pho- These ideas were harnessed in silicon-on-insulator waveguides tonic and phononic gaps have been proposed and demonstrated to confine both photons and phonons to silicon waveguide cores in technologically relevant material systems. [14, 29, 31] over milli- to centimeter propagation lengths. The Beyond enabling 3D-confined wavelength-scale phononic acoustic impedances of silicon and silica are quite similar, so cavities, phononic bandgaps also support waveguides or wires, in these systems the silica needs to be removed to realize low which are 2D-confined defect states. These have been realized in phonon leakage from the silicon core. In one approach [14], the silicon slabs with a pattern of cross-shaped holes supporting a silicon waveguide was partially under-etched to leave a small full phononic bandgap, with an incorporated line defect within silica pillar that supports the waveguide [Fig.2c]. In another, the the bandgap material [41, 53–55]. Robustness to scattering is silicon waveguide was fully suspended while leaving periodic particularly important to consider in such nano-confined guided silica or silicon anchors [29, 31]. wave structures, since as in photonics, intermodal scattering due to fabrication imperfections increases with decreasing cross- C. Geometric softening sectional area of the guided modes [56]. Single-mode phononic The guided wave structures considered above utilize full or par- wires are intrinsically more robust as they remove all intermodal tial under-etching of the oxide layer to prevent leakage of acous- scattering except backscattering. They have been demonstrated tic energy from the silicon into the oxide. Geometric softening is to allow robust and low-loss phonon propagation over millime- a technique that allows us to achieve simultaneous guiding of ter length scales [55]. Currently both multi- and single-mode light and sound in a material system without under-etching and phononic waveguides are actively being considered as a means regardless of the bulk wave velocities. Although phonons and of generating connectivity and functionality in chip-scale solid- photons behave similarly in bulk media, their interactions with state quantum emitter systems based on defects in diamond boundaries are markedly different. In particular, a solid-vacuum [57, 58]. boundary geometrically softens the structural response of the material below and thus lowers the effective mechanical phase E. Other confinement mechanisms velocity [Fig.2b]. This is the principle underpinning the 1D con- The above mechanisms for confinement cover many if not most finement of Rayleigh SAWs [25, 36]. This mechanism was used current systems. However, there are alternative mechanisms for in the 1970s in the megahertz range [36–38] to achieve 2D con- photonic and phononic confinement, including but not limited finement and was recently rediscovered for gigahertz phonons to: bound states in the continuum [59–61], Anderson localization where it was found that both light and motion can be guided [62, 63] and topological edge states [64, 65]. We do not cover in unreleased silicon-on-insulator structures [34]. More recently, these approaches here. fully 3D-confined acoustic waves have been demonstrated [35] with this approach on silicon-on-insulator where a narrow sil- F. Material limits icon fin, clamped to a silicon dioxide substrate, supports both localized photons and phonons. Phononic confinement, propagation losses and lifetimes are limited by various imperfections such as geometric disorder D. Phononic bandgaps [29, 40, 55, 66–68], thermo-elastic and Akhiezer damping [25, 69], Structures patterned periodically, such as a silicon slab with a two-level systems [70–72] and clamping losses [14, 73, 74]. grid of holes, with a period a close to half the phonons’ wave- Losses in 2D-confined waveguides are typically quantified by −1 length Λ = 2π/K result in strong mechanical reflections as in the a propagation length Lm = αm . In 3D-confined cavities one optical case. At this X-point – where K = π/a – in the dispersion usually quotes linewidths γ or quality factors Qm = ωm/γ. diagram forward- and backward-traveling phonons are strongly A cavity’s internal loss rate can be computed from the decay coupled, resulting in the formation of a phononic bandgap [Fig. length Lm through γ = vmαm in high-finesse cavities with negli- 2d] whose size scales with the strength of the periodic perturba- gible bending losses [75] with vm the mechanical group velocity. tion. The states just below and above the bandgap can be tuned Mechanical propagation lengths in bulk crystalline silicon are by locally and smoothly modifying geometric properties of the limited to Lm ≈ 1 cm at room-temperature and at a frequency lattice, resulting in the formation of line- or point-defects. This of ωm/(2π) = 1 GHz by thermo-elastic and Akhiezer damping. technique is pervasive in photonic crystals [44] and was adapted Equivalently, taking vm ≈ 5000 m/s one can expect material- to the mechanical case in the last decade [45–50]. This led to limited minimum linewidths of γ/2π ≈ 0.1 MHz and maximum 4 the demonstration of optomechanical crystals that 3D-confine quality factors of Qm ≈ 10 [25, 69] at ωm/(2π) = 1 GHz. both photons and gigahertz phonons to a wavelength-scale sus- These limits deteriorate rapidly at higher frequencies, typ- −2 −1 pended silicon nanobeams [40, 51, 52]. In these experiments, ically scaling as Lm ∝ ωm and Qm ∝ ωm [69, 72] or worse. confinement in one or two dimensions was obtained by periodic This makes the fm · Qm product a natural figure of merit for patterning of a bandgap structure, while in the remaining di- mechanical systems. For gigahertz frequency resonators at room mension confinement is due to the material being removed to temperature, the highest demonstrated values of fm · Qm are 13 obtain a suspended beam or film. on the order of 10 in several materials [76]. Intriguingly, the Conflicting demands similar to those discussed in sec- maximum length of time that a quantum state can persist inside tion2B complicate the design of simultaneous photonic-phononic a mechanical resonator with quality factor Qm at temperature hQ¯ m bandgap structures [49]. For example, a hexagonal lattice of T is given by tdecoherence = kT , and so requiring that the infor- circular holes in a silicon slab as is often used in photonic mation survive for more than a mechanical cycle is equivalent 5

−1 12 to the condition tdecoherence > ωm , or fm · Qm > 6 × 10 Hz as in Brillouin and Raman scattering and optomechanics, at room temperature [10]. Recently, new loss mitigation mech- where the latter includes capacitive electromechanics. anisms called “strain ” and “soft clamping” have been invented for megahertz mechanical resonators that enable • Direct coupling [Fig.3b]: one photon and one phonon inter- 8 mechanical quality factors and fm · Qm products beyond 10 act with each other directly as in piezoelectrics. and 1015 Hz respectively under high vacuum but without re- The parametric three-wave mixing takes place via two routes: frigeration [77–79]. This unlocks exciting new possibilities for quantum-coherent operations at room temperature. Finally, we • Difference-frequency driving (DFD): two photons with fre- note that many material loss processes, with the exception of quencies ω and ω0 drive the mechanical system through a 0 two-level systems [70], vanish rapidly at low temperatures (sec- beat note at frequency ω − ω = Ω ≈ ωm in the forces. tion4). • Sum-frequency driving (SFD): two photons with frequen- cies ω and ω0 drive the mechanical system through a beat (a) Parametric coupling 0 note at frequency ω + ω = Ω ≈ ωm in the forces. Photons Phonons

y ω ω’ Three-wave DFD is the only possible mechanism when the pho- ensi t ω tons and phonons have a large energy gap, as in interactions Ω al D

t r Ω ω’ Ω between phonons and optical photons. In contrast, microwave K photons can interact with phonons through any of the three- Spe c β’ β K wave and direct processes. Ω

(b) Direct coupling A. Interactions between phonons and optical photons Parametric DFD in a cavity is generally described by an interac-

y Ω tion Hamiltonian of the form (see Appendix) ensi t ω al D H = ( ) † t r ω Ω int h¯ ∂xωo a ax (1) Spe c β K with ∂xωo the sensitivity of the optical cavity frequency ωo to mechanical motion x and a the photonic annihilation opera- tor. The terminology “parametric” refers to the parameter ω , Fig. 3. Generating and detecting phonons. (a) Interactions o essentially the photonic energy, being modulated by the me- between phonons and high-frequency photons occur through chanical motion [82–85], whereas the term “three-wave mix- parametric three-wave mixing: two high-frequency photons ing” points out that there are three operators in the Hamilto- couple to one phonon via third-order nonlinearities such as nian given by equation1. This does not restrict the interaction photoelasticity and the moving-boundary effect [80, 81]. In- to only three waves, as discussed further on. Describing the teractions between low-frequency photons and a phonon can Hamiltonian H in this manner is a concise way of captur- also occur through these mechanisms. Depending on which int ing all the consequences of the interaction between the electro- of the three waves is pumped, the interaction results in ei- magnetic field a and the mechanical motion x. The detailed ther down/up-conversion (δa δb + h.c.) or state-swapping dynamics can be studied via the Heisenberg equations of mo- (δa δb† + h.c.) events. The frequency-difference between the tion defined by a˙ = − i [a, H ] when making use of the har- two high-frequency photons ω − ω0 must approximately equal h¯ int † the phononic frequency Ω for efficient parametric interactions monic oscillator commutator [a, a ] = 1 [10]. Since by definition † to occur (left). In addition, in structures with translational x = xzp δb + δb with xzp the mechanical zero-point fluctua- symmetry the wavevector-difference between the two high- tions and δb the phonon annihilation operator, this is equivalent frequency photons β − β0 must also approximately equal the to † † phononic wavevector K for efficient coupling (right). Here we Hint = hg¯ 0a a(δb + δb ) (2) depict only DFD, SFD proceeds analogously but with minus with signs replaced by plus signs. (b) Direct conversion via second- g = (∂xωo)xzp (3) order nonlinearities such as is possible when 0 the photonic and phononic energies match. Stronger mechan- the zero-point optomechanical coupling rate, which quantifies ical waves can typically be generated by direct conversion the shift in the optical cavity frequency ωo induced by the zero- than by indirect three-wave mixing of two optical waves. A point fluctuations xzp of the mechanical oscillator. Here we microwave photon can be converted into a phonon and sub- neglect the static mechanical motion [10, 86, 87]. Achieving sequently into an optical photon by cascading two of these large g0 thus generally requires small structures with large sen- 1/2 processes. sitivity ∂xωo and zero-point motion xzp = (h¯ /(2ωmmeff)) , where meff is the effective mass of the mechanical mode. This is brought about by ensuring a good overlap between the phononic 3. PHOTON-PHONON INTERACTIONS field and the photonic forces acting on the mechanical system [14, 80, 81] and by focusing the photonic and phononic energy In this section we describe the key mechanisms underpinning into a small volume to reduce meff. There are typically separate the coupling between photons and phonons. Photon-phonon bulk and boundary contributions to the overlap integral. The interactions occur via two main mechanisms: bulk contribution is associated with photoelasticity, while the • Parametric coupling [Fig.3a]: two photons and one phonon boundary contribution results from deformation of the interfaces interact with each other in a three-wave mixing process between materials [80, 81, 88, 89]. Achieving strong interactions 6 requires careful engineering of a constructive interference be- an infinitely long waveguide where phase-matching is strictly tween these contributions [14, 80, 81, 88]. Optimized nanoscale enforced. In contrast, a finite-length waveguide allows for in- silicon structures with mechanical modes at gigahertz frequen- teractions between a wider set of modes, although it suppresses cies typically have xzp ≈ 1 fm and g0/(2π) ≈ 1 MHz (section those with a large phase-mismatch (see Appendix). In essence, 4). The zero-point fluctuation amplitude increases with lower shorter waveguides permit larger violations of momentum con- frequency leading to an increase in g0: megahertz-frequency servation. The momentum selectivity can enable non-reciprocal mechanical systems with g0/(2π) ≈ 10 MHz have been demon- transport of both photons [95–98] and phonons [50, 99]. It is a strated [90]. continuum version of interference-based synthetic magnetism The dynamics generated by the Hamiltonian of equation2 schemes using discrete optomechanical elements [54, 100]. can lead to a feedback loop. The beat note between two photons Cavities can be realized by coiling up or terminating a 2D- with slightly different frequencies ω and ω0 generates a force that confined waveguide with mirrors. Then the cavity’s optome- 0 drives phonons at frequency ω − ω = Ω. Conversely, phonons chanical coupling rate g0 is connected to the waveguide’s cou- modulate, at frequency Ω, the optical field, scattering photons pling rate g0|β+K by into up- and downconverted sidebands. This feedback loop g0|β+K g0 = √ (8) can amplify light or sound, lead to electromagnetically induced L transparency, or cooling of mechanical modes. In principle, this with L the roundtrip length of the cavity (see Appendix). The interaction can even cause strong nonlinear interactions at the parameters g and g are directly related to the so-called few photon or phonon limit if g /κ > 1 [91], though current 0|β+K 0 0 Brillouin gain coefficient G that is often used to quantify photon- solid-state systems are more than two orders of magnitude away B phonon interactions in waveguides [11, 14, 31]. In particular from this regime (see Figure5 and section5). [75], Assuming g /κ  1, valid in nearly all systems, we linearize 0 4g2 the Hamiltonian of equation2 by setting a = α + δa with α a 0|β+K GB = (9) classical, coherent pump amplitude, yielding vpvs(h¯ ω)γ     with v and v the group velocities of the interacting photons, H = hg¯ δa + δa† δb + δb† (4) p s int h¯ ω the photon energy and γ the phononic decay rate. Equations 8 and9 enable comparison of the photon-phonon interaction with g = g α the enhanced interaction rate – taking α real – and 0 strengths of waveguides and cavities. Since this gain coefficient δa and δb the annihilation operators representing photonic and depends on the mechanical quality factor Q via γ = ω /Q , phononic signals respectively. Often there are experimental con- m m m it is occasionally worth comparing waveguides in terms of the ditions that suppress a subset of interactions present in Hamil- ratio G /Q . The g /(2π) ≈ 1 MHz measured in silicon op- tonian4. For instance, in sideband-resolved optomechanical B m 0 tomechanical crystals [15] is via equation9 in correspondence cavities (ωm > κ) a blue-detuned pump α sets up an entangling −1 −1 with the GB/Qm ≈ 10 W m measured in silicon nanowires interaction   † † at slightly higher frequencies [14, 29]. Both g0 and GB/Qm have Hint = hg¯ δaδb + δa δb (5) an important dependence on mechanical frequency ωm: lower- that creates or annihilates photon-phonon pairs. Similarly, a frequency structures are generally more flexible and thus gener- red-detuned pump α sets up a beam-splitter interaction ate larger interaction rates. The Hamiltonians given in equations5,6 and7 describe a  † †  Hint = hg¯ δaδb + δa δb (6) wide variety of effects. The detailed consequences of the three- wave mixing depend on the damping, intensity, dispersion and that converts photons into phonons or vice versa. This beam- momentum of the interacting fields. Next, we describe some splitter Hamiltonian can also be realized by pumping the of the potential dynamics. We quantify the dissipation experi- phononic instead of the photonic mode. In that case, α rep- enced by the photons and phonons with decay rates κ and γ resents the phononic pump amplitude, whereas both δa and δb respectively. The following regimes appear: are then photonic signals. • Weak coupling: g  κ + γ. The phonons and photons In multi-mode systems, such as in 3D-confined cavities with can be seen as independent entities that interact weakly. A several modes or in 2D-confined continuum systems, the in- common figure of merit for the interaction is the coopera- teraction Hamiltonian is a summation or integration over each tivity C = 4g2/(κγ), which quantifies the strength of the of the possible interactions between the individual photonic feedback loop discussed above. In particular, for C  1 and phononic modes. For instance, linearized photon-phonon the optomechanical back-action dominates the dynamics. interactions in a 2D-confined waveguide with continuous trans- The pair-generation Hamiltonian5 generates amplification, lational symmetry are described by [92–94]: whereas the beam-splitter interaction6 generates cooling ZZ h¯  †  and loss. Whether the phonons or the photons dominantly Hint = √ dβ dK gβ+K aβbK + gβ−K aβbK + h.c. (7) 2π experience this amplification and loss depends on the ratio κ/γ of their decay rates. The linewidth of the phonons is In this case the three-wave mixing interaction rate gβ+K = effectively (1 ∓ C)γ when κ  γ where the minus-sign in ∓ ? g0|β+Kαβ+K is proportional to the amplitude αβ+K of the mode holds for the amplification case (Hamiltonian5). In contrast, with wavevector β + K, which is usually considered to be the linewidth of the photons is effectively (1 ∓ C)κ when pumped strongly. In contrast to the single-mode cavity de- γ  κ. A lasing threshold is reached for the phonons or the scribed by equation4, in the waveguide case the symmetry photons when C = 1. In waveguide systems described by between the two-mode-squeezing δaδb and the beam-splitter equation7, C = 1 is equivalent to the transparency point † δa δb terms is broken by momentum selection from the onset as GBPp/α = 1 with Pp the pump power and α the waveguide generally gβ+K 6= gβ−K. The Hamiltonian of equation7 assumes propagation loss per meter. In fact, interactions between 7

0 photons and phonons in a waveguide can also be captured sum-frequency ω + ω = Ω ≈ ωm [3]. Such interactions can be in terms of a cooperativity which is identical to C under realized in capacitive electromechanics, where the capacitance of only weakly restrictive conditions [75]. an electrical circuit depends on mechanical motion. In particular, this sets up an interaction • Strong coupling: g  κ + γ. The phonons and photons interact so strongly that they can no longer be considered (∂ C)V2x H = − x (11) independent entities. Instead, they form a photon-phonon int 2 polariton with an effective decay rate (κ + γ)/2. The beam- with ∂ C the sensitivity of the capacitance C(x) to the me- splitter interaction6 sets up Rabi oscillations between pho- x chanical motion x and V the voltage across the capacitor. In tons and phonons with a period of 2π/g [10, 33, 101]. This terms of ladder operators we have V = V (a + a†) and is a necessary requirement for broadband intra-cavity state zp = ( + †) swapping, but is not strictly required for narrowband itin- x xzp δb δb such that erant state conversion [102, 103].  aa + a†a† 1    H = hg¯ a†a + + δb + δb† (12) Neglecting dynamics and when the detuning from the mechani- int 0 2 2 cal resonance is large (∆Ω  γ), the phonon ladder operator is This interaction contains three-wave DFD (Hamiltonian2) as a δb = (g /∆Ω)a†a such that Hamiltonian2 generates an effective 0 subset via the a†a term with an interaction rate g given by dispersive Kerr nonlinearity described by 0 hg¯ = −(∂ C)V2 x (13) g2 0 x zp zp H = h¯ 0 a†aa†a (10) Kerr ∆Ω In addition to three-wave DFD, it also contains three-wave SFD via the aa and a†a† terms. These little explored terms enable This effective Kerr nonlinearity [104–108] is often much stronger electromechanical interactions beyond the canonical three-wave than the intrinsic material nonlinearities. Thus a single optome- DFD optomechanical and Brillouin interactions. chanical system can mediate efficient and tunable interactions be- Further, by applying a strong bias voltage V the capacitive tween up to four photons in a four-wave mixing process that an- b interaction gets linearized: using V = V + δV and keeping only nihilates and creates two photons. The mechanics enhances the b the 2V δV term in V2 yields intrinsic optical material nonlinearities for applications such as b wavelength conversion [109] and photon-pair generation [110]. Hint = −(∂xC)VbδVx (14) Additional dynamical effects exist in the multi-mode case. † For instance, in a waveguide described by equation7 there is With δV = Vzp(δa + δa ) this generates an interaction a spatial variation of the photonic and phononic fields that is † † absent in the optomechanical systems described by equation4. Hint = hg¯ (δa + δa )(δb + δb ) (15) This includes: which is identical to the linearized optomechanics Hamiltonian • The steady-state spatial Brillouin amplification of an optical in expression4 with an interaction rate set by sideband. This has been the topic of recent research in chip- hg¯ = −(∂ C)V V x (16) scale photonic platforms. One can show that an optical x b zp zp Stokes sideband experiences a modified propagation loss that is enhanced with respect to g0 by g = g0α and α = Vb/Vzp (1 − C)α with C = GBPp/α the waveguide’s cooperativity the enhancement factor. The linearized Hamiltonian 15 realizes [75]. This Brillouin gain or loss is accompanied by slow or a tunable, effective piezoelectric interaction that can directly fast light [111]. Here we assumed an optical decay length convert microwave photons into phonons and vice versa. Piezo- exceeding the mechanical decay length, which is valid in electric structures are described by equation 15 as well with an nearly all systems. In the reverse case, the mechanical wave intrinsically fixed bias Vb determined by material properties. ( − C) 1/2 experiences a modified propagation loss 1 αm and Since Vzp = ((h¯ ωµ)/(2C)) with ωµ the microwave fre- there is slow and fast sound [75, 112, 113]. quency and C the total capacitance, the electromechanical cou- pling rate can be written as • Traveling photonic pulses can be converted into traveling phononic pulses in a bandwidth far exceeding the mechan- ∂ C g = − x ω x (17) ical linewidth. This is often called Brillouin light storage 0 2C µ zp [114–117]. The traveling optical pump and signal pulses or alternatively as g = (∂ ω )x – precisely as in section3A may counterpropagate or occupy different optical modes. 0 x µ zp but with the optical frequency√ ωo replaced by the microwave Several of these and other multi-mode effects have received lit- frequency ωµ with ωµ = 1/ LinC and Lin the circuit’s induc- tle attention so far. This may change with the advent of new tance. Typically the capacitance C = Cm(x) + Cp consists of a nanoscale systems realizing multi-mode and continuum Hamil- part that responds to mechanical motion Cm(x) and a part Cp tonians with strong coupling rates [14, 31, 92, 93, 118]. that is fixed and usually considered parasitic. This leads to ∂ C B. Interactions between phonons and microwave photons g = − x m η ω x (18) 0 2C p µ zp The above section3A on parametric three-wave DFD also applies m to interactions between phonons and microwave photons. How- with ηp = Cm/C the participation ratio that measures the frac- ever, microwave photons may interact with phonons via two tion of the capacitance responding to mechanical motion. For the additional routes: (1) three-wave SFD and (2) direct coupling. canonical parallel-plate capacitor with electrode separation s, we In three-wave SFD, two microwave photons with a frequency have ∂xCm = Cm/s such that g0 = −ηpxzpωµ/(2s). Similar to below the phonon frequency ωm excite mechanical motion at the the optomechanics case, this often drives research towards small 8 structures with large zero-point motion xzp and small electrode Comparably high quality factors are measured electrically in separation s. Contrary to the optomechanics case, however, in- quartz and sapphire at lower frequencies [159, 161]. It is an open creasing the participation ratio ηp motivates increasing the size question whether these extreme lifetimes have reached intrinsic and thus the motional capacitance Cm of the structures until material limits. The long lifetimes make mechanical systems ηp ≈ 1. In gigahertz-range microwave circuits with unity partic- attractive for delay lines and qubit storage [162] (section5). ipation and electrode separations on the order of s ≈ 100 nm, we Next, we look at the coupling strengths in these systems 5 have g0/(2π) ≈ −10 Hz, about a factor ωo/ωµ ≈ 10 smaller [Fig.5]. As discussed in section3, a few different figures of merit than the optomechanical g0/(2π) ≈ 1 MHz (section3A). De- are commonly used depending on the type of system. We believe spite the much smaller g0, it is still possible to achieve large the dimensionless ratios g0/κ and the cooperativity C are two 2 cooperativity C = 4g /(κγ) in electromechanics as the typical of the most powerful figures of merit (section5). The ratio g0/κ microwave linewidths are much smaller and the enhancement determines the single-photon nonlinearity, the energy-per-bit in factors α can be larger than in the optical case [119–121]. optical modulators as well as the energy-per-qubit in microwave- to-optical photon converters. The cooperativity C must be unity

1011 for efficient state conversion as well as for phonon and photon lasing. In the context of waveguides it measures the maximum Brillouin gain as C = GBPp/α [75]. 1010 optical cavity Thus we compute g0/κ for about fifty opto- and electrome- microwave cavity chanical cavities and waveguides [Fig.5a]. We convert the optical waveguide simulation waveguide Brillouin coefficients GB to g0 via expressions8 and 109 other 9 by estimating the minimum roundtrip length L a cavity made from the waveguide would have. In addition, we convert the

108 waveguide propagation loss α to the intrinsic loss rate κin = αvg with vg the group velocity. This brings a diverse set of systems together in single figure. No systems exceed g0/κ ≈ 0.01, with 107 the highest values obtained in silicon optomechanical crystals [15, 52], Brillouin-active waveguides [14, 29, 163] and Raman cavities [145]. There is no strong relation between g0/κ and C: 106 systems with low interactions rates g0 often have low decay rates κ and γ as well since they do not have quite as stringent fabrication requirements on the surface quality. 105 The absolute zero-point coupling rates g0 illustrate the power of moving to the nanoscale. We plot them as a function of the 104 maximum quantum cooperativity Cq = C/n¯th with n¯th the ther- mal phonon occupation [Fig.5b]. When Cq > 1, the state transfer between photons and phonons takes place more rapidly than 103 the mechanical thermal decoherence [10]. This is a requirement for hybrid quantum systems such as efficient microwave-to- optical photon converters (section5). There are several chip- 102 105 106 107 108 109 1010 1011 1012 1013 scale electro- and optomechanical systems that obtained Cq > 1, with promising values demonstrated in silicon photonic crys- tals. A main impediment to large quantum cooperativities in Fig. 4. Mechanical quality factors. The systems include opto- optomechanics is the heating of the mechanics caused by optical and electromechanical cavities and waveguides at room tem- absorption [128]. perature (red), at a few kelvin (blue) and at millikelvin temper- Further, we give an overview of the Brillouin coefficients G atures (blue). The highest quality factors are demonstrated in B found in 2D-confined waveguides [Fig.5c]. The current record millikelvin cavities. The data points correspond to references 4 −1 −1 [14, 15, 28–31, 33, 35, 40, 52, 67, 81, 90, 101, 120, 122–160]. GB = 10 W m in the gigahertz range was measured in a suspended series of silicon nanowires [29]. However, larger Bril- louin amplification was obtained with silicon and chalcogenide rib waveguides which have disproportionately lower optical 4. STATE OF THE ART propagation losses α and can handle larger optical pump pow- Here we give a concise overview of the current state of the art in ers Pp [163, 164]. We stress that the maximum Brillouin gain opto- and electromechanical systems by summarizing the param- is identical to the cooperativity [75]. They are both limited by eters obtained in about fifty opto- and electromechanical cavities the maximum power and electromagnetic energy density the and waveguides. First, we plot the mechanical quality factors system in question can withstand. At room temperature in sili- as a function of mechanical frequency [Fig.4] including room con, the upper limit is usually set by two-photon and free-carrier temperature (red) and cold (blue) systems. As discussed in sec- absorption [14, 31, 89]. Moving beyond the two-photon bandgap tion2F, cold systems usually reach much higher quality factors. of 2200 nm in silicon or switching to materials such as silicon ni- The current record is held by a 5 GHz silicon optomechanical tride, lithium niobate or chalcogenides can drastically improve 10 crystal with Qm > 10 , yielding a lifetime longer than a sec- the power handling [89, 144, 165, 166]. In cold systems, it is ond [122] at millikelvin temperatures. Measuring these quality instead set by the cooling power of the refrigerator and the heat- factors requires careful optically pulsed read-out techniques, as ing of the mechanical system [128]. Another challenge for 2D- the intrinsic dissipation of continuous-wave optical photons eas- confined waveguides is the inhomogeneous broadening of the ily heats up the mechanics thus destroying its coherence [128]. mechanical resonance. This arises from atomic-scale fluctuations 9

(a)10-2 optical cavity

10-3 microwave cavity optical waveguide simulation 10-4 other

10-5

10-6 10-2 10-1 100 101 102 103 104 105

(b) (c) 106

106

104

104

102

102 100

100 10-3 10-2 10-1 100 101 102 103 104 100 101 102 103

Fig. 5. Interaction rates in 3D-confined cavities and 2D-confined waveguides. (a) The dimensionless nonlinearity g0/κ versus the maximum cooperativity C. The largest interaction rates are achieved in nanoscale cavities and waveguides. Cooperativities are typically highest in cold systems (blue) and can be as high in less tightly confined systems since they often have lower photonic and phononic decay rates than the smallest systems. (b) The zero-point coupling rate g0 versus the maximum quantum cooperativity Cq = C/n¯th with n¯th the thermal phonon occupation for a selection of cold 3D-confined systems. It is significantly harder to reach Cq > 1 than C > 1. In particular, optical absorption easily heats up mechanical systems – effectively increasing n¯th [128]. (c) The Brillouin gain coefficient GB versus maximum net Brillouin gain (GBPp/α − 1) for a selection of 2D-confined waveguides. Up to extrinsic cavity losses, C − 1 = (GBPp/α − 1). Achieving this maximum Brillouin gain requires the waveguide to be have a length L = 1/α with α the optical propagation loss per meter. However, this is challenging as longer waveguides can suffer from an inhomogeneous broadening of the mechanical resonance due to atomic-scale disorder in the waveguide geometry. in the waveguide geometry along its length effectively smearing tuning from the mechanical resonance in the above expressions: out the mechanical response [14, 29, 31, 66]. Finally, compared ωm → 2∆Ω with the detuning ∆Ω = ∆ω − ωm. This enhances to gigahertz systems, flexible megahertz mechanical systems the shift per photon so that quantum nonlinearities are realized 6 −1 −1 give much higher efficiencies of GB ≈ 10 W m as measured at [170, 171] 9 −1 −1 g2 in dual-nanoweb [167] fibers and of GB ≈ 10 W m as pre- 2 0 ϑcav = ≈ π (19) dicted in silicon double-slot waveguides [168]. κ∆Ω with ∆Ω  ωm. The photon blockade effect also requires > 5. PERSPECTIVES sideband-resolution (∆Ω κ) so g 0 > 1 (20) A. Single-photon nonlinear optics κ The three-wave mixing interactions discussed in section3 in is generally a necessary condition for single-photon nonlinear principle enable single-photon nonlinear optics in opto- and optics with opto- and electromechanical cavities [10, 172]. In the electromechanical systems [91, 169, 170]. For instance, in the case of 2D-confined waveguides, it can similarly be shown [118] photon blockade effect a single incoming photon excites the that a single photon drives a mechanically-mediated cross-Kerr motion of a mechanical system in a cavity, which then shifts phase shift 2 the cavity resonance and thus blocks the entrance of another g0|β+K photon. Realizing such quantum nonlinearities sets stringent ϑwg = (21) vg∆Ω requirements on the interaction strengths and decay rates. on another photon with v the optical group velocity (see Ap- For instance, in an optomechanical cavity the force exerted g † pendix). The cross-Kerr phase-shift ϑwg can be enhanced drasti- by a single photon is hFi = −h∂xHinti = −h¯ (g0/xzp)ha ai = cally by reducing the group velocity vg via Brillouin slow light −h¯ (g /xzp). To greatly affect the optical response seen by an- 0 [111, 118, 173]. If sufficiently large, the phase-shifts ϑ and ϑ other photon impinging on the cavity, this force must drive a cav wg can be used to realize controlled-phase gates between photonic mechanical displacement that shifts the optical resonance by qubits – an elementary building block for quantum information about a linewidth κ or xπ = κ/(∂xωo) = (κ/g0)xzp. In other 2 processors [118, 174–176]. Using equation8, we have words, we require F/(meffωm) = xπ which leads to ϑcav ≡ 4g2/(κω ) ≈ π where ϑ is the mechanically-mediated cross- F 0 m cav ϑcav = ϑwg (22) phase shift experienced by the other photon assuming critical π coupling to the cavity. This extremely challenging condition is with F = 2π/(κTrt) the cavity finesse and Trt the cavity relaxed when two photonic modes with a frequency difference roundtrip time. Therefore, cavities generally yield larger single- ∆ω roughly resonant with the mechanical frequency are used. photon cross-Kerr phase shifts than their corresponding optome- In this case, the mechanical frequency can be replaced by the de- chanical waveguides. 10

Currently state-of-the-art solid-state and sideband-resolved a microwave cavity [187, 188]. The required Vπ corresponds to 2 (ωm > κ) opto- and electromechanical systems yield at best an energy-per-bit Ebit = CVπ/2 which again can be rewritten as g0/κ ≈ 0.01 in any material [Fig.5]. Significant advances in g0 expression 23. Electro-optic materials such as lithium niobate may be made in e.g. nanoscale slotted structures [90, 153, 177], [189] may yield up to g0/(2π) ≈ 10 kHz, corresponding to an but it remains an open challenge to not only increase g0 but also energy-per-bit Ebit ≈ 10 fJ/bit keeping the optical linewidth κ g0/κ by a few orders of magnitude [178]. Beyond exploring constant – on the order of today’s world records [19]. novel structures, other potential approaches include effectively Although full system demonstrations using mechanics for boosting g0/κ by parametrically amplifying the mechanical mo- electro-optic modulation are lacking, based on estimates like tion [179], by employing delayed quantum feedback [180] or these we believe that mechanics will unlock highly efficient via collectively enhanced interactions in optomechanical arrays electro-optic systems. The expected much lower energy-per-bit [181, 182]. Although single-photon nonlinear optics may be out implies that future electro-opto-mechanical modulators could of reach for now, many-photon nonlinear optics can be enhanced achieve much higher bitrates at fixed power, or alternatively, very effectively with mechanics. Specifically, mechanics realizes much lower dissipated power at fixed bitrate than current direct Kerr nonlinearities orders of magnitude beyond those of typical electro-optic modulators. Although the mechanical linewidth intrinsic material effects. This is especially so for highly flexible, does not enter expression 23, bandwidths of a single device are low-frequency mechanical systems [104, 105, 183, 184] but has usually limited by the phononic quality factor or transit time been shown in gigahertz silicon optomechanical cavities and across the device. Interestingly, the mechanical displacement waveguides as well [14, 109]. corresponding to the estimated 1 aJ/bit is only xπ ≈ 10 pm. Here we highlighted the potential for optical modulation B. Efficient optical modulation based on mechanical motion at gigahertz-frequencies. However, Phonons provide a natural means for the spatiotemporal mod- similar arguments can be made for optical switching networks ulation of optical photons via electro- and optomechanical in- based on lower frequency mechanical structures. In particular, teractions. Hybrid circuits that marry photonic and phononic voltage-driven capacitive or piezoelectric optical phase-shifters excitations give us access to novel opto-electro-mechanical sys- exploiting mechanical motion do not draw static power and can tems. Two aspects of the physics make phononic circuits very generate large optical phase shifts in small devices [12, 190–196]. attractive for the modulation of optical fields. These “photonic MEMS” are thus an attractive elementary build- First, there is excellent spatial matching between light and ing block in reconfigurable and densely integrated photonic sound. As touched upon above, the wavelengths of microwave networks used for high-dimensional classical [19, 197–200] and phonons and telecom photons are both about a micron in tech- quantum [201–203] photonic information processors. They may nologically relevant materials such as silicon. The matching meet the challenging power- and space-constraints involved in follows from the four to five orders of magnitude difference running a complex programmable network. between the speed of sound and the speed of light. Momentum Demonstrating fully integrated acousto-optic systems re- conservation, i.e. phase-matching, between phonons and opti- quires that we properly confine, excite and route phonons on a cal photons (as discussed in section3) is key for non-reciprocal chip. Among the currently proposed and demonstrated systems nonlinear processes and modulation schemes with traveling are acousto-optic modulators [97] as well as optomechanical phonons [95, 97, 185, 186]. beam-steering systems [68, 204]. Besides showing the power of Second, the optomechanical nonlinearity is strong and essen- sound to process light with minuscule amounts of energy, these tially lossless. Small deformations can induce major changes phononic systems have features that are absent in competing on the optical response of a system. For instance, in an optome- approaches. For instance, gigahertz traveling mechanical waves chanical cavity (equation1) the mechanical motion required to with large momentum naturally enable non-reciprocal features encode a bit onto a light field has an amplitude of approximately in both modulators [97] and beam-steering systems [68]. This is essential for isolators and circulators based on indirect photonic xπ = (κ/g0)xzp. Generating this motion requires energy, and 2 2 transitions [96, 98, 205–208]. this corresponds to an energy-per-bit Ebit = meffωmxπ/2 which we rewrite as In order to realize these and other acousto-optic systems, it  2 is crucial to efficiently excite mechanical excitations on the sur- h¯ ωm κ E = (23) face of a chip. In this context, electrical excitation is especially bit 4 g 0 promising as it allows for stronger mechanical waves than opti- Thus the energy-per-bit also depends on the dimensionless cal excitation. With optical excitation of mechanical waves, the quantity g0/κ: a single phonon can switch a photon when this flux of phonons is upper-bounded by the flux of optical pho- quantity reaches unity, in agreement with section5A. For sil- tons injected into the structure. The ratio of photon to phonon −3 icon optomechanical crystals with g0/κ ≈ 10 , this yields energy limits the mechanical power to less than a microwatt, Ebit ≈ 1 aJ/bit: orders of magnitude more efficient than com- corresponding to 10-100 milliwatts of optical power. Neverthe- monly deployed electro-optic technologies [19]. less, proof-of-concept demonstrations [96, 98] have successfully The similarity between the fundamental interactions in op- generated non-reciprocity on a chip using optically generated tomechanics [10] and electro-optics [187, 188] allows to com- phonons. In contrast, microwave photons have a factor 105 pare the two types of modulation head-to-head. In particu- larger fluxes than optical photons for the same power. Therefore, lar, in an optical cavity made of an electro-optic material the microwave photons can drive milliwatt-level mechanical waves voltage drop across the electrodes required to encode a bit is in nanoscale cavities and waveguides. Such mechanical waves Vπ = κ/(∂V ωo) = (κ/g0)Vzp with g0 = (∂V ωo)Vzp the electro- can have displacements up to a nanometer and strains of a few optic interaction rate [187, 188], which is defined analogously to percent – close to material yield strengths. the optomechanical interaction rate. It is the parameter appear- Electrical generation of gigahertz phonons in nanoscale struc- † † ing in the interaction Hamiltonian Hint = hg¯ 0a a(b + b ), with tures has received little attention so far, especially in non- b + b† now proportional to the voltage across the capacitor of piezoelectric materials such as silicon and silicon nitride. As 11 discussed in section3B, this can be realized either via capacitive states from the microwave to the optical domain and vice versa. or via piezoelectric electromechanics. Capacitive approaches To realize a microwave-to-optical photon converter, one can work in any material [209] and have recently been demonstrated start from a classical electro-optic modulator and modify it to in a silicon photonic waveguide [158]. They require small capaci- protect quantum coherence. Several proposals aim to achieve tor gaps and large bias voltages to generate effects of magnitude this by coupling a superconducting microwave cavity to an comparable to piezoelectric approaches. More commonly, piezo- optical cavity made of an electro-optic material. For instance, electrics such as lithium niobate, aluminum nitride [210, 211] the beam-splitter Hamiltonian can be engineered by injecting a and lead-zirconate titanate can be used as the photonic plat- strong optical pump red-detuned from the cavity resonance in form, or be integrated with existing photonic platforms such as an electro-optic cavity. In order to suppress undesired Stokes silicon and silicon nitride in order to combine the best of both scattering events, the frequency of the microwave cavity needs worlds [68, 212–214]. Such hybrid integration typically comes to exceed the optical cavity linewidth, i.e. sideband-resolution with challenging incompatibilities in material properties [215], is necessary. In this scenario, continuous-wave state conver- especially when more than one material needs to be integrated sion with high fidelity requires an electro-optic cooperativity Ceo on a single chip. Efficient electrically-driven acoustic waves in close to unity: 2 2 photonic structures have the potential to enable isolation and 4g0|α| Ceo = = 1 (24) circulation with an optical bandwidth beyond 1 THz – limited κγµ only by optical walk-off [95, 98, 216–218]. with g0 the electro-optic interaction rate as defined in the pre- 2 C. Hybrid quantum systems vious section, |α| the number of optical pump photons in the cavity and γ the microwave cavity linewidth. The quantum Strain and displacement alter the properties of many different µ conversion is accompanied by an optical power dissipation systems and therefore provide excellent opportunities for con- P = h¯ ω |α|2κ with κ the intrinsic decay rate of the op- necting dissimilar degrees of freedom. In addition, mechanical diss o in in tical cavity. Operating the converter in a bandwidth of γ and systems can possess very long coherence times and can be used µ inserting condition 24, this leads to an energy-per-qubit of to store quantum information. In the field of hybrid quantum systems, researchers find ways to couple different degrees of ! h¯ ωo κκin freedom over which quantum control is possible to scale up and Eqbit = (25) 4 g2 extend the power of quantum systems. Realizing hybrid systems 0 by combining mechanical elements with other excitations is a which is the quantum version of the energy-per-bit 23. This widely pursued research goal. Studies on both static tuning of yields an interesting relation between the efficiency of classical quantum systems using nanomechanical forces [219–221] as well and quantum modulators as on quantum dynamics mediated by mechanical resonances Eqbit ω and waveguides [18, 57, 219, 222, 223] are being pursued. ≈ o (26) Among the emerging hybrid quantum systems, microwave- Ebit ωm to-optical photon converters utilizing mechanical degrees of We stress that Eqbit is the optical dissipated energy in a quan- freedom have attracted particular interest recently [123, 222, 224– tum converter, whereas Ebit is the microwave or mechanical 227]. In particular, one of the leading platforms to realize scal- energy necessary to switch an optical field in a classical mod- able, error-corrected quantum processors [228, 229] are supercon- ulator [238]. The quantum electro-optic modulator dissipates ducting microwave circuits in which qubits are realized using roughly five orders of magnitude more energy per converted Josephson junctions [230, 231] in a platform compatible with sili- qubit as it requires an optical pump field to drive the conver- con photonics [232]. To suppress decoherence, these microwave sion process. Strategies developed to minimize Ebit, as pursued circuits are operated at millikelvin temperatures inside dilution for decades by academic groups and the optical communica- refrigerators. Heat generation must be restricted in these cold tions industry, also tend to minimize Eqbit. Recently a coupling environments [233]. The most advanced prototypes currently rate of g0/(2π) = 310 Hz was demonstrated in an integrated consist of on the order of fifty qubits on which gates with at best aluminum nitride electro-optic resonator [239]. Switching to 0.1% error rates can be applied [233, 234]. Scaling up these sys- lithium niobate and harnessing improvements in the electro- tems to millions of qubits, as required for a fully error-corrected optic modal overlap may increase this to g /(2π) ≈ 10 kHz, quantum , is a formidable unresolved challenge [229]. 0 corresponding to Eqbit ≈ 1 nJ/qbit. Electro-optic polymers [240] Also, the flow of microwave quantum information is hindered may yield higher interaction rates g but bring along challenges outside of the dilution refrigerators by the microwave thermal 0 in optical and microwave losses κ and γµ. Cooling powers of noise present at room temperature [235, 236]. Optical photons roughly 10 µW at the low temperature stage of current dilution travel for kilometers at room temperature along today’s opti- refrigerators [233] imply that conversion rates with common cal fiber networks. Thus quantum interfaces that convert mi- electro-optic materials will likely not exceed about 10 kqbits/s. crowave to optical photons with high efficiency and low noise Considering that the g0/κ demonstrated optomechanical de- should help address the scaling and communication barriers vices is much larger than those found in electro-optic systems, hindering microwave quantum processors. They may pave the and following a reasoning similar to that presented in section way for distributed quantum computing systems or a “quantum 5B for classical modulators, it is likely that microwave-to-optical internet” [237]. Besides, such interfaces would give optical sys- photon converters based on mechanical elements as interme- tems access to the large nonlinearities generated by Josephson diaries will be able to achieve large efficiencies. It has been junctions, which enables a new approach for nonlinear optics. theoretically shown that electro-opto-mechanical cavities with The envisioned microwave-to-optical photon converters are dynamics described in section3 allow for efficient state trans- in essence electro-optic modulators that operate on single pho- duction between microwave and optical fields when tons and preserve entanglement [187]. They exploit the beam- splitter Hamiltonian discussed in section3 to swap quantum Cem ≈ Com  1 (27) 12 with Cem and Com the electro- and optomechanical coopera- tive level, as excessive fiber-to-chip losses and high modulator tivities. Noiseless conversion additionally requires negligible drive voltages associated with the discrete photonic devices cur- thermal microwave and mechanical occupations [102, 103, 224]. rently being used are the main origin of the low system efficiency. Since the dominant dissipation still arises from the optical pump, Also, the photonic-phononic emit-receive scheme proposed in the energy-per-qubit can still be expressed as in equation 25 for [254, 255] results in a lower RF-insertion loss. Although it gives a electro-opto-mechanical cavity. Given the large nonlinearity up tunability, additional advantages of this approach are that g0/κ enabled by nanoscale mechanical systems [Fig.5], we ex- its engineerable filter response [254, 256] and its cascadability pect conversion rates up to 100 Mqbits/s are feasible by operat- [255]. Exploiting the phase response of the SBS resonance also ing multiple electro-opto-mechanical photon converters in paral- phase control of RF signals has been demonstrated [250]. Both lel inside the refrigerator. State-of-the-art integrated electro- and pure phase shifters and relative time delay have been demon- opto-mechanical cavities have achieved Cem > 1 and Com > 1 in strated. Again interferometric approaches allow to amplify the separate systems [Fig.5]. It is an open challenge to achieve con- intrinsic phase delay of the system, which is limited by the avail- dition 27 in a single integrated electro-opto-mechanical device. able SBS gain. In the examples above, the filter is driven by a Finally, the long lifetimes and compact nature of mechanical single-frequency pump, resulting in a Lorentzian filter response. systems also makes them attractive for the storage of classi- More complex filter responses can be obtained by combining cal and quantum information [13, 157, 160, 162, 223, 241–244]. multiple pumps [111]. However, this comes at the cost of the Mechanical memories are currently pursued both with purely overall system response since the total power handling capacity electromechanical [18, 157, 160] and purely optomechanical of the system is typically limited. As such there is still a need for [114, 115] systems. Interfaces between mechanical systems waveguide platforms that can handle large optical powers and and superconducting qubits may lead to the generation of non- at the same time provide high SBS-gain. classical states of mesoscopic mechanical systems [10, 245, 246], Further, low-noise oscillators are also a key building block probing the boundary between quantum and classical behavior. in RF-systems. Two approaches, equivalent with the two dis- sipation hierarchies (γ  κ and γ  κ) identified in section D. Microwave signal processing 3, have been studied. In the first case, if the photon lifetime In particular in the context of wireless communications, com- exceeds the phonon lifetime (γ  κ), optical line narrowing and pact and cost-effective solutions for radio-frequency (RF) signal eventually self-oscillation is obtained at the transparency con- processing are rapidly gaining importance. Compared to purely dition C = 1 – resulting in substantial narrowing of the Stokes electronic and MEMS-based approaches, RF processing in the wave and thus a purified laser beam [257–259]. Cascading this photonics domain – microwave photonics – promises compact- process leads to higher-order Stokes waves with increasingly ness and light weight, rapid tunability and integration density narrowed linewidths. Photomixing a pair of cascaded Brillouin [247, 248]. Currently demonstrated optical solutions however lines gives an RF carrier with phase noise determined by the still suffer from high RF-insertion loss and an unfavorable trade- lowest order Stokes wave. Using this approach in a very low- off between achieving sufficiently narrow bandwidth, high rejec- loss silica disk resonator a phase noise suppression of 110 dBc at tion ratio and linearity. Solutions mediated by phonons might 100 kHz offset from a 21.7 GHz carrier was demonstrated [260]. overcome this limit as they offer a narrow linewidth without suf- In the alternate case, with the phonon lifetime exceeding the fering from the power limits experienced in high-quality optical photon lifetime (γ  κ), the Stokes wave is a frequency-shifted cavities [249, 250] copy of the pump wave apart from the phase noise added by the Given the high power requirements, 2D-confined waveg- mechanical oscillator. At the transparency condition C = 1 the uides lend themselves more naturally to many RF-applications. phonon noise goes down, eventually reaching the mechanical As such, stimulated Brillouin scattering (SBS) has been exten- Schawlow-Townes limit. Several such “phonon lasers” have sively exploited. Original work focused on phonon-photon in- been demonstrated already, relying on very different integration teractions in optical fibers, which allows for high SBS-gain and platforms [150, 156, 261–263]. Further work is needed to deter- high optical power but lacks compactness and integrability. Fol- mine if these devices can deliver the performance required to lowing the demonstration of SBS-gain in integrated waveguide compete with existing microwave oscillators. platforms [14, 31, 144], several groups now also demonstrated In the examples above, the mechanical mode is excited all- RF-signal processing using integrated photonics chips. In the optically via a strong pump beam. Both in terms of efficiency and most straightforward approach, the RF-signal is modulated on in terms of preventing the pump beam from propagating further a sideband of an optical carrier which is then overlayed with through the optical circuit this may be not the most appropriate the narrowband SBS loss-spectrum generated by a strong pump method. Recently, several authors have demonstrated electrical [251]. Tuning the carrier frequency allows rapid and straight- actuation of optomechanical circuits [53, 97, 123, 158, 264–266]. forward tuning of the notch filter over several GHz and a band- While this provides a more direct way to drive the acousto- width below 130 MHz was demonstrated. The suppression optic circuit, considerable efforts are still needed to improve the was only 20 dB however, limited by the SBS gain achievable overall efficiency of these systems and to develop a platform in the waveguide platform used, in this case a chalcogenide where all relevant building blocks including e.g. actuators and waveguide. This issue is further exacerbated in more CMOS- detectors, optomechanical oscillators and acoustic delay lines compatible platforms, where the SBS gain is typically limited can be co-integrated without loss in performance. to a few dB. This can be overcome by using interferometric ap- proaches, which enable over 45 dB suppression with only 1 dB E. General challenges of SBS gain [252, 253]. Each of the perspectives discussed above potentially benefits While this approach outperforms existing photonic and non- enormously from miniaturizing photonic and phononic systems photonic approaches on almost all specifications (see table 1 in in order to maximize interaction rates and pack more function- [252]), a remaining issue is the high RF insertion loss of about 30 ality into a constrained space. Current nanoscale electro- and dB. Integration might be key in bringing the latter to a competi- optomechanical devices indeed demonstrate some of the highest 13 interaction rates (section4). However, the fabrication of high- REFERENCES quality nanoscale systems requires exquisite process control. Even atomic-scale disorder in the geometric properties can ham- 1. K.-y. Hashimoto, Surface Acoustic Wave Devices in Telecom- per device performance, especially when extended structures munications, vol. 1 (Springer, 2000). or many elements are required [29, 267, 268]. This can be con- 2. K. Hashimoto, “RF bulk acoustic wave filters for communi- sidered the curse of moving to the nanoscale. It manifests itself cations,” in “Artech House,” (2009), p. 267. as photonic and phononic propagation loss [56, 67], backscat- 3. T. B. Jones and N. G. Nenadic, Electromechanics and MEMS tering [56], inter-modal scattering as well as inhomogeneous (Cambridge University Press, Cambridge, 2013). broadening [66], dephasing [68] and resonance splitting [29, 40]. 4. R. C. Ruby, P. B. Sm, Y. Oshmyansky, A. Chien, A. Technolo- To give a feel for the sensitivity of these systems, a 10 GHz me- gies, and C. St, “Thin Film Bulk Wave Acoustic Resonators chanical breathing mode undergoes a frequency shift of about (FBAR) for Wireless Applications,” pp. 813–821 (2001). 10 MHz per added monolayer of silicon atoms [14]. Therefore 5. R. Ruby, “Review and comparison of bulk acoustic wave nanometer-level disorder is easily resolvable in current devices FBAR, SMR technology,” Proceedings - IEEE Ultrasonics with room-temperature quality factors on the order of 103. De- Symposium pp. 1029–1040 (2007). veloping better process control and local tuning [269] methods 6. M. Hara, J. Kuypers, T. Abe, and M. Esashi, “Surface mi- is thus a major task for decades to come. In addition, shrinking cromachined AlN thin film 2 GHz resonator for CMOS systems to the nanoscale leads to large surface-to-volume ratios integration,” and Actuators, A: Physical 117, 211– that imply generally ill-understood surface physics determines 216 (2005). key device properties, even with heavily studied materials such 7. C. T.-C. Nguyen, “MEMS-based RF channel selection for as silicon [81, 270, 271]. This is a particular impediment for true software-defined cognitive radio and low-power sen- emerging material platforms such as thin-film aluminum nitride sor communications,” IEEE Communications Magazine 51, [272], lithium niobate [273] and diamond [274]. The flip side 110–119 (2013). of these large sensitivities is that opto- and electromechanical 8. D. Van Thourhout and J. Roels, “Optomechanical device systems may generate exquisite sensors of various perturbations. actuation through the optical gradient force,” Nature Pho- Amongst others, current research takes aim at inertial tonics 4, 211–217 (2010). and mass sensing [275–277] as well as local temperature [278] 9. B. Eggleton, C. Poulton, and R. Pant, “Inducing and har- and geometry mapping [279–282]. nessing stimulated Brillouin scattering in photonic inte- grated circuits,” Adv. Opt. Photon. pp. 536–587 (2013). 6. CONCLUSION 10. M. Aspelmeyer, T. Kippenberg, and F. Marquardt, “Cavity optomechanics,” Rev. Mod. Phys. 86, 1391–1452 (2014). New hybrid electro- and optomechanical nanoscale systems 11. M. Merklein, A. Casas-Bedoya, D. Marpaung, T. Buttner, have emerged in the last decade. These systems confine both M. Pagani, B. Morrison, I. Kabakova, and B. Eggleton, photons and phonons in structures about one wavelength across “Stimulated Brillouin scattering in photonic integrated cir- to set up large interaction rates in a compact space. Similar to cuits: novel applications and devices,” IEEE Journal of silicon photonics more than a decade ago, nanoscale phononic Selected Topics in Quantum p. 1 (2016). circuitry is in its infancy and severe challenges such as geomet- 12. L. Midolo, A. Schliesser, and A. Fiore, “Nano-opto-electro- ric disorder hinder its development. Still, we expect much to mechanical systems,” Nature 13 (2018). come in the years ahead. We believe mechanical systems are par- 13. K. L. Ekinci and M. L. Roukes, “Nanoelectromechanical ticularly interesting as low-energy electro-optic interfaces with systems,” Review of Scientific Instruments 76 (2005). potential use in classical and quantum information processors 14. R. Van Laer, B. Kuyken, D. Van Thourhout, and R. Baets, and sensors. Phonons are a gateway for photons to a world with “Interaction between light and highly confined hypersound five orders of magnitude slower timescales. Linking the two in a silicon photonic nanowire,” Nature Photonics 9, 199– excitations has the potential for major impact on our information 203 (2015). infrastructure in ways we have yet to fully explore. 15. A. H. Safavi-Naeini, J. Chan, J. Hill, T. Alegre, A. Krause, and O. Painter, “Observation of Quantum Motion of a FUNDING INFORMATION Nanomechanical Resonator,” Physical Review Letters 108, 1–5 (2012). We acknowledge the support from the U.S. government through 16. V. M. N. Passaro, F. Magno, and F. De Leonardis, “Opti- the National Science Foundation (Grant Nos. ECCS-1509107 and mization of Bragg reflectors in AlGaAs/GaAs VCSELs,” ECCS-1708734) and the Air Force Office of Scientific Research Laser Physics Letters 2, 239–246 (2005). under MURI. A.S.N. acknowledges the support of a David and 17. M. Goryachev, D. L. Creedon, E. N. Ivanov, S. Galliou, Lucile Packard Fellowship. R.V.L. acknowledges funding from R. Bourquin, and M. E. Tobar, “Extremely low-loss acoustic VOCATIO and from the European Union’s Horizon 2020 re- phonons in a quartz bulk acoustic wave resonator at mil- search and innovation program under Marie Skłodowska-Curie likelvin temperature,” Applied Physics Letters 100, 243504 grant agreement No. 665501 with the research foundation Flan- (2012). ders (FWO). 18. A. D. O’Connell, M. Hofheinz, M. Ansmann, R. C. Bialczak, M. Lenander, E. Lucero, M. Neeley, D. Sank, H. Wang, CONTRIBUTIONS M. Weides, J. Wenner, J. M. Martinis, and A. N. Cleland, “Quantum ground state and single-phonon control of a R.V.L. organized and wrote much of the manuscript along with mechanical resonator,” Nature 464, 697–703 (2010). A.S.N. The section on microwave signal processing was written 19. D. A. Miller, “Attojoule Optoelectronics for Low-Energy by D.V.T. All authors read, discussed and gave critical feedback Information Processing and Communications,” Journal of on the paper. Lightwave Technology 35, 346–396 (2017). 14

20. Z. Wang, A. Abbasi, U. Dave, A. De Groote, S. Kumari, Brillouin scattering and optomechanics,” APL Photonics B. Kunert, C. Merckling, M. Pantouvaki, Y. Shi, B. Tian, 071301 (2016). K. Van Gasse, J. Verbist, R. Wang, W. Xie, J. Zhang, 35. C. J. Sarabalis, Y. D. Dahmani, R. N. Patel, J. T. Hill, and Y. Zhu, J. Bauwelinck, X. Yin, Z. Hens, J. Van Campen- A. H. Safavi-Naeini, “Release-free silicon-on-insulator cav- hout, B. Kuyken, R. Baets, G. Morthier, D. Van Thourhout, ity optomechanics,” Optica 4, 1147 (2017). and G. Roelkens, “Novel Light Source Integration Ap- 36. A. A. Oliner, “Waveguide for Acoustic Surface Waves:A proaches for Silicon Photonics,” Laser & Photonics Reviews Review,” IEEE 64 (1973). 11, 1700063 (2017). 37. E. Ash, R. M. D. L. Rue, and R. F. Humphryes, “Microsound 21. C. Sun, M. T. Wade, Y. Lee, J. S. Orcutt, L. Alloatti, M. S. Surface Waveguides,” IEEE Transactions on Microwave Georgas, A. S. Waterman, J. M. Shainline, R. R. Avizienis, Theory and Techniques 17, 882–892 (1969). S. Lin, B. R. Moss, R. Kumar, F. Pavanello, A. H. Atabaki, 38. P. E. Lagasse, “Analysis of silicon ridge waveguides for H. M. Cook, A. J. Ou, J. C. Leu, Y.-H. Chen, K. Asanovi´c, acoustic surface waves,” IEEE . R. J. Ram, M. A. Popovi´c,and V. M. Stojanovi´c,“Single-chip 39. P. E. Lagasse, “Higher-order finite-element analysis of to- that communicates directly using light,” pographic guides supporting elastic surface waves,” The Nature 528, 534–538 (2015). Journal of the Acoustical Society of America 53, 1116–1122 22. A. Rahim, E. Ryckeboer, A. Subramanian, S. Clemmen, (1973). b. kuyken, A. Dhakal, A. Raza, A. Hermans, M. Muneeb, 40. A. H. Safavi-Naeini, J. T. Hill, S. Meenehan, J. Chan, S. Dhoore, Y. Li, U. Dave, P. Bienstman, N. LeThomas, S. Gröblacher, and O. Painter, “Two-Dimensional Phononic- G. Roelkens, D. Van Thourhout, X. Rottenberg, P. Helin, Photonic Band Gap Optomechanical Crystal Cavity,” Phys- S. Severi, and R. Baets, “Expanding the Silicon Photonics ical Review Letters 112, 153603 (2014). Portfolio with Silicon Nitride Photonic Integrated Circuits,” 41. K. Fang, M. H. Matheny, X. Luan, and O. Painter, “Optical Journal of Lightwave Technology XX, 1–1 (2016). transduction and routing of microwave phonons in cavity- 23. A. H. Atabaki, S. Moazeni, F. Pavanello, H. Gevorgyan, optomechanical circuits,” Nature Photonics 10, 489–496 J. Notaros, L. Alloatti, M. T. Wade, C. Sun, S. A. Kruger, (2016). H. Meng, K. A. Qubaisi, I. Wang, B. Zhang, A. Khilo, C. V. 42. F. G. S. Santos, Y. A. V. Espinel, G. O. Luiz, R. S. Benevides, Baiocco, M. A. Popovi´c,V. M. Stojanovi´c,and R. J. Ram, G. S. Wiederhecker, and T. P. Mayer Alegre, “Hybrid con- “Integrating photonics with silicon nanoelectronics for the finement of optical and mechanical modes in a bullseye next generation of systems on a chip,” Nature 556, 349–354 optomechanical resonator,” Optics Express 25, 508 (2017). (2018). 43. D. Royer and E. Dieulesaint, Elastic Waves in Solids I & II 24. R. Soref, “The Past, Present, and Future of Silicon Photon- (Springer-Verlag Berlin Heidelberg, 2000). ics,” IEEE Journal of Selected Topics in Quantum Electron- 44. J. D. Joannopoulos, S. G. Johnson, J. N. Winn, and R. D. ics 12, 1678–1687 (2006). Meade, Photonic crystals: molding the flow of light (Princeton 25. B. A. Auld, Acoustic Fields and Waves in Solids: volume I$&$II University Press, 2008). (John Wiley $&$ Sons, Inc., Stanford, 1973). 45. T. Gorishnyy, C. Ullal, M. Maldovan, G. Fytas, and 26. B. Saleh and M. Teich, Fundamentals of Photonics (John Wiley E. Thomas, “Hypersonic phononic crystals,” Phys. Rev. & Sons, Inc., 1991). Lett. 94, 115501 (2005). 27. Y. Liu, N. Dostart, and M. A. Popovi´c, “Toward Mi- 46. V. Laude, A. Khelif, S. Benchabane, M. Wilm, T. Sylvestre, crophononic Circuits on Chip: An Evaluation of Compo- B. Kibler, A. Mussot, J. Dudley, and H. Maillotte, “Phononic nents based on High-Contrast Evanescent Confinement of band-gap guidance of acoustic modes in photonic crystal Acoustic Waves,” pp. 1–11 (2017). fibers,” Physical Review B 71, 045107 (2005). 28. A. Kobyakov, M. Sauer, and D. Chowdhury, “Stimulated 47. M. Maldovan and E. L. Thomas, “Simultaneous localiza- Brillouin scattering in optical fibers,” Adv. Opt. Photon. 2, tion of photons and phonons in two-dimensional periodic 1 (2009). structures,” Applied Physics Letters 88, 251907 (2006). 29. R. Van Laer, A. Bazin, B. Kuyken, R. Baets, and 48. M. Eichenfield, R. Camacho, J. Chan, K. J. Vahala, and D. Van Thourhout, “Net on-chip Brillouin gain based on O. Painter, “A picogram- and nanometre-scale photonic- suspended silicon nanowires,” New Journal of Physics 17 crystal optomechanical cavity.” Nature 459, 550–555 (2009). (2015). 49. A. H. Safavi-Naeini and O. Painter, “Design of optomechan- 30. H. Shin, W. Qiu, R. Jarecki, J. Cox, R. Olsson, A. Starbuck, ical cavities and waveguides on a simultaneous bandgap Z. Wang, and P. Rakich, “Tailorable stimulated Brillouin phononic-photonic crystal slab.” Optics express 18, 14926– scattering in nanoscale silicon waveguides,” Nat. Commun. 14943 (2010). 4, 1944 (2013). 50. M. Maldovan, “Sound and heat revolutions in phononics,” 31. E. A. Kittlaus, H. Shin, and P. T. Rakich, “Large Brillouin Nature 503, 209–217 (2013). amplification in silicon,” Nature Photonics 10, 463–467 51. M. Eichenfield, J. Chan, R. M. Camacho, K. J. Vahala, and (2016). O. Painter, “Optomechanical crystals.” Nature 462, 78–82 32. X. Sun, X. Zhang, and H. X. Tang, “High- Q silicon optome- (2009). chanical microdisk resonators at gigahertz frequencies,” 52. J. Chan, T. Alegre, A. H. Safavi-Naeini, J. Hill, A. Krause, Applied Physics Letters 100, 173116 (2012). S. Gröblacher, M. Aspelmeyer, and O. Painter, “Laser cool- 33. E. Verhagen, S. Deléglise, S. Weis, A. Schliesser, and T. Kip- ing of a nanomechanical oscillator into its quantum ground penberg, “Quantum-coherent coupling of a mechanical state,” Nature 478, 89–92 (2011). oscillator to an optical cavity mode,” Nature 482, 63–67 53. K. C. Balram, M. Davanco, J. D. Song, and K. Srinivasan, (2012). “Coherent coupling between radio frequency, optical, and 34. C. J. Sarabalis, J. T. Hill, and A. H. Safavi-Naeini, “Guided acoustic waves in piezo-optomechanical circuits,” Nature acoustic and optical waves in silicon-on-insulator for Photonics 10, 1–20 (2015). 15

54. K. Fang, J. Luo, A. Metelmann, M. H. Matheny, F. Mar- 1–5 (2013). quardt, A. A. Clerk, and O. Painter, “Generalized non- 72. A. N. Cleland, Foundations of Nanomechanics: From Solid- reciprocity in an optomechanical circuit via synthetic mag- State Theory to Device Applications (Springer-Verlag Berlin netism and reservoir engineering,” Nature Physics 13, 465– Heidelberg, New York, USA, 2003). 471 (2017). 73. M. Zhang, G. Luiz, S. Shah, G. Wiederhecker, and M. Lip- 55. R. N. Patel, Z. Wang, W. Jiang, C. J. Sarabalis, J. T. Hill, son, “Eliminating anchor loss in optomechanical resonators and A. H. Safavi-Naeini, “Single-Mode Phononic Wire,” using elastic wave interference,” Applied Physics Letters Physical Review Letters 121, 040501 (2018). 105, 051904 (2014). 56. D. Melati, A. Melloni, F. Morichetti, D. Elettronica, 74. G. Anetsberger, R. Rivière, a. Schliesser, O. Arcizet, and I. Bioingegneria, and P. Milano, “Real photonic waveg- T. Kippenberg, “Ultralow-dissipation optomechanical res- uides : guiding light through imperfections,” 224, 156–224 onators on a chip,” Nature Photonics 2, 627–633 (2008). (2014). 75. R. Van Laer, R. Baets, and D. Van Thourhout, “Unifying 57. M. A. Lemonde, S. Meesala, A. Sipahigil, M. J. Schuetz, Brillouin scattering and cavity optomechanics,” Phys. Rev. M. D. Lukin, M. Loncar, and P. Rabl, “Phonon Networks A 93, 53828 (2016). with Silicon-Vacancy Centers in Diamond Waveguides,” 76. C. Nguyen, “MEMS technology for timing and frequency Physical Review Letters 120, 213603 (2018). control,” IEEE Transactions on Ultrasonics, Ferroelectrics 58. M. C. Kuzyk and H. Wang, “Phononic Quantum Net- and Frequency Control 54, 251–270 (2007). works of Solid-State Spins with Alternating and Frequency- 77. R. A. Norte, J. P. Moura, and S. Gröblacher, “Mechani- Selective Waveguides,” arXiv:1804.07862 p. 1804.07862 cal Resonators for Quantum Optomechanics Experiments (2018). at Room Temperature,” Physical Review Letters 116, 1–6 59. C. W. Hsu, B. Zhen, J. Lee, S.-L. Chua, S. G. Johnson, J. D. (2016). Joannopoulos, and M. Soljaˇci´c,“Observation of trapped 78. Y. Tsaturyan, A. Barg, E. S. Polzik, and A. Schliesser, “Ultra- light within the radiation continuum.” Nature 499, 188–91 coherent nanomechanical resonators via soft clamping and (2013). dissipation dilution,” Nature Nanotechnology 12, 776–783 60. J. Gomis-Bresco, D. Artigas, and L. Torner, “Anisotropy- (2017). Induced Photonic Bound States in the Continuum,” Nature 79. A. H. Ghadimi, S. A. Fedorov, N. J. Engelsen, M. J. Bereyhi, Photonics 11, accepted (2017). R. Schilling, D. J. Wilson, and T. J. Kippenberg, “Elastic 61. Y. Plotnik, O. Peleg, F. Dreisow, M. Heinrich, S. Nolte, strain engineering for ultralow mechanical dissipation,” A. Szameit, and M. Segev, “Experimental observation of Science 360, 764–768 (2018). optical bound states in the continuum,” Physical Review 80. P. Rakich, C. Reinke, R. Camacho, P. Davids, and Z. Wang, Letters 107, 28–31 (2011). “Giant enhancement of stimulated Brillouin scattering in 62. D. S. Wiersma, P. Bartolini, A. Lagendijk, and R. Righini, the subwavelength limit,” Phys. Rev. X 2, 1–15 (2012). “Localization of light in a disordered medium,” Nature 390, 81. J. Chan, A. H. Safavi-Naeini, J. T. Hill, S. Meenehan, and 671–673 (1997). O. Painter, “Optimized optomechanical crystal cavity with 63. M. Segev, Y. Silberberg, and D. Christofilos, “Anderson acoustic radiation shield,” Applied Physics Letters 101, localization of light,” Nature Photonics 7, 197–204 (2013). 081115 (2012). 64. S. H. Mousavi, A. B. Khanikaev, and Z. Wang, “Topologi- 82. W. H. Louisell, A. Yariv, and A. E. Siegman, “Quantum cally protected elastic waves in phononic metamaterials,” Fluctuations and Noise in Parametric Processes. I.” Physi- Nature Communications 6, 8682 (2015). cal Review 124, 1646–1654 (1961). 65. M. Hafezi, S. Mittal, J. Fan, a. Migdall, and J. M. Taylor, 83. N. Bloembergen and Y. Shen, “Coupling between vibra- “Imaging topological edge states in silicon photonics,” Na- tions and light waves in Raman laser media,” Physical ture Photonics 7, 1001–1006 (2013). Review Letters 12, 504–507 (1964). 66. C. Wolff, R. Van Laer, M. J. Steel, B. J. Eggleton, and C. G. 84. Y. R. Shen and N. Bloembergen, “Theory of stimulated Bril- Poulton, “Brillouin resonance broadening due to struc- louin and Raman scattering,” Physical Review 290 (1965). tural variations in nanoscale waveguides,” New Journal of 85. A. Yariv, “Quantum theory for parametric interactions of Physics 18 (2016). light and hypersound,” IEEE Journal of Quantum Electron- 67. R. N. Patel, C. J. Sarabalis, W. Jiang, J. T. Hill, and A. H. ics pp. 28–36 (1965). Safavi-Naeini, “Engineering Phonon Leakage in Nanome- 86. G. S. Wiederhecker, L. Chen, A. Gondarenko, and M. Lip- chanical Resonators,” Physical Review Applied 8, 041001 son, “Controlling photonic structures using optical forces,” (2017). Nature 462, 633–636 (2009). 68. C. J. Sarabalis, R. Van Laer, and A. H. Safavi-Naeini, “Op- 87. J. Rosenberg, Q. Lin, and O. Painter, “Static and dynamic tomechanical antennas for on-chip beam-steering,” Optics wavelength routing via the gradient optical force,” Nature Express 26, 22075 (2018). Photonics 3, 478–483 (2009). 69. S. Ghaffari, S. a. Chandorkar, S. Wang, E. J. Ng, C. H. Ahn, 88. W. Qiu, P. T. Rakich, H. Shin, H. Dong, M. Soljaˇci´c, and V. Hong, Y. Yang, and T. W. Kenny, “Quantum limit of qual- Z. Wang, “Stimulated Brillouin scattering in nanoscale sili- ity factor in silicon micro and nano mechanical resonators.” con step-index waveguides: a general framework of selec- Scientific reports 3, 3244 (2013). tion rules and calculating SBS gain,” Opt. Express 21, 31402 70. R. O. Behunin, P. Kharel, W. H. Renninger, and P. T. Ra- (2013). kich, “Engineering dissipation with phononic spectral hole 89. C. Wolff, M. J. Steel, B. J. Eggleton, and C. G. Poulton, “Stim- burning,” Nature Materials 16, 315–321 (2016). ulated Brillouin Scattering in integrated photonic waveg- 71. T. Ramos, V. Sudhir, K. Stannigel, P. Zoller, and T. J. Kippen- uides: forces, scattering mechanisms and coupled mode berg, “Nonlinear quantum optomechanics via individual analysis,” Physical Review A pp. 1–17 (2015). intrinsic two-level defects,” Physical Review Letters 110, 90. R. Leijssen and E. Verhagen, “Strong optomechanical inter- 16

actions in a sliced photonic crystal nanobeam,” Scientific chanics.” Nature communications 3, 1196 (2012). Reports 5, 15974 (2015). 110. R. Riedinger, S. Hong, R. A. Norte, J. A. Slater, J. Shang, 91. P. Rabl, “Photon Blockade Effect in Optomechanical Sys- A. G. Krause, V. Anant, M. Aspelmeyer, and S. Gröblacher, tems,” Physical Review Letters 107, 063601 (2011). “Non-classical correlations between single photons and 92. J. E. Sipe and M. J. Steel, “A Hamiltonian treatment of stim- phonons from a mechanical oscillator,” Nature 530, 1–12 ulated Brillouin scattering in nanoscale integrated waveg- (2015). uides,” New Journal of Physics 045004, 1–26 (2016). 111. L. Thévenaz, “Slow and fast light in optical fibres,” Nature 93. H. Zoubi and K. Hammerer, “Optomechanical multimode Photonics 2 (2008). Hamiltonian for nanophotonic waveguides,” Physical Re- 112. Y.-C. Chen, S. Kim, and G. Bahl, “Brillouin cooling in a lin- view A 94, 1–16 (2016). ear waveguide,” New Journal of Physics 18, 115004 (2016). 94. P. Rakich and F. Marquardt, “Quantum theory of contin- 113. N. T. Otterstrom, R. O. Behunin, E. A. Kittlaus, and P. T. uum optomechanics,” New Journal of Physics 20, 045005 Rakich, “Optomechanical cooling in a continuous system,” (2018). arXiv:1805.02502 (2018). 95. Zongfu Yu and Shanhui Fan, “Integrated Nonmagnetic Op- 114. Z. Zhu, D. Gauthier, and R. Boyd, “Stored light in an optical tical Isolators Based on Photonic Transitions,” IEEE Jour- fiber via stimulated Brillouin scattering.” Science 318, 1748– nal of Selected Topics in Quantum Electronics 16, 459–466 1750 (2007). (2010). 115. M. Merklein, B. Stiller, K. Vu, S. J. Madden, and B. J. Eggle- 96. J. Kim, M. C. Kuzyk, K. Han, H. Wang, and G. Bahl, “Non- ton, “A chip-integrated coherent photonic-phononic mem- reciprocal Brillouin scattering induced transparency,” Na- ory,” Nature Communications 8, 574 (2017). ture Physics 11, 275–280 (2015). 116. M. Merklein, B. Stiller, K. Vu, P. Ma, S. J. Madden, and B. J. 97. D. B. Sohn, S. Kim, and G. Bahl, “Time-reversal symmetry Eggleton, “Broad-bandwidth non-reciprocal Brillouin light breaking with acoustic pumping of nanophotonic circuits,” storage,” arXiv:1806.00146 pp. 3–5 (2018). Nature Photonics 12, 91–97 (2018). 117. B. Stiller, M. Merklein, C. Wolff, K. Vu, P. Ma, C. G. Poulton, 98. E. A. Kittlaus, N. T. Otterstrom, P.Kharel, S. Gertler, and P.T. S. J. Madden, and B. J. Eggleton, “On-chip multi-stage Rakich, “Nonreciprocal Inter-band Brillouin Modulation,” optical delay based on cascaded Brillouin light storage,” arXiv:1808.09865 (2018). arXiv:1804.08799 pp. 1–5 (2018). 99. A. Seif, W. DeGottardi, K. Esfarjani, and M. Hafezi, “Ther- 118. H. Zoubi and K. Hammerer, “Quantum Nonlinear Optics in mal management and non-reciprocal control of phonon Optomechanical Nanoscale Waveguides,” Physical Review flow via optomechanics,” Nature Communications 9, 1207 Letters 119, 123602 (2017). (2018). 119. J. M. Fink, M. Kalaee, A. Pitanti, R. Norte, L. Heinzle, M. Da- 100. V. Peano, C. Brendel, M. Schmidt, and F. Marquardt, “Topo- vanço, K. Srinivasan, and O. Painter, “Quantum electrome- logical Phases of Sound and Light,” Physical Review X chanics on silicon nitride nanomembranes,” Nature Com- 031011, 1–18 (2015). munications 7, 12396 (2016). 101. S. Gröblacher, K. Hammerer, M. Vanner, and M. As- 120. P. B. Dieterle, M. Kalaee, J. M. Fink, and O. Painter, “Su- pelmeyer, “Observation of strong coupling between a mi- perconducting Cavity Electromechanics on a Silicon-on- cromechanical resonator and an optical cavity field.” Na- Insulator Platform,” Physical Review Applied 6, 014013 ture 460, 724–727 (2009). (2016). 102. Y.-D. Wang and A. a. Clerk, “Using Interference for High 121. M. Kalaee, M. Mirhosseni, P. B. Dieterle, M. Peruzzo, J. M. Fidelity Quantum State Transfer in Optomechanics,” Phys- Fink, and O. Painter, “Quantum electromechanics of a hy- ical Review Letters 108, 153603 (2012). personic crystal,” arXiv:1808.04874 pp. 1–16 (2018). 103. Y. D. Wang and A. A. Clerk, “Using dark modes for high- 122. G. MacCabe, H. Ren, J. Luo, J. Cohen, H. Zhou, A. Ardizzi, fidelity optomechanical quantum state transfer,” New Jour- and O. Painter, “Optomechanical Measurements of Ultra- nal of Physics 14 (2012). Long-Lived Microwave Phonon Modes in a Phononic 104. J. Ma and M. L. Povinelli, “Mechanical Kerr nonlinearities Bandgap Cavity,” in “Bulletin of the American Physical due to bipolar optical forces between deformable silicon Society,” (2018). waveguides.” Optics Express 19, 10102–10110 (2011). 123. J. Bochmann, A. Vainsencher, D. D. Awschalom, and A. N. 105. W. H. P. Pernice, M. Li, and H. X. Tang, “A mechanical Kerr Cleland, “Nanomechanical coupling between microwave effect in deformable photonic media,” Appl. Phys. Lett. 95, and optical photons,” Nature Physics 9, 712–716 (2013). 123507 (2009). 124. K. E. Grutter, M. I. Davanco, K. Srinivasan, K. A. E. G. 106. S. Aldana, C. Bruder, and A. Nunnenkamp, “Equivalence Rutter, M. A. I. D. Avanço, and K. A. S. Rinivasan, “Slot- between an optomechanical system and a Kerr medium,” Mode Optomechanical Crystals: A Versatile Platform for Physical Review A - Atomic, Molecular, and Optical Multimode Optomechanics,” 1, 1–11 (2015). Physics 88, 1–10 (2013). 125. Y. Liu, M. Davanço, V. Aksyuk, and K. Srinivasan, “Electro- 107. J. Gea-Banacloche, “Impossibility of large phase shifts magnetically induced transparency and wideband wave- via the giant Kerr effect with single-photon wave pack- length conversion in silicon nitride microdisk optomechan- ets,” Physical Review A - Atomic, Molecular, and Optical ical resonators,” Physical Review Letters 110, 1–5 (2013). Physics 81 (2010). 126. M. J. Burek, J. D. Cohen, S. M. Meenehan, N. El-Sawah, 108. A. Butsch, C. Conti, F. Biancalana, and P. Russell, “Optome- C. Chia, T. Ruelle, S. Meesala, J. Rochman, H. A. Atikian, chanical Self-Channeling of Light in a Suspended Planar M. Markham, D. J. Twitchen, M. D. Lukin, O. Painter, and Dual-Nanoweb Waveguide,” Physical Review Letters 108, M. Lonˇcar, “Diamond optomechanical crystals,” Optica 3, 093903 (2012). 1404 (2016). 109. J. T. Hill, A. H. Safavi-Naeini, J. Chan, and O. Painter, “Co- 127. S. M. Meenehan, J. D. Cohen, S. Gröblacher, J. T. Hill, A. H. herent optical wavelength conversion via cavity optome- Safavi-Naeini, M. Aspelmeyer, and O. Painter, “Silicon op- 17

tomechanical crystal resonator at millikelvin temperatures,” 115, 243601 (2015). Physical Review A 90, 1–5 (2014). 143. J. M. Fink, M. Kalaee, A. Pitanti, R. Norte, L. Heinzle, M. Da- 128. S. M. Meenehan, J. D. Cohen, G. S. MacCabe, F. Marsili, vanço, K. Srinivasan, and O. Painter, “Quantum electrome- M. D. Shaw, and O. Painter, “Pulsed Excitation Dynamics chanics on silicon nitride nanomembranes,” Nature Com- of an Optomechanical Crystal Resonator near Its Quantum munications 7, 12396 (2016). Ground State of Motion,” Physical Review X 5, 041002 144. R. Pant, C. G. Poulton, D.-Y. Choi, H. Mcfarlane, S. Hile, (2015). E. Li, L. Thevenaz, B. Luther-Davies, S. J. Madden, and B. J. 129. R. Rivière, S. Deléglise, S. Weis, E. Gavartin, O. Arcizet, Eggleton, “On-chip stimulated Brillouin scattering.” Optics A. Schliesser, and T. J. Kippenberg, “Optomechanical side- Express 19, 8285–8290 (2011). band cooling of a micromechanical oscillator close to the 145. H. Rong, S. Xu, Y.-H. Kuo, V. Sih, O. Cohen, O. Raday, quantum ground state,” Physical Review A - Atomic, and M. Paniccia, “Low-threshold continuous-wave Raman Molecular, and Optical Physics 83, 1–9 (2011). silicon laser,” Nature Photonics 1, 232–237 (2007). 130. A. Schliesser, O. Arcizet, R. Rivière, G. Anetsberger, and 146. M. S. Kang, A. Nazarkin, A. Brenn, and P. S. T. Russell, T. J. Kippenberg, “Resolved-sideband cooling and position “Tightly trapped acoustic phonons in photonic crystal fibres measurement of a micromechanical oscillator close to the as highly nonlinear artificial Raman oscillators,” Nature Heisenberg uncertainty limit,” Nature Physics 5, 509–514 Phys. 5, 276–280 (2009). (2009). 147. K. S. Abedin, “Brillouin amplification and lasing in a single- 131. D. J. Wilson, V. Sudhir, N. Piro, R. Schilling, a. Ghadimi, mode As2Se 3 chalcogenide fiber,” Optics Letters 31, 1615– and T. J. Kippenberg, “Measurement-based control of a me- 1617 (2006). chanical oscillator at its thermal decoherence rate,” Nature 148. A. Butsch, J. Koehler, R. Noskov, and P. Russell, “CW- 524, 325–329 (2015). pumped single-pass frequency comb generation by res- 132. A. H. Safavi-Naeini, S. Gröblacher, J. T. Hill, J. Chan, M. As- onant optomechanical nonlinearity in dual-nanoweb fiber,” pelmeyer, and O. Painter, “Squeezed light from a silicon Optica 1, 158–164 (2014). micromechanical resonator,” Nature 500, 185–189 (2013). 149. E. A. Kittlaus, N. T. Otterstrom, and P. T. Rakich, “On-chip 133. A. G. Krause, J. T. Hill, M. Ludwig, A. H. Safavi-Naeini, inter-modal Brillouin scattering,” Nature Communications J. Chan, F. Marquardt, and O. Painter, “Nonlinear Radi- 8, 15819 (2017). ation Pressure Dynamics in an Optomechanical Crystal,” 150. B. Morrison, A. Casas-Bedoya, G. Ren, K. Vu, Y. Liu, Physical Review Letters 115, 1–5 (2015). A. Zarifi, T. G. Nguyen, D.-Y. Choi, D. Marpaung, S. J. Mad- 134. X. Sun, X. Zhang, M. Poot, C. Xiong, and H. X. Tang, “A den, A. Mitchell, and B. J. Eggleton, “Compact Brillouin superhigh-frequency optoelectromechanical system based devices through hybrid integration on silicon,” Optica 4, on a slotted photonic crystal cavity,” Applied Physics Let- 847 (2017). ters 101, 221116 (2012). 151. R. Van Laer, C. J. Sarabalis, R. Baets, D. Van Thourhout, and 135. C. Xiong, L. Fan, X. Sun, and H. X. Tang, “Cavity piezoop- A. H. Safavi-Naeini, “Thermal Brillouin noise observed in tomechanics: Piezoelectrically excited, optically trans- silicon optomechanical waveguide,” Journal of Optics 19, duced optomechanical resonators,” Applied Physics Let- 044002 (2017). ters 102, 1–5 (2013). 152. S. R. Mirnaziry, C. Wolff, M. J. Steel, B. J. Eggleton, and 136. L. Fan, X. Sun, C. Xiong, C. Schuck, and H. X. Tang, “Alu- C. G. Poulton, “Stimulated Brillouin scattering in sili- minum nitride piezo-acousto-photonic crystal nanocavity con/chalcogenide slot waveguides,” Optics Express 24, with high quality factors,” Applied Physics Letters 102, 4786 (2016). 153507 (2013). 153. R. Van Laer, B. Kuyken, D. Van Thourhout, and R. Baets, 137. T. P. Purdy, R. W. Peterson, P. L. Yu, and C. A. Regal, “Cav- “Analysis of enhanced stimulated Brillouin scattering in sil- ity optomechanics with Si3N4 membranes at cryogenic icon slot waveguides,” Optics Letters 39, 1242–1245 (2014). temperatures,” (2012). 154. C. Wolff, R. Soref, C. Poulton, and B. Eggleton, “Germa- 138. I. A. Williamson, S. H. Mousavi, and Z. Wang, “Large nium as a material for stimulated Brillouin scattering in the Cavity-Optomechanical Coupling with Graphene at In- mid-infrared,” Optics Express 22, 30735 (2014). frared and Terahertz Frequencies,” ACS Photonics 3, 2353– 155. R. Van Laer and A. H. Safavi-Naeini, “Enabling Strong Cou- 2361 (2016). pling in Nanoscale Silicon Optomechanical Waveguides,” 139. J. D. Teufel, D. Li, M. S. Allman, K. Cicak, a. J. Sirois, J. D. in “Conference on Lasers and Electro-Optics,” (OSA, Wash- Whittaker, and R. W. Simmonds, “Circuit cavity electrome- ington, D.C., 2017), p. FM1F.2. chanics in the strong-coupling regime.” Nature 471, 204– 156. N. T. Otterstrom, R. O. Behunin, E. A. Kittlaus, Z. Wang, 208 (2011). and P. T. Rakich, “A silicon Brillouin laser,” Science 360, 140. A. J. Weinstein, C. U. Lei, E. E. Wollman, J. Suh, A. Metel- 1113–1116 (2018). mann, A. A. Clerk, and K. C. Schwab, “Observation and 157. P. Arrangoiz-Arriola, E. A. Wollack, M. Pechal, J. D. Witmer, Interpretation of Motional Sideband Asymmetry in a Quan- J. T. Hill, and A. H. Safavi-Naeini, “Coupling a Supercon- tum Electromechanical Device,” Physical Review X 4, ducting Quantum Circuit to a Phononic Crystal Defect 041003 (2014). Cavity,” Physical Review X 8, 031007 (2018). 141. E. E. Wollman, C. U. Lei, a. J. Weinstein, J. Suh, A. Kron- 158. R. Van Laer, R. N. Patel, T. P. McKenna, J. D. Witmer, and wald, F. Marquardt, a. a. Clerk, and K. C. Schwab, “Quan- A. H. Safavi-Naeini, “Electrical driving of X-band mechani- tum squeezing of motion in a mechanical resonator,” Sci- cal waves in a silicon photonic circuit,” APL Photonics 3, ence 349, 952–955 (2015). 086102 (2018). 142. J.-M. Pirkkalainen, E. Damskägg, M. Brandt, F. Massel, and 159. M. Goryachev, D. L. Creedon, E. N. Ivanov, S. Galliou, M. A. Sillanpää, “Squeezing of Quantum Noise of Motion R. Bourquin, and M. E. Tobar, “Extremely Low-Loss Acous- in a Micromechanical Resonator,” Physical Review Letters tic Phonons in a Quartz Bulk Acoustic Wave Resonator,” 18

243504, 3–6 (2012). 177. K. Schneider and P. Seidler, “Strong optomechanical cou- 160. Y. Chu, P. Kharel, W. H. Renninger, L. D. Burkhart, L. Frun- pling in a slotted photonic crystal nanobeam cavity with zio, P. T. Rakich, and R. J. Schoelkopf, “Quantum acoustics an ultrahigh quality factor-to-mode volume ratio,” Optics with superconducting qubits,” Science 358, 199–202 (2017). Express 24, 13850 (2016). 161. S. Galliou, M. Goryachev, R. Bourquin, P. Abbé, J. P. Aubry, 178. C. Baker, W. Hease, D.-T. Nguyen, A. Andronico, S. Ducci, and M. E. Tobar, “Extremely Low Loss Phonon-Trapping G. Leo, and I. Favero, “Photoelastic coupling in gallium Cryogenic Acoustic Cavities for Future Physical Experi- arsenide optomechanical disk resonators,” Optics Express ments,” Scientific Reports 3, 2132 (2013). 22, 14072 (2014). 162. A. N. Cleland and M. R. Geller, “Superconducting 179. M.-A. Lemonde, N. Didier, and A. a. Clerk, “Enhanced qubit storage and entanglement with nanomechanical res- nonlinear interactions in quantum optomechanics via me- onators,” Physical Review Letters 93, 1–4 (2004). chanical amplification,” Nature Communications 7, 11338 163. E. A. Kittlaus, N. T. Otterstrom, and P. T. Rakich, “On-chip (2016). Inter-modal Brillouin Scattering,” arXiv:1611.03556 (2016). 180. Z. Wang and A. H. Safavi-Naeini, “Enhancing a slow and 164. R. Pant, D. Marpaung, I. V. Kabakova, B. Morrison, C. G. weak optomechanical nonlinearity with delayed quantum Poulton, and B. J. Eggleton, “On-chip stimulated Brillouin feedback,” Nature Communications 8, 15886 (2017). scattering for microwave signal processing and generation,” 181. A. Xuereb, C. Genes, and A. Dantan, “Strong Coupling Laser Photon. Rev. (2014). and Long-Range Collective Interactions in Optomechanical 165. B. Kuyken, H. Ji, S. Clemmen, S. K. Selvaraja, H. Hu, M. Pu, Arrays,” Physical Review Letters 109, 223601 (2012). M. Galili, P. Jeppesen, G. Morthier, S. Massar, L. K. Ox- 182. C. Gärtner, J. P. Moura, W. Haaxman, R. A. Norte, and enløwe, G. Roelkens, and R. Baets, “Nonlinear properties S. Gröblacher, “Integrated optomechanical arrays of two of and nonlinear processing in hydrogenated amorphous high reflectivity SiN membranes,” arXiv:1809.06372 pp. 1– silicon waveguides.” Optics Express 19, B146–53 (2011). 11 (2018). 166. B. Kuyken, F. Leo, S. Clemmen, U. Dave, R. Van Laer, 183. J. Roels, I. De Vlaminck, L. Lagae, B. Maes, T. Ideguchi, H. Zhao, X. Liu, J. Safioui, S. Coen, S. Gorza, D. Van Thourhout, and R. Baets, “Tunable optical S. K. Selvaraja, S. Massar, R. M. Osgood, P. Verheyen, forces between nanophotonic waveguides,” Nature J. Van Campenhout, R. Baets, W. J. Green, and G. Roelkens, Nanotech. 4, 510–513 (2009). “Nonlinear optical interactions in silicon waveguides,” 184. D. Van Thourhout, T. Spuesens, S. K. Selvaraja, L. Liu, Nanophotonics 0, 1–16 (2016). G. Roelkens, R. Kumar, G. Morthier, P. Rojo-Romeo, F. Man- 167. J. R. Koehler, R. E. Noskov, A. A. Sukhorukov, D. Novoa, dorlo, P. Regreny, O. Raz, C. Kopp, and L. Grenouillet, and P. S. J. Russell, “Coherent control of flexural vibra- “Nanophotonic devices for optical interconnect,” IEEE Jour- tions in dual-nanoweb fiber using phase-modulated two- nal on Selected Topics in Quantum Electronics 16, 1363– frequency light,” (2017). 1375 (2010). 168. E. A. Kittlaus, N. T. Otterstrom, and P. T. Rakich, “On-chip 185. C. Poulton, R. Pant, A. Byrnes, and S. Fan, “Design for Inter-modal Brillouin Scattering,” Conference on Lasers broadband on-chip isolator using stimulated Brillouin scat- and Electro-Optics p. FM1F.2 (2016). tering in dispersion-engineered chalcogenide waveguides,” 169. A. Nunnenkamp, K. Børkje, and S. M. Girvin, “Single- Optics Express 20, 11740–11745 (2012). Photon Optomechanics,” Physical Review Letters 107, 186. K. Fang, J. Luo, A. Metelmann, M. H. Matheny, F. Mar- 063602 (2011). quardt, A. A. Clerk, and O. Painter, “Generalized non- 170. P. Kómár, S. D. Bennett, K. Stannigel, S. J. M. Habraken, reciprocity in an optomechanical circuit via synthetic mag- P. Rabl, P. Zoller, and M. D. Lukin, “Single-photon nonlin- netism and reservoir engineering,” Nature Physics 13, 465– earities in two-mode optomechanics,” Physical Review A 471 (2017). 87, 013839 (2013). 187. M. Tsang, “Cavity quantum electro-optics,” Physical Re- 171. M. Ludwig, A. H. Safavi-Naeini, O. Painter, and F. Mar- view A - Atomic, Molecular, and Optical Physics 81, 1–5 quardt, “Enhanced quantum nonlinearities in a two-mode (2010). optomechanical system,” Physical Review Letters 109, 1–5 188. M. Tsang, “Cavity quantum electro-optics. II. Input-output (2012). relations between traveling optical and microwave fields,” 172. A. Kronwald and F. Marquardt, “Optomechanically In- Physical Review A 84, 043845 (2011). duced Transparency in the Nonlinear Quantum Regime,” 189. C. Wang, M. Zhang, B. Stern, M. Lipson, and M. Lonˇcar, Physical Review Letters 111, 133601 (2013). “Nanophotonic lithium niobate electro-optic modulators,” 173. Y. Okawachi, M. S. Bigelow, J. E. Sharping, Z. Zhu, Optics Express 26, 1547 (2018). A. Schweinsberg, D. J. Gauthier, R. W. Boyd, and A. L. 190. K. Van Acoleyen, J. Roels, P. Mechet, T. Claes, Gaeta, “Tunable all-optical delays via Brillouin slow light D. Van Thourhout, and R. Baets, “Ultracompact Phase in an optical fiber,” Physical Review Letters 153902, 1–4 Modulator Based on a Cascade of NEMS-Operated Slot (2005). Waveguides Fabricated in Silicon-on-Insulator,” IEEE Pho- 174. D. J. Brod and J. Combes, “Passive CPHASE Gate via tonics Journal 4, 779–788 (2012). Cross-Kerr Nonlinearities,” Physical Review Letters 117, 191. A. Pitanti, J. M. Fink, A. H. Safavi-Naeini, J. T. Hill, C. U. 1–6 (2016). Lei, A. Tredicucci, and O. Painter, “Strong opto-electro- 175. D. J. Brod, J. Combes, and J. Gea-Banacloche, “Two photons mechanical coupling in a silicon photonic crystal cavity,” co- and counterpropagating through N cross-Kerr sites,” Optics Express 23, 3196 (2015). Physical Review A 94 (2016). 192. M. Winger, T. D. Blasius, T. P. Mayer Alegre, A. H. Safavi- 176. J. Gea-Banacloche and N. Német, “Conditional phase gate Naeini, S. Meenehan, J. Cohen, S. Stobbe, and O. Painter, using an optomechanical resonator,” Physical Review A - “A chip-scale integrated cavity-electro-optomechanics plat- Atomic, Molecular, and Optical Physics 89, 1–13 (2014). form,” Optics Express 19, 24905 (2011). 19

193. M. Poot and H. X. Tang, “Broadband nanoelectromechan- duced optomechanical resonators,” Applied Physics Let- ical phase shifting of light on a chip,” Applied Physics ters 102, 1–5 (2013). Letters 104, 061101 (2014). 212. K. Alexander, J. P. George, J. Verbist, K. Neyts, B. Kuyken, 194. W. Jin, R. G. Polcawich, P. A. Morton, and J. E. Bowers, D. Van Thourhout, and J. Beeckman, “Nanophotonic Pock- “Phase tuning by length contraction,” Optics Express 26, els modulators on a silicon nitride platform,” Nature Com- 3174 (2018). munications 9, 3444 (2018). 195. C. Errando-Herranz, F. Niklaus, G. Stemme, and K. B. Gyl- 213. P. O. Weigel, J. Zhao, K. Fang, H. Al-Rubaye, D. Trot- fason, “Low-power microelectromechanically tunable sili- ter, D. Hood, J. Mudrick, C. Dallo, A. T. Pomerene, A. L. con photonic ring resonator add-drop filter,” Optics Letters Starbuck, C. T. DeRose, A. L. Lentine, G. Rebeiz, and 40, 3556 (2015). S. Mookherjea, “Hybrid Silicon Photonic-Lithium Niobate 196. T. J. Seok, N. Quack, S. Han, R. S. Muller, and M. C. Wu, Electro-Optic Mach-Zehnder Modulator Beyond 100 GHz “Large-scale broadband digital silicon photonic switches Bandwidth,” arXiv:1803.10365 26, 14810–14816 (2018). with vertical adiabatic couplers,” Optica 3, 64 (2016). 214. A. Siddiqui, R. H. Olsson, and M. Eichenfield, “Lamb Wave 197. D. a. B. Miller, “Self-aligning universal beam coupler.” Op- Focusing Transducer for Efficient Coupling to Wavelength- tics express 21, 6360–70 (2013). Scale Structures in Thin Piezoelectric Films,” Journal of 198. D. A. B. Miller, “Perfect optics with imperfect components,” Microelectromechanical Systems pp. 1–17 (2018). Optica 2, 747–750 (2015). 215. O. Marshall, M. Hsu, Z. Wang, B. Kunert, C. Koos, and 199. J. Carolan, C. Harrold, C. Sparrow, E. Martin-Lopez, N. J. D. Van Thourhout, “Heterogeneous Integration on Silicon Russell, J. W. Silverstone, P. J. Shadbolt, N. Matsuda, Photonics,” Proceedings of the IEEE pp. 1–12 (2018). M. Oguma, M. Itoh, G. D. Marshall, M. G. Thompson, J. C. F. 216. Z. Yu and S. Fan, “Optical isolation based on nonrecipro- Matthews, T. Hashimoto, J. L. O’Brien, and A. Laing, “Uni- cal phase shift induced by interband photonic transitions,” versal linear optics,” Science 349, 711–716 (2015). Applied Physics Letters 94, 171116 (2009). 200. Y. Shen, N. C. Harris, S. Skirlo, M. Prabhu, T. Baehr-Jones, 217. K. Fang, Z. Yu, and S. Fan, “Realizing effective magnetic M. Hochberg, X. Sun, S. Zhao, H. Larochelle, D. Englund, field for photons by controlling the phase of dynamic mod- and M. Soljaˇci´c,“Deep learning with coherent nanopho- ulation,” Nature Photonics 6, 782–787 (2012). tonic circuits,” Nature Photonics 11, 441–446 (2017). 218. Y. Shi, S. Han, and S. Fan, “Optical Circulation and Iso- 201. T. Rudolph, “Why I am optimistic about the silicon- lation Based on Indirect Photonic Transitions of Guided photonic route to quantum computing,” APL Photonics Resonance Modes,” ACS Photonics 4, 1639–1645 (2017). 2, 030901 (2017). 219. D. Lee, K. W. Lee, J. V. Cady, P. Ovartchaiyapong, and 202. P. J. Shadbolt, M. R. Verde, A. Peruzzo, A. Politi, A. Laing, A. C. B. Jayich, “Topical review: spins and mechanics in M. Lobino, J. C. F. Matthews, M. G. Thompson, and J. L. diamond,” Journal of Optics 19, 033001 (2017). O’Brien, “Generating, manipulating and measuring en- 220. M. S. Barson, P. Peddibhotla, P. Ovartchaiyapong, K. Gane- tanglement and mixture with a reconfigurable photonic san, R. L. Taylor, M. Gebert, Z. Mielens, B. Koslowski, D. A. circuit,” Nature Photonics 6, 45–49 (2012). Simpson, L. P. McGuinness, J. McCallum, S. Prawer, S. On- 203. X. Qiang, X. Zhou, J. Wang, C. M. Wilkes, T. Loke, S. O’Gara, oda, T. Ohshima, A. C. Bleszynski Jayich, F. Jelezko, N. B. L. Kling, G. D. Marshall, R. Santagati, T. C. Ralph, J. B. Manson, and M. W. Doherty, “Nanomechanical Sensing Us- Wang, J. L. O’Brien, M. G. Thompson, and J. C. F. Matthews, ing Spins in Diamond,” Nano Letters 17, 1496–1503 (2017). “Large-scale silicon quantum photonics implementing arbi- 221. Y. I. Sohn, S. Meesala, B. Pingault, H. A. Atikian, trary two-qubit processing,” Nature Photonics (2018). J. Holzgrafe, M. Gündogan,ˇ C. Stavrakas, M. J. Stanley, 204. C. Errando-Herranz, N. L. Thomas, and K. B. Gylfason, A. Sipahigil, J. Choi, M. Zhang, J. L. Pacheco, J. Abraham, “Low-power optical beam steering by microelectromechan- E. Bielejec, M. D. Lukin, M. Atatüre, and M. Lonˇcar, “Con- ical waveguide gratings,” arXiv:1809.04483 (2018). trolling the coherence of a diamond spin qubit through 205. D. Jalas, A. Petrov, M. Eich, W. Freude, S. Fan, Z. Yu, its strain environment,” Nature Communications 9, 17–22 R. Baets, M. Popovi´c, A. Melloni, J. D. Joannopoulos, (2018). M. Vanwolleghem, C. R. Doerr, and H. Renner, “What is — 222. R. W. Andrews, R. W. Peterson, T. P. Purdy, K. Cicak, R. W. and what is not — an optical isolator,” Nature Photonics 7, Simmonds, C. A. Regal, and K. W. Lehnert, “Bidirectional 579–582 (2013). and efficient conversion between microwave and optical 206. D. L. Sounas and A. Alù, “Non-reciprocal photonics based light,” Nature Physics 10, 321–326 (2014). on time modulation,” Nature Photonics 11, 774–783 (2017). 223. M. Pechal, P. Arrangoiz-Arriola, and A. H. Safavi- 207. M.-A. Miri, F. Ruesink, E. Verhagen, and A. Alù, “Opti- Naeini, “Superconducting circuit quantum computing with cal Nonreciprocity Based on Optomechanical Coupling,” nanomechanical resonators as storage,” Quantum Science Physical Review Applied 7, 064014 (2017). and Technology 4, 015006 (2018). 208. E. Verhagen and A. Alù, “Optomechanical nonreciprocity,” 224. A. H. Safavi-Naeini, T. P. Mayer Alegre, J. Chan, M. Eichen- Nature Physics 13, 922–924 (2017). field, M. Winger, Q. Lin, J. T. Hill, D. E. Chang, and 209. Q. P. Unterreithmeier, E. M. Weig, and J. P. Kotthaus, “Uni- O. Painter, “Electromagnetically induced transparency and versal transduction scheme for nanomechanical systems slow light with optomechanics.” Nature 472, 69–73 (2011). based on dielectric forces,” Nature 458, 1001–1004 (2009). 225. T. Bagci, A. Simonsen, S. Schmid, L. G. Villanueva, 210. S. A. Tadesse and M. Li, “Sub-optical wavelength acous- E. Zeuthen, J. Appel, J. M. Taylor, A. Sørensen, K. Usami, tic wave modulation of integrated photonic resonators at A. Schliesser, and E. S. Polzik, “Optical detection of radio microwave frequencies,” Nature Communications 5, 5402 waves through a nanomechanical transducer.” Nature 507, (2014). 81–85 (2014). 211. C. Xiong, L. Fan, X. Sun, and H. X. Tang, “Cavity piezoop- 226. A. Rueda, F. Sedlmeir, M. C. Collodo, U. Vogl, B. Stiller, tomechanics: Piezoelectrically excited, optically trans- G. Schunk, D. V. Strekalov, C. Marquardt, J. M. Fink, 20

O. Painter, G. Leuchs, and H. G. L. Schwefel, “Efficient 243. M. K. Ekström, T. Aref, J. Runeson, J. Björck, I. Boström, microwave to optical photon conversion: an electro-optical and P. Delsing, “Surface acoustic wave unidirectional trans- realization,” Optica 3, 597 (2016). ducers for quantum applications,” Applied Physics Letters 227. A. Vainsencher, “Building an Optomechanical Interface for 110, 073105 (2017). Superconducting Qubits,” (2016). 244. I. Mahboob and H. Yamaguchi, “Bit storage and bit flip 228. J. Preskill, “Quantum computing and the entanglement operations in an electromechanical oscillator.” Nature nan- frontier,” arXiv:1203.5813 pp. 1–18 (2012). otechnology 3, 275–9 (2008). 229. J. Preskill, “Quantum Computing in the NISQ era and be- 245. Y. Chu, P. Kharel, T. Yoon, L. Frunzio, P. T. Rakich, and yond,” arXiv:1801.00862 pp. 1–22 (2018). R. J. Schoelkopf, “Climbing the phonon Fock state ladder,” 230. M. H. Devoret and R. J. Schoelkopf, “Superconducting arXiv:1804.07426 (2018). Circuits for Quantum Information: An Outlook,” Science 246. E. J. Davis, Z. Wang, A. H. Safavi-Naeini, and M. H. 339, 1169–1174 (2013). Schleier-Smith, “Painting Nonclassical States of Spin or 231. R. Barends, J. Kelly, A. Megrant, A. Veitia, D. Sank, E. Jeffrey, Motion with Shaped Single Photons,” Physical Review Let- T. C. White, J. Mutus, A. G. Fowler, B. Campbell, Y. Chen, ters 121, 123602 (2018). Z. Chen, B. Chiaro, A. Dunsworth, C. Neill, P. O’Malley, 247. J. Capmany and D. Novak, “Microwave photonics com- P. Roushan, A. Vainsencher, J. Wenner, A. N. Korotkov, A. N. bines two worlds,” Nature Photonics 1, 319–330 (2007). Cleland, and J. M. Martinis, “Superconducting quantum 248. D. Marpaung, B. Morrison, R. Pant, and B. J. Eggleton, “Fre- circuits at the surface code threshold for fault tolerance,” quency agile microwave photonic notch filter with anoma- Nature 508, 500–503 (2014). lously high stopband rejection,” Optics Letters 38, 4300– 232. A. J. Keller, P. B. Dieterle, M. Fang, B. Berger, J. M. Fink, 4303 (2013). and O. Painter, “Al transmon qubits on silicon-on-insulator 249. R. Pant, D. Marpaung, I. V. Kabakova, B. Morrison, C. G. for quantum device integration,” Applied Physics Letters Poulton, and B. J. Eggleton, “On-chip stimulated Brillouin 111, 042603 (2017). Scattering for microwave signal processing and generation,” 233. S. Krinner, S. Storz, P. Kurpiers, P. Magnard, J. Heinsoo, Laser & Photonics Reviews 8, 653–666 (2014). R. Keller, J. Lütolf, C. Eichler, and A. Wallraff, “Engineer- 250. Y. Liu, A. Choudhary, D. Marpaung, and B. J. Eggleton, ing cryogenic setups for 100-qubit scale superconducting “Chip-Based Brillouin Processing for Phase Control of RF circuit systems,” arXiv:1806.07862 (2018). Signals,” IEEE Journal of Quantum Electronics 54, 1–13 234. C. Neill, P. Roushan, K. Kechedzhi, S. Boixo, S. V. Isakov, (2018). V. Smelyanskiy, A. Megrant, B. Chiaro, A. Dunsworth, 251. B. Morrison, D. Marpaung, R. Pant, E. Li, D.-Y. Choi, K. Arya, R. Barends, B. Burkett, Y. Chen, Z. Chen, A. Fowler, S. Madden, B. Luther-Davies, and B. J. Eggleton, “Tunable B. Foxen, M. Giustina, R. Graff, E. Jeffrey, T. Huang, J. Kelly, microwave photonic notch filter using on-chip stimulated P. Klimov, E. Lucero, J. Mutus, M. Neeley, C. Quintana, Brillouin scattering,” Optics Communications 313, 85–89 D. Sank, A. Vainsencher, J. Wenner, T. C. White, H. Neven, (2014). and J. M. Martinis, “A blueprint for demonstrating quan- 252. D. Marpaung, B. Morrison, M. Pagani, R. Pant, D. Choi, tum supremacy with superconducting qubits,” Science 360, B. Davies, S. Madden, and B. Eggleton, “Low-power, chip- 195–199 (2018). based stimulated Brillouin scattering microwave photonic 235. B. Vermersch, P.-O. Guimond, H. Pichler, and P. Zoller, filter with ultrahigh selectivity,” Optica 2, 76–83 (2015). “Quantum State Transfer via Noisy Photonic and Phononic 253. A. Casas-bedoya, B. Morrison, M. Pagani, D. Marpaung, Waveguides,” Physical Review Letters 118, 133601 (2017). and B. J. Eggleton, “Tunable narrowband microwave pho- 236. Z.-L. Xiang, M. Zhang, L. Jiang, and P. Rabl, “Intracity tonic filter created by stimulated Brillouin scattering from Quantum Communication via Thermal Microwave Net- a Silicon nanowire,” Optics Letters 40, 4154–4157 (2015). works,” Physical Review X 7, 011035 (2017). 254. H. Shin, J. a. Cox, R. Jarecki, A. Starbuck, Z. Wang, and 237. H. J. Kimble, “The quantum internet.” Nature 453, 1023– P. T. Rakich, “Control of coherent information via on-chip 1030 (2008). photonic-phononic emitter-receivers,” Nature Communica- 238. M. Pechal and A. H. Safavi-Naeini, “Millimeter-wave in- tions 6, 6427 (2015). terconnects for microwave-frequency quantum machines,” 255. E. A. Kittlaus, P. Kharel, N. T. Otterstrom, Z. Wang, Physical Review A 96, 042305 (2017). and P. T. Rakich, “RF-Photonic Filters via On-Chip Pho- 239. L. Fan, C.-L. Zou, R. Cheng, X. Guo, X. Han, Z. Gong, tonic–Phononic Emit–Receive Operations,” Journal of S. Wang, and H. X. Tang, “Superconducting cavity electro- Lightwave Technology 36, 2803–2809 (2018). optics: a platform for coherent photon conversion between 256. A. Zadok, A. Eyal, and M. Tur, “Gigahertz-wide optically superconducting and photonic circuits,” arXiv:1805.04509 reconfigurable filters using stimulated Brillouin scattering,” (2018). J. Lightw. Technol. 25, 2168–2174 (2007). 240. C. Kieninger, Y. Kutuvantavida, D. L. Elder, S. Wolf, 257. J. Li, H. Lee, T. Chen, and K. J. Vahala, “Characterization H. Zwickel, M. Blaicher, J. N. Kemal, M. Lauermann, S. Ran- of a high coherence, Brillouin microcavity laser on silicon,” del, W. Freude, L. R. Dalton, and C. Koos, “Ultra-high Optics Express 20, 369–373 (2012). electro-optic activity demonstrated in a silicon-organic hy- 258. M. G. Suh, Q. F. Yang, and K. J. Vahala, “Phonon-Limited- brid modulator,” Optica 5, 739 (2018). Linewidth of Brillouin Lasers at Cryogenic Temperatures,” 241. P. Arrangoiz-Arriola and A. H. Safavi-Naeini, “Engineering Physical Review Letters 119, 1–6 (2017). interactions between superconducting qubits and phononic 259. J. Li, M.-G. Suh, and K. Vahala, “Microresonator Brillouin nanostructures,” Physical Review A 94, 063864 (2016). gyroscope,” Optica 4, 346 (2017). 242. M. V. Gustafsson, T. Aref, a. F. Kockum, M. K. Ekstrom, 260. J. Li, H. Lee, and K. J. Vahala, “Microwave synthesizer G. Johansson, and P. Delsing, “Propagating phonons cou- using an on-chip Brillouin oscillator,” Nat. Commun. 4, 1–7 pled to an artificial atom,” Science 346, 207–211 (2014). (2013). 21

261. T. Kippenberg, H. Rokhsari, T. Carmon, A. Scherer, and fibers,” Optics Letters 32, 2006–2008 (2007). K. Vahala, “Analysis of Radiation-Pressure Induced Me- 280. J.-C. Beugnot, S. Lebrun, G. Pauliat, H. Maillotte, V. Laude, chanical Oscillation of an Optical Microcavity,” Phys. Rev. and T. Sylvestre, “Brillouin light scattering from surface Lett. 95, 1–4 (2005). acoustic waves in a subwavelength-diameter optical fibre.” 262. I. S. Grudinin, H. Lee, O. Painter, and K. J. Vahala, “Phonon Nature communications 5, 5242 (2014). laser action in a tunable two-level system,” Phys. Rev. Lett. 281. R. Cohen, Y. London, Y. Antman, and A. Zadok, “Brillouin 104, 83901 (2010). optical correlation domain analysis with 4 millimeter reso- 263. W. C. Jiang, X. Lu, J. Zhang, and Q. Lin, “High-frequency lution based on amplified spontaneous emission,” Optics silicon optomechanical oscillator with an ultralow thresh- Express 22, 12070 (2014). old,” Optics Express 20, 15991 (2012). 282. A. Zarifi, B. Stiller, M. Merklein, N. Li, K. Vu, D.-Y. Choi, 264. L. Fan, C.-L. Zou, M. Poot, R. Cheng, X. Guo, X. Han, P. Ma, S. J. Madden, and B. J. Eggleton, “Highly localized and H. X. Tang, “Integrated optomechanical single-photon distributed Brillouin scattering response in a photonic inte- frequency shifter,” Nature Photonics 10, 766–770 (2016). grated circuit,” APL Photonics 3, 036101 (2018). 265. H. Li, S. A. Tadesse, Q. Liu, and M. Li, “Nanophotonic 283. S. Johnson, M. Ibanescu, M. Skorobogatiy, O. Weisberg, cavity optomechanics with propagating acoustic waves at J. Joannopoulos, and Y. Fink, “Perturbation theory for frequencies up to 12 GHz,” Optica 2, 826 (2015). Maxwell’s equations with shifting material boundaries,” 266. C. L. Zou, X. Han, L. Jiang, and H. X. Tang, “Cavity Phys. Rev. E 65, 1–7 (2002). piezomechanical strong coupling and frequency conver- 284. D. F. Nelson and M. Lax, “New Symmetry for Acousto- sion on an aluminum nitride chip,” Physical Review A 94, Optic Scattering,” Physical Review Letters 24, 379–380 1–7 (2016). (1970). 267. D. E. Chang, a. H. Safavi-Naeini, M. Hafezi, and O. Painter, 285. R. S. Weis and T. K. Gaylord, “Lithium niobate: summary of “Slowing and stopping light with an optomechanical crystal physical properties and crystal structure,” Applied Physics array,” in “New Journal of Physics,” , vol. 1291 (2010), vol. A 203, 191–203 (1985). 1291, pp. 13–17. 286. M. J. A. Smith, C. M. de Sterke, C. Wolff, M. Lapine, and 268. M. Zhang, G. S. Wiederhecker, S. Manipatruni, A. Barnard, C. G. Poulton, “Enhanced acousto-optic properties in lay- P. McEuen, and M. Lipson, “Synchronization of Microme- ered media,” Physical Review B 96, 064114 (2017). chanical Oscillators Using Light,” Physical Review Letters 287. C. Wolff, M. J. Steel, B. J. Eggleton, and C. G. Poulton, 109, 233906 (2012). “Acoustic build-up in on-chip stimulated Brillouin scatter- 269. H. Pfeifer, T. Paraïso, L. Zang, and O. Painter, “Design of ing,” Scientific Reports pp. 1–7 (2015). tunable GHz-frequency optomechanical crystal resonators,” 288. W. H. Renninger, P. Kharel, R. O. Behunin, and P. T. Rakich, Optics Express 24, 11407 (2016). “Bulk crystalline optomechanics,” Nature Physics 14, 601– 270. M. Borselli, T. J. Johnson, and O. Painter, “Measuring the 607 (2018). role of surface in silicon microphotonics,” Ap- 289. P. T. Rakich, M. Popovi´c,and Z. Wang, “General treatment plied Physics Letters 88, 131114 (2006). of optical forces and potentials in mechanically variable 271. T. Asano, Y. Ochi, Y. Takahashi, K. Kishimoto, and S. Noda, photonic systems.” Opt. Express 17, 18116–18135 (2009). “Photonic crystal nanocavity with a Q factor exceeding eleven million,” Optics Express 25, 1769 (2017). 272. L. Fan, X. Sun, C. Xiong, C. Schuck, and H. X. Tang, “Alu- A. THEORETICAL DESCRIPTION minum nitride piezo-acousto-photonic crystal nanocavity with high quality factors,” Applied Physics Letters 102 We first give a description of 3D-confined cavity optomechanics. (2013). Next, we connect to the description of 2D-confined waveguides. 273. H. Liang, R. Luo, Y. He, H. Jiang, and Q. Lin, “High-quality lithium niobate photonic crystal nanocavities,” Optica 4, A. Cavities: 3D-confined 1251 (2017). We focus on cavity optomechanics in this section although much 274. M. Mitchell, D. P. Lake, and P. E. Barclay, “Realizing of it applies to electromechanics as well. The dynamics of an op- Q>300000 in diamond microdisks for optomechanics via tomechanical system involves taking into account the interplay etch optimization,” arXiv:1808.09883 pp. 44–46 (2018). between the coupled acoustic and optical degrees of freedom 275. A. Krause, M. Winger, T. Blasius, Q. Lin, and O. Painter, “A in a system. The frequencies of these two coupled degrees of high-resolution microchip optomechanical accelerometer,” freedom are typically different by many orders of magnitude Nature Photonics 6 (2012). so that the only physical significant coupling arises paramet- 276. K. L. Ekinci, Y. T. Yang, and M. L. Roukes, “Ultimate limits rically in the form described below. As a first step, a modal to inertial mass sensing based upon nanoelectromechanical decomposition of time varying deformations in the elastic struc- systems,” Journal of Applied Physics 95, 2682–2689 (2004). ture is considered, so a set of parameters xj(t), each encoding 277. C. Baker, D. T. Nguyen, W. Hease, C. Gomez, A. Lemaître, the deformation due to a particular vibrational mode is consid- S. Ducci, G. Leo, and I. Favero, “High-frequency nano- ered. A similar decomposition of Maxwell’s equations leads optomechanical disk resonators in liquids,” Nature Nan- to a set of electromagnetic modes of the structure with ampli- otechnology 10, 810–816 (2015). tudes αj(t), which with the correct normalization would lead 2 2 278. P. Tiebot, R. Van Laer, and D. Van Thourout, “Thermal to Uj = h¯ ωo|αj| being the energy and |αj| the average photon tuning of Brillouin resonance in free standing silicon number in mode j. Each optical (mechanical) mode of the struc- nanowire,” in “Conference on Lasers and Electro-Optics,” ture has a frequency ωo (ωm) and their associated dynamics. At (OSA, Washington, D.C., 2018), p. JW2A.35. first we will only consider the interaction between two modes: 279. J.-C. Beugnot, T. Sylvestre, and H. Maillotte, “Guided a single optical and a single mechanical mode of the structure. acoustic wave Brillouin scattering in photonic crystal Optomechanical interactions give rise to coupling between these 22 modes in the following way: the deformation of the structure in The left-hand side of the above equation is simply the equation of a specific vibrational mode parametrized by x, causes a change motion for a damped harmonic oscillator and takes into account in the optical frequency given by ωo(x) = ωo(0) + Gx, where the dynamics of the modal degree of freedom being considered. the optomechanical coupling parameter G = ∂xωo has units The right-hand side of the equation are the forcing terms: FBA(t) −1 of Hz · m . The modal equation for the electromagnetic field, is the optical back-action, while Finput(t) is an input force which under laser excitation at frequency ωin = ωo(0) − ∆ with input we use to understand the linear response of the mechanical 2 photon flux given by |αin| is then expressed as system. The back-action force is given by radiation pressure described via the Maxwell stress tensor, which is quadratic in dα(t)  κ  √ 2 = − i(∆ + Gx) + α − κ α . (28) the field or proportional to |α| . By considering the total energy ex in 2 dt 2 of the system (see section A.3), we find that FBA(t) = hG¯ |α(t)| . The cavity decay rate κ represents the full-width half-maximum Equations 28 and 31 now describe the dynamics of the coupled (FWHM) of the optical mode excitation spectrum and contains system and can be solved to obtain the effects of back-action in all decay channels coupling to the photonic system. Typically, κ the classical domain. In particular, we are primarily interested in the modification of the linear response of the mechanical system consists of an engineered extrinsic decay rate κex as well as the to an input force, i.e. changes to its damping rate and frequency. intrinsic loss rate κin. These changes come about from the mechanical motion x modi- A.1. Linear detection of motion fying the intracavity field α(t) which then applies a force back First, we consider how motion is detected optically in such a onto x which can be proportional, lagging, or leading, leading setup, completely ignoring at first the effect of the light on the to a redefinition of the mechanical system’s complex frequency. mechanical system. We make a few approximations for this To calculate the laser power and frequency dependence of these particular analysis that are useful though not generally valid. modifications, we choose an operating point (α, x) and linearize First we assume that x(t) is slow compared to optical bandwidth the equations of motion by taking to account only the dynamics κ, or equivalently ωm  κ. Also we assume that the laser field is of the fluctuations δx(t) and δα(t). This gives us a set of three driving the photonic cavity on resonance so ∆ = 0, and that the coupled linear differential equations optical decay rate is dominated by the out-coupling so κ = κ . √ ex δx¨(t) = −ω2 δx − γ δx˙ + The output field is then given by α = α + κα, which for m in out in ∗ oscillations that are small, i.e. when the laser-cavity detuning is hG¯ (α δα + c.c.)/meff + Finput(t)/meff (32) √ being modulated by the motion within the linear region of the δα˙ (t) = −(i∆ + κ/2)δα − iGαδx − κexδαin (33) cavity phase response such that Gx  κ, can be solved to obtain ∗ ∗ ∗ √ ? δα˙ (t) = (i∆ − κ/2)δα + iGα δx − κ δα (34) a relation representing phase modulation of the output field: ex in αout = −αin(1 − i4Gx(t)/κ). This is a first order expansion which can be solved for input forces Finput(t) taking δαin = 0 of e−iφ(t) where φ(t) = Gx(t)/κ. An alternate way of writing for now. Solving these equations in the Fourier domain, we ob- tain an expression for the small-signal response of the mechan- this expression is in terms of the intracavity field α(t) which,√ neglecting the mechanical motion, is given by α = −2αin/ κ. ical system to the input force in terms of a dispersion relation, The output field is then δx(ω) ≡ χx(ω)Finput(ω), with p 1 αout = −α − i Γmeasx/xzp (29) χ (ω) = , (35) in x 2 2 meff(ωm − ω − iωγin + Σopt(ω)) with the measurement rate defined as where 2 2 2 2 2 4G xzp|α| 4g |α| 2 2 ∗ Γ ≡ = 0 (30) Σopt(ω) = −ihG¯ |α| (χα(ω) − χα(−ω)) /meff (36) meas κ κ −1 and χα(ω) = (i(∆ − ω) + κ/2) is the optical resonance re- being the rate at which photons are scattered from the laser beam sponse function. The expression in equation 35 represents the to sidebands due to motion of amplitude xzp and an intracavity response of a damped mechanical resonance that is modified by | |2 zero-point optical field intensity of α . The subscript represents , a “self-energy” term, Σopt(ω), due to interaction with optical res- assigned in anticipation of the quantum analysis below though onance. The real and imaginary parts of this self-energy cause an for a classical description xzp can be used to normalize the above effective modification of the mechanical frequency and linewidth expression without changing the physics. The measurement rate ωm and γin. This shift in the complex frequency, often referred as defined here is central to understanding the operation of to as the “optical spring” and “optical damping/amplification” optomechanical systems in the linear regime and will be used effects can be expressed succinctly in terms of Σopt(ω): throughout the text below often denoted alternatively as γOM ≡ ( ) ΓOM ≡ Γmeas. The zero-point coupling rate g0 = Gxzp defined ImΣopt ωm γOM ≈ − , and (37) here is consistent with section3 in the main text. ωm ReΣopt(ωm) A.2. Back-action on the mechanical mode δω ≈ . (38) m 2ω Now we consider how the motion of the mechanical system m is modified due to interaction with the optical field resonating These expression are good approximations in the weak-coupling in the structure. In addition to equation 28, to understand the regime (g  κ + γ). In the strong-coupling regime (g  κ + γ), back-action arising from the interplay between the optical field the full frequency-dependence of the self-energy Σopt(ω) should and mechanical motion, we must consider the dynamics of the be considered. motional degree of freedom: A common mode of operation of optomechanical systems that are sideband-resolved (ωm  κ) is to tune the laser approx- 2 x¨(t) + γinx˙ + ωmx = (FBA(t) + Finput(t))/meff. (31) imately a mechanical frequency to the red side of the optical 23 resonance so ∆ = ωm. In this case, the above relations lead us in the equations above are then physically significant since, e.g., 2 2 2 2 2 † to γOM ≈ 2hG¯ |α| /(κmeffωm) = 4G xzp|α| /κ which is seen the expectation value of h¯ ωjaj aj represents the energy stored in to be equal to the measurement rate calculated in equation 30, the jth mode of the electromagnetic field. We can use the classical though that was for a different regime of operation. The equality expressions for energy in the fields to calculate the energy for of these two rates can be understood as such: with the red- a single photon/phonon above the vacuum state, |ψi = |1ij, detuned scheme of driving, all of the sideband scattering, which which we will then set equal toh ¯ ωj: occurs at rate Γmeas, causes up-shifting of the laser photons into the photonic mode, and thus effectively damps the mechanical |ψi Z h¯ ω ≡ U = hψ| dr Qˆ˙ (r)ρ(r)Qˆ˙ (r)|ψi resonator’s motion. In the above we focused on the effect of j mech the optomechanical interaction on the mechanical resonator’s re- Z −h | r Qˆ˙ (r) (r)Qˆ˙ (r)| i sponse function. However, there are equally important changes vac d ρ vac in the electromagnetic response. These effects, including Bril- Z = 2 ∗( ) ( ) ( ) louin gain/loss and slow/fast light, can be derived similarly 2ωj dr Qj r ρ r Qj r [75]. 2 2 = 2meffωj max[|Qj(r)| ]. (43) A.3. Understanding optomechanical coupling in a nanophotonic sys- tem where we’ve defined the effective mass and zero-point motion In the previous section we studied how optomechanical cou- of the mode to be pling can be used to detect mechanical motion and modify the R ∗ dr Qj (r)ρ(r)Qj(r) linear response of a mechanical resonator. Here we will see m = , and eff,j max[|Q (r)|2] how such an interaction comes about in a realistic nanophotonic j s system. Though a toy model with a one-dimensional scalar h¯ wave-equation and a simplified mass-spring system has long xzpf,j ≡ max[|Qj(r)|] = . (44) 2meff,jωj been used in studying optomechanical systems, obtaining a precise understanding of the coupling rates given nontrivial A similar consideration for the electromagnetic fields leads to wavelength-scale optical and elastic mode profiles requires care- Z ful consideration of the fields and calculation of the interactions. h¯ ωj ≡ Uem = hψ| dr Eˆ (r)e(r)Eˆ (r)|ψi − The goal of this section is to show how we can obtain equa- Z tions similar to equations 32-34 where x(t) and α(t) now repre- hvac| dr Eˆ (r)e(r)Eˆ (r)|vaci sent mode amplitudes for acoustic and optical excitations in a Z nanophotonic device. ∗ = 2 dr ej (r)e(r)ej(r) We start by solving separately the dynamical equations for ∗ electromagnetics and elastodynamics which can be expressed = 2Veffmax[ej (r)e(r)ej(r)]. (45) as eigenvalue equations for the magnetic field H(r) and elastic displacement field Q(r) respectively: with R ∗( ) ( ) ( )  e  dr ej r e r ej r 2 2 0 V = Lhj = ωj hj, L(·) = c curl curl (·) . (39) eff,j ∗ ., and e(r) max[ej (r)e(r)ej(r)] s h¯ ωj ω2Q (r) = LQ (r) max[|ej(r)|] = . (46) j j j 2Veff,jediel λ + µ µ L(·) = − ∇(∇ · (·)) − ∇2(·). (40) ρ ρ Having defined the quantization of the fields and mode nor- malizations, we can now write a Hamiltonian for the optome- The set of solutions of these two equations are the normal elec- chanical system, tromagnetic and acoustic modes of the structure, and define the † † spectrum. Typically, a software package such as COMSOL is H = h¯ ∑ ωjaj aj + h¯ ∑ ωjbj bj +Hint (47) used to obtain these solutions in dielectric structures that don’t j j | {z } | {z } permit analytic analysis. Valid solutions of the electromagnetic Ho Hm and elastic field in the structure can then be expressed as normal mode expansions that can capture the quantum dynamics. The challenge remains calculation of the optomechanical interaction term Hint. We are ˆ ( ) = ( ) ( ) + † † E r, t ∑ ej r aj t h.c., (41) interested in interactions of the type aj ak(bl + bl ) which are the j simplest type that allow energy conservation, assuming that the photonic frequencies for mode j and k are many orders of ˆ ˆ magnitude larger than the mechanical frequency of mode l. For Q(r, t) = ∑ Qk(r)bk(t) + h.c. (42) k the case where j = k, we are considering a shift in the frequency of optical mode j due to a mechanical displacement in mode l. ˙ ˆ with a˙ j(t) = −iωjaj(t), and bk(t) = −iωkbk(t). In defining The relevant interaction energy or rate can be calculated using the quantum field theory, we assign to each mode j a Hilbert first-order cavity perturbation theory. In a dielectric structure space {|nij, n = 0, 1, 2, 3, ..} where |nij is the state representing characterized by e0(r), modifications due to deformations can n photons or phonons in the jth mode. Phonons and photons be taken into account with the expression in each of these Hilbert spaces are then annihilated with the operators aj and bj respectively. The normalizations of ej and Qk e(r) = e0(r) + δe(r). (48) 24

To first order, such a modification of the dielectric causes a shift B. Waveguides: 2D-confined in the optical resonance frequency of a mode with mode profile A theory for 2D-confined waveguides with translational sym- e(r) of metry can be developed similarly to the previous section on 3D-confined cavities. We refer to [89, 92, 93] for a thorough treat- ment from first principles. Here, we focus on connecting the (1) ω0 he|δe|ei ω = − . (49) waveguides’ dynamics to the cavities’ dynamics described in the 2 he|e|ei previous section. For a waveguide the translational symmetry implies that the photonic and phononic eigenproblems can be There are two ways that the dielectric constant changes due to expressed in terms of wavevectors β and K, yielding as solution this deformation. First, a deformation of the optical resonator dispersion relations ω(β) and Ω(K). As shown in [89, 92, 93], boundaries affects the dielectric tensor at the between different the Hamiltonian of the waveguide is H = H + H with the (r) free int materials. This is because the high-contrast step profile of e free Hamiltonian given by across a boundary is shifted by deformations of the structure. By Z Z relating a deformation to a change in the dielectric constant, we H = dβ h¯ ω a† a + dK h¯ ω b† b (54) can use equation 49 to calculate the optomechanical coupling. free β β β K K K Johnson has derived a useful expression [283] for this shift in and the interaction Hamiltonian given by frequency, which when adapted to optomechanics [48], gives a ZZ frequency shift per unit displacement of h¯  †  Hint = √ dβ dK gβ+K aβbK + gβ−K aβbK + h.c. (55) 2π R k 2 −1 ⊥ 2 ω Qn(r)(∆e|e | − ∆(e )|d | )dA in the momentum-description where the three-wave mixing = − 0 gOM,B R 3 (50) † 2 max(|Q|) e(r)|e(r)|2d r interaction rate gβ±K = g0|β±Kαβ±K is proportional to the pump amplitude αβ±K of the mode with wavevector β ± K. Here we for a mechanical vector displacement field Q(r) with component assume phase-matching (∆K = −βp + β ± K = 0 with βp the pump wavevector) and usually consider α to represent a strong Qn(r) normal to the boundary. Secondly, a photoelastic contribu- tion to the optomechanical coupling arises from local changes pump that can be treated classically. in the refractive index due to stress in the material induced by Previous work [75] linked the waveguide’s coupling rate the mechanical deformation. For a particular displacement vec- g0|β±K and Brillouin gain coefficient GB to an optomechanical   cavity’s coupling rate g via a mean-field transition performed tor Q(r) corresponding strain tensor S = 1 ∂ Q + ∂ Q , the 0 ij 2 i j j i on the photonic and phononic equations of motion both in the dielectric perturbation is given by limit of large and small cavity finesse. Here, we derive the same connection but now via a mean-field transition in the large- p · S finesse limit performed directly on the waveguide’s Hamiltonian δe(r) = e · · e, (51) given by expression 55. We consider a cavity of roundtrip length e0 L constructed from a waveguide described by equation 55 and focus on a triplet of two photonic and one phononic mode(s). = − 4 which reduces to δeij e0n pijklSkl for an isotropic medium. The operator corresponding to the number of excitations in each The fourth-rank tensor p with components pijkl is called the pho- of the modes can be expressed as toelastic tensor and is a property of the material. The roto-optic Z k +π/L effect, which captures permittivity changes induced by rota- † c † c c = dk ck ck (56) tion, must be included as well in optically anisotropic materials kc −π/L [284–286]. Composite metamaterials may yield enhanced pho- 2π = c† c toelasticity [286]. Often, when considering the symmetries in L kc kc the atomic structure of the material, a reduced tensor is used with kc a relevant center wavevector. Therefore, the cavity and with elements pij. The perturbation integral can then be used to calculate the shift in frequency per unit displacement: waveguide operators are linked by r 2π R 3 c = ck (57) ω e · δe · e d r L c = − 0 gOM,PE R 3 . (52) 2 max(|Q|) e(r)|e(r)|2d r with c either a photonic or phononic ladder operator. Consider- ing the first term in Hint, we therefore have These expressions give the boundary and photoelastic compo- ZZ  2 nents for a shift in the optical cavity frequency per unit displace- 2π dβ dK gβ+K aβbK = gβ+K aβbK (58) ment of the maximum deflection point of a deformation profile L Q(r). A natural unit for displacement is the zero-point fluctua- 2π = g α δaδb (59) tion amplitude found by multiplying the expressions (50) and L 0|β+K β+K p (52) by the zero-point fluctuation length x = h¯ /(2m ω ) g0|β+K zpf eff m = √ αδaδb (60) (see equation 44). The coupling rate is g0 = g0,Bnd + g0,PE, and L the corresponding optomechanical interaction Hamiltonian can = g αδaδb (61) be written as 0 with δa and δb the cavity’s photonic and phononic ladder opera- † Hint = h¯ (gOM,PE + gOM,Bnd)xa a tors. Thus we obtain g0|β+K † † g0 = √ (62) = hg¯ 0(b + b)a a. (53) L 25 as in equation8 in the main text. This link between the coupling suppresses the interaction in the equations of motion via a rotat- rates holds both for forward and backward as well as for intra- ing term, for instance and inter-modal scattering [75]. (∂ + v ∂ )b = −iω b − ig αa† e−i∆Kcz (71) Next, we consider the waveguide’s dynamics in slightly more t m z m 0|βp detail. The dynamics of an optomechanical waveguide is usually The range of spatiotemporal effects described by these and ex- considered after transforming from momentum- to real-space tended versions of these equations of motion are considered in operators Z detail in amongst others [75, 89, 92, 94, 108, 146, 167, 287, 288]. dk −i(k−kc)z c(z) = e ck (63) 2π B. SINGLE-PHOTON NONLINEARITY We are usually concerned with narrow bandwidths relative to the group-velocity-dispersion so the frequencies can be ex- In this section we give a derivation for relationships 21 and 22. panded to first-order as We consider a 2D-confined waveguide of length L and inject a photon flux hΦi = vg/L with vg the optical group velocity that ωk ≈ ωc + vg(k − kc) (64) corresponds to one photon on average in the waveguide. This photon excites the mechanical system with a displacement x1, A few manipulations [92] starting from equation 54 lead to which in turn yields a phase-shift ϑwg on a second probe photon. Z The phase-shift can be expressed as h † † i Hfree = h¯ dz a (z)ωˆ aa(z) + b (z)ωˆ mb(z) (65) ϑwg = k0(∂xneff)x1L (72) with ωˆ k = ωc − ivg∂z the real-space operator corresponding with k0 the vacuum optical wavevector and ∂xneff the sensitivity to the momentum-space dispersion relation ωk. Higher-order of the effective optical index to mechanical motion x1. Assuming expansions of the dispersion relation in equation 64 yield higher- a static displacement, we have x1 = hFi/keff with hFi the force order spatial derivatives the operator ωˆ k. exerted by the first photon and keff the effective mechanical Further, dropping the second term in equation 55 the real- stiffness per unit length. In addition, from power-conservation 1 space interaction Hamiltonian becomes it can be shown that hFi = c ∂xneff(h¯ ω)hΦi [289]. Substitution Z leads to † h¯ ω2  1 2 Hint = h¯ dz g | + α (z)a(z)b(z) + h.c. (66) 0 β K ϑwg = ∂xneff vg (73) keff c Further, the Heisenberg equations of motion c˙(z) = Since the Brillouin gain coefficient can be expressed as [14] − i [ ( ) H ] h¯ c z , int together with the equal-time commutator  2 † 0 0 Qm 1 [c(z), c (z )] = δ(z − z ) yield GB = 2ω ∂xneff (74) keff c (∂t + vα∂z)α = −iωαα − ig0|β+K a b this yields † h¯ ω GB (∂t + va∂z)a = −iωaa − ig0|β+Kα b (67) ϑwg = vg (75) 2 Qm ( + ) = − − † ∂t vm∂z b iωmb ig0|β+Kα a Making use of equation9 gives These equations describe the spatiotemporal evolution of the 2 2g0|β+K three interacting fields in absence of dissipation and phase- ϑ = (76) wg v mismatch. Intrinsic propagation losses can be included via a gωm dissipative term, e.g. This is the single-photon cross-phase shift for a statically driven mechanical waveguide. When the mechanics is driven with † (∂t + vm∂z)b = −iωmb − γb − ig0|β+Kαa (68) a detuning ∆Ω > γ close to the resonance, one must replace x → (ωm/(2∆Ω))x so the cross-Kerr phase shift increases to −1 1 1 for the phononic field with γ = vmαm with Lm = αm the 2 phononic decay length. Similarly, a non-zero phase-mismatch g0|β+K ∆K 6= 0 effectively reduces the interaction rate. In particular, ϑwg = (77) vg∆Ω dropping the second term for simplicity the momentum-space interaction Hamiltonian 55 becomes which is in agreement with more rigorous analysis [118]. This phase-shift is independent of the waveguide length L since the ZZZ     h¯ † ∆KL photon flux and the optical forces are inversely proportional Hint = √ dβ dK dβp g0|β αβ aβbK L sinc + h.c. 2π p p 2 to length L when there is on average a single photon in the (69) waveguide. L with the waveguide length. Thus the finite length weakens the Next, we link the waveguide single-photon phase-shift ϑwg wavevector-selectivity, generating interactions between a larger to the cavity single-photon phase shift set of modes. The strongest interactions are obtained between modes for which ∆K = 0. This corresponds to the real-space 2g2 ϑ = 0 (78) Hamiltonian cav κ∆Ω Z derived in section5A of the main paper. Making use of equation † i∆Kcz Hint = h¯ dz g α (z)a(z)b(z) e + h.c. (70) 0|β+K 8, the cavity phase-shift ϑcav can be expressed as 2 with ∆Kc = −βpc + βc + Kc the phase-mismatch between the 2g0|β+K center wavevectors of the photonic and phononic fields. This ϑcav = (79) Lrtκ∆Ω 26 with Lrt the cavity roundtrip length. The cavity finesse is F = 2π/(κTrt) with Trt = Lrt/vg the cavity roundtrip time so we obtain F ϑ = ϑ (80) cav π wg The expression for the cavity phase-shift assumed critical cou- pling and a small phase shift given by ϑcav = 2∆/κ with ∆ = ϑwg/Trt the mechanically-induced detuning from the cavity resonance.