arxiv:1907.13058v1 [physics.app-ph] 30 jul 2019the promise of on-chip phononics. the modes of small...

9
Piezoelectric transduction of a wavelength-scale mechanical waveguide Yanni D. Dahmani, * Christopher J. Sarabalis, * Wentao Jiang, Felix M. Mayor, and Amir H. Safavi-Naeini Department of Applied Physics and Ginzton Laboratory, Stanford University 348 Via Pueblo Mall, Stanford, California 94305, USA (Dated: July 31, 2019) We present a piezoelectric transducer in thin-film lithium niobate that converts a 1.7 GHz mi- crowave signal to a mechanical wave in a single mode of a 1 micron-wide waveguide. We measure a -12 dB conversion efficiency that is limited by material loss. The design method we employ is widely applicable to the transduction of wavelength-scale structures used in emerging phononic circuits like those at the heart of many optomechanical microwave-to-optical converters. I. INTRODUCTION Mechanical waves play important roles in classical and quantum microwave systems from low-power, passive fil- tering and signal processing [1–3] to compact resonators for state storage [4–6] and low-loss channels for routing and delays [7–9]. In optical systems, phonons mediate interactions between photons enabling efficient and com- pact modulators [10–15] and full-spectrum beam steer- ing [16] with promising routes to microwave-to-optical conversion for quantum state transfer [17–20]. Emerging applications of on-chip mechanics has led to the development of phononic circuits that utilize compo- nents which have dimensions on the order of the mechan- ical wavelength [21–23]. Efficient and selective transduc- ers of wavelength-scale waveguides are needed to fulfill the promise of on-chip phononics. The modes of small waveguides have been selectively transduced with op- tical forces in cavity optomechanics [7, 8] and stimu- lated Brillouin scattering [24–26]. In these devices, an optical-frequency photon scatters, emitting a microwave- frequency phonon with 10 5 times less energy. Generating a large phonon flux requires impractically large optical power because of this difference in energy scales [27]. Ca- pacitive transducers [28, 29] and piezoelectric transducers convert microwave photons to phonons of the same fre- quency and thus near unity power efficiency conversion can be realized. Piezoelectric transducers have been used to ex- cite wavelength-scale devices including nanobeam res- onators [17, 19, 20, 30] and waveguides [11, 31] in a host of material platforms including aluminum nitride, alu- minum nitride on silicon, galium arsenide, and more re- cently lithium niobate (LN). Some recent approaches to integrated acousto-optic modulation [11] and microwave- to-optical quantum conversion [20, 30] have suffered primarily from inefficient microwave-to-mechanical con- version and thus can benefit from a more systematic, component-based approach to the design and character- ization of efficient, single-mode transducers. One of the difficulties in designing such a device is in coupling a * These authors contributed equally to this work. [email protected] many-wavelength-scale transducer to a waveguide with a much smaller cross-section. As discussed in Section II, the area of these transducers scales inversely with the strength of the piezoelectric coupling which is captured in a parameter called k 2 eff . In suspended LN, resonators [32] and delay line transducers [2, 33] have been demonstrated with k 2 eff on the order of 10%. Compared to previous work in aluminum nitride and galium arsenide where k 2 eff is on the order of 1% [34], suspended LN opens opportu- nities to make more compact transducers enabling new approaches to their design. We present an approach to designing efficient trans- ducers of wavelength-scale guided modes and demon- strate a single-mode transducer for GHz-frequency, hor- izontal shear (SH) waves of a 1 μm-wide rectangular waveguide in suspended, thin-film LN. We restrict our attention to narrow transducers that stay in the low- mode-density limit where normal-mode analysis is prac- ticable, thereby avoiding the challenges of beam forming and horn design [31]. In this work, we find that engi- neering a transducer requires careful consideration of the trade-offs between device footprint, bandwidth, and ef- ficiency. Modeling our transducers with the finite ele- ment method (FEM) [35], we decompose the radiated mechanical waves in the basis of modes of the 1 μm- wide waveguide to find that 96% of the emitted mechan- ical power is into the SH0 mode. We use these mod- els to understand the limitations on the efficiency and bandwidth imposed by loss. We demonstrate compact 260 μm 2 transducers with 10 μm long horns that exhibit an electrical-to-SH0-mode transmission coefficient |t bμ | 2 of -12 dB limited primarily by material loss. Finally we characterize the SH0 mode at 1.7 GHz, finding a group velocity of 4.0 × 10 3 m/s and an attenuation coefficient of α =6.8 dB/mm in close agreement with material losses in other work on LN films [32, 33, 36]. II. TRANSDUCER AREA We begin by considering the footprint A needed to impedance match to a 50 Ω transmission line. Even without detailed knowledge of an IDT’s modes, its foot- print is constrained by the piezoelectric coupling coeffi- arXiv:1907.13058v1 [physics.app-ph] 30 Jul 2019

Upload: others

Post on 28-Mar-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: arXiv:1907.13058v1 [physics.app-ph] 30 Jul 2019the promise of on-chip phononics. The modes of small waveguides have been selectively transduced with op-tical forces in cavity optomechanics

Piezoelectric transduction of a wavelength-scale mechanical waveguide

Yanni D. Dahmani,∗ Christopher J. Sarabalis,∗ Wentao Jiang, Felix M. Mayor, and Amir H. Safavi-Naeini†

Department of Applied Physics and Ginzton Laboratory, Stanford University348 Via Pueblo Mall, Stanford, California 94305, USA

(Dated: July 31, 2019)

We present a piezoelectric transducer in thin-film lithium niobate that converts a 1.7 GHz mi-crowave signal to a mechanical wave in a single mode of a 1 micron-wide waveguide. We measure a-12 dB conversion efficiency that is limited by material loss. The design method we employ is widelyapplicable to the transduction of wavelength-scale structures used in emerging phononic circuits likethose at the heart of many optomechanical microwave-to-optical converters.

I. INTRODUCTION

Mechanical waves play important roles in classical andquantum microwave systems from low-power, passive fil-tering and signal processing [1–3] to compact resonatorsfor state storage [4–6] and low-loss channels for routingand delays [7–9]. In optical systems, phonons mediateinteractions between photons enabling efficient and com-pact modulators [10–15] and full-spectrum beam steer-ing [16] with promising routes to microwave-to-opticalconversion for quantum state transfer [17–20].

Emerging applications of on-chip mechanics has led tothe development of phononic circuits that utilize compo-nents which have dimensions on the order of the mechan-ical wavelength [21–23]. Efficient and selective transduc-ers of wavelength-scale waveguides are needed to fulfillthe promise of on-chip phononics. The modes of smallwaveguides have been selectively transduced with op-tical forces in cavity optomechanics [7, 8] and stimu-lated Brillouin scattering [24–26]. In these devices, anoptical-frequency photon scatters, emitting a microwave-frequency phonon with 105 times less energy. Generatinga large phonon flux requires impractically large opticalpower because of this difference in energy scales [27]. Ca-pacitive transducers [28, 29] and piezoelectric transducersconvert microwave photons to phonons of the same fre-quency and thus near unity power efficiency conversioncan be realized.

Piezoelectric transducers have been used to ex-cite wavelength-scale devices including nanobeam res-onators [17, 19, 20, 30] and waveguides [11, 31] in a hostof material platforms including aluminum nitride, alu-minum nitride on silicon, galium arsenide, and more re-cently lithium niobate (LN). Some recent approaches tointegrated acousto-optic modulation [11] and microwave-to-optical quantum conversion [20, 30] have sufferedprimarily from inefficient microwave-to-mechanical con-version and thus can benefit from a more systematic,component-based approach to the design and character-ization of efficient, single-mode transducers. One of thedifficulties in designing such a device is in coupling a

∗ These authors contributed equally to this work.† [email protected]

many-wavelength-scale transducer to a waveguide with amuch smaller cross-section. As discussed in Section II,the area of these transducers scales inversely with thestrength of the piezoelectric coupling which is captured ina parameter called k2

eff. In suspended LN, resonators [32]and delay line transducers [2, 33] have been demonstratedwith k2

eff on the order of 10%. Compared to previouswork in aluminum nitride and galium arsenide where k2

effis on the order of 1% [34], suspended LN opens opportu-nities to make more compact transducers enabling newapproaches to their design.

We present an approach to designing efficient trans-ducers of wavelength-scale guided modes and demon-strate a single-mode transducer for GHz-frequency, hor-izontal shear (SH) waves of a 1 µm-wide rectangularwaveguide in suspended, thin-film LN. We restrict ourattention to narrow transducers that stay in the low-mode-density limit where normal-mode analysis is prac-ticable, thereby avoiding the challenges of beam formingand horn design [31]. In this work, we find that engi-neering a transducer requires careful consideration of thetrade-offs between device footprint, bandwidth, and ef-ficiency. Modeling our transducers with the finite ele-ment method (FEM) [35], we decompose the radiatedmechanical waves in the basis of modes of the 1 µm-wide waveguide to find that 96% of the emitted mechan-ical power is into the SH0 mode. We use these mod-els to understand the limitations on the efficiency andbandwidth imposed by loss. We demonstrate compact260 µm2 transducers with 10 µm long horns that exhibitan electrical-to-SH0-mode transmission coefficient |tbµ|2of −12 dB limited primarily by material loss. Finally wecharacterize the SH0 mode at 1.7 GHz, finding a groupvelocity of 4.0×103 m/s and an attenuation coefficient ofα = 6.8 dB/mm in close agreement with material lossesin other work on LN films [32, 33, 36].

II. TRANSDUCER AREA

We begin by considering the footprint A needed toimpedance match to a 50 Ω transmission line. Evenwithout detailed knowledge of an IDT’s modes, its foot-print is constrained by the piezoelectric coupling coeffi-

arX

iv:1

907.

1305

8v1

[ph

ysic

s.ap

p-ph

] 3

0 Ju

l 201

9

Page 2: arXiv:1907.13058v1 [physics.app-ph] 30 Jul 2019the promise of on-chip phononics. The modes of small waveguides have been selectively transduced with op-tical forces in cavity optomechanics

2

50 µm

a b

c

dy

xz

yz

y

x z

0 0.2 0.4 0.60

0.5

1

1.5

2

2.5

3

(GH

z)

0

0.5

1

1.5

2

2.5

3

(GH

z)

SH0

SH1

SH2

SH0

SH1

(1/µm)

FIG. 1. Suspended transducers patterned in 300 nm-thick, X-cut LN on silicon are designed to excite the SH0 mode of a1 µm-wide waveguide at 1.7 GHz. They are composed of a 3.4 µm wide, 100 nm thick aluminum IDT and a 10 µm long linearhorn. In the false color SEM c, LN is blue, aluminum is yellow, and the XeF2 release etch front is burgundy. FEM analysisb shows the horn scatters the SH0 mode of the IDT efficiently into the SH0 mode of the 1 µm-wide waveguide. Bands andBloch functions of the IDT and waveguide which constitute the asymptotic state of the horn are plotted at left a and right d,respectively. Waves propagate along y.

cient k2eff [33, 37]

A =π

4

1

ω20csk

2eff

∫dωG (ω) (1)

8

2 G0

ω20cs

γ

k2eff

(for Lorentzian G(ω)) (2)

Here G is the conductance of the IDT (the real part ofthe admittance Y ), cs is its capacitance per unit area,ω0 is its center frequency, and the integral is evaluatedover an interval about ω0. From Equation 2 we see thatmatching to 50 Ω over a large bandwidth comes at thecost of footprint. Materials like LN with high εk2

eff, whereε is the dielectric permittivity, enable small transducerswith large bandwidth.

Our goal is to efficiently excite SH waves in a 300 nmthick, suspended, X-cut LN slab using a transducermatched to a 50 Ω transmission line over a bandwidthof γ = 2π×10 MHz. We choose a γ, taken to be the full-width-half-maximum of the conductance, that is largerthan the limit imposed by material loss. From FEM mod-els of an IDT unit cell, we determine an a = 1.9 µm pitchIDT with fingers parallel to the extraordinary axis ex-cites SH waves at 1.7 GHz and has a capacitance densitycs = 155 µF/m2. We assume a Lorentzian lineshape forG(ω) and k2

eff of 35% [38] leading to an IDT footprint ofA = 250 µm2. To go from our footprint estimate to theprecise dimensions of the transducer, we first considerguided modes of the mechanical waveguide and trans-ducer.

III. THE MODES OF LN WAVEGUIDES:CHOOSING IDT DIMENSIONS

A piezoelectric waveguide with continuous transla-tional symmetry such as the rectangular waveguide inFigure 1d supports a power-orthogonal basis of modesat each frequency ω. These modes solve an eigenvalueproblem on a 2D cross-section of the waveguide in whichthe stress σ and velocity v fields of the theory of elas-ticity and the electrostatic potential Φ of electrostaticsare coupled by the piezoelectric tensor d. The modes|ψm〉 ≡ (σm,vm,Φm), indexed by m, vary along thewaveguide as eiKmy for complex eigenvalue Km. IfKi 6= K∗j , modes i and j are power-orthogonal and satisfy

〈ψi| ψj〉 ≡ −∫

d~S · (σ∗i vj + σjv∗i + iωD∗iΦj − iωDjΦ

∗i )

= 0 (3)

forming an inner-product space in which band structurescan be computed and scattering can be studied [39]. HereD = −ε∇Φ + dσ is the electric displacement field, andwe normalize our basis such that 〈ψi| ψj〉 = δij . For moredetail on our choice of Fourier conventions and the rela-tionship between the inner-product and power see Ap-pendix A.

The wavelength-scale LN waveguide we are trying toexcite is 300 nm thick and 1 µm wide, and supports fourmodes between 0 and 1.6 GHz: the Lamb (A0), the shear(SH0), the first excited Lamb (A1), and the longitudinal(S0) mode [40]. The band structure is plotted in Fig-ure 1d. For X-cut LN with crystal Z transverse to thewaveguide, the shear mode couples to the electric field

Page 3: arXiv:1907.13058v1 [physics.app-ph] 30 Jul 2019the promise of on-chip phononics. The modes of small waveguides have been selectively transduced with op-tical forces in cavity optomechanics

3

component Ey by the largest component of the piezo-electric tensor dYZY = 68 pC/N [41, 42]. For this reasonwe focus our attention on the shear bands.

In these waveguides the number of modes increaseswith width and the first two SH modes become degen-erate. For ω = 2π × 1.7 GHz, we numerically solve forthe wavevectors K of the modes of a 2D cross sectionof the waveguide. At widths greater than 6 µm (Fig-ure 2), the SH0 and SH1 modes localize to the edge,decaying exponentially away from the edges analogousto Rayleigh waves. These waves limit continuously todegenerate antisymmetric and symmetric supermodes ofthe edge states, respectively.

As shown in Figure 2a, an adiabatic horn scatters theSH0 mode of a narrow waveguide to the antisymmetricsupermode of a wide waveguide. Since the edge statesare bound to the edge, their coupling to a wide IDTis limited. A wide IDT will excite modes with a sig-nificant amount of energy in the center of the waveg-uide. Mechanical power localized in the center of theIDT must either reflect off an adiabatic horn or scat-ter into modes other than the SH0 mode of the narrowwaveguide. Therefore, for a wide IDT to efficiently excitethe SH0 mode of the waveguide, the horn would need tomediate a non-adiabatic transition.

We choose 3.4 µm for the width of the IDT so thatan adiabatically tapered horn can efficiently scatter thetransduced mode into the 1 µm waveguide. The narrowIDT allows us to simplify the design, make full use of thewidth of the transducer, spectrally resolve the SH0 andSH1 modes, and keep spurious shear modes in cutoff. Wedesign a 3.4 µm-wide transducer and a horn from 3.4 µmto 1 µm that efficiently scatters the output of the IDTinto a single mode of the waveguide.

IV. DESIGNING THE IDT AND HORN

From our footprint estimate in Section II, we expecta 3.4 µm × 74 µm IDT to match to 50 Ω correspondingto a N = 39 finger-pair transducer. As we will see insimulation and experiment, material loss dominates theresponse at this length, reducing transmission from mi-crowaves in the 50 Ω line to phonons in the SH0 modeof the waveguide. We focus our attention on a 40 finger-pair IDT and, after discussing its performance, analyzethe relationship between the transmission tbµ and N .

Our waveguide transducer consists of three cascadedcomponents, a 3.4 µm-wide transducer, a 1 µm-wide out-put waveguide with modes analyzed in Section III, and alinear taper which connects the two. A good transducerhas a conductance reaching 20 mS (matching to 50 Ω)and radiates efficiently into a single-mode of the outputwaveguide.

By orienting the IDT perpendicular to the crystal Zaxis, we can leverage the dYZY component of the piezo-electric tensor to efficiently excite the SH waves of theslab. The 100 nm aluminum electrodes mechanically and,

(GHz)1.6 1.62 1.64 1.66 1.68 1.7 1.72 1.74 1.76 1.78

-25

-20

-15

-10

-5c

yzb

2 4 6 8 10 12 140.3

0.35

0.4

0.45

0.5

0.55

0.6

0.65

0.7

0.75

0.8

Waveguide width (µm)

(1/µ

m)

a

y

x z

Mod

e fr

actio

n (d

B)

FIG. 2. a. The number of modes supported by a suspendedLN waveguide increases with width. The SH0 and SH1 modes(uz plotted) limit to degenerate antisymmetric and symmet-ric edge supermodes, respectively. Below 5 µm the SH waves(red) are well-resolved. The lamb waves are plotted in lightblue. b. The linear horn scatters the SH0 mode of the3 µm-wide IDT efficiently into SH0 of the 1 µm-wide out-put waveguide. c. Decomposing the power in the waveguidewe find that transduction of spurious modes, the isolation,is no greater than −10 dB away from the nodes in the con-ductance over a 200 MHz bandwidth. The largest spuriouscomponent, the SH1 mode, is plotted in red.

for high k2eff devices [33, 43], electrically load the free slab

of Section III. The band diagrams and mode plots shownin Figure 1a take this loading into account. The 1.92 µm-pitch IDT has a duty cycle of 50% with fingers that end300 nm away from the 400 nm wide bus wires that runalong the edges of the waveguide. The 1.7 and 2.2 GHzSH0 and SH2 Γ-point modes of the IDT are efficientlytransduced as seen in the measured S11 and conductanceG (ω) (Figure 4b and c). The symmetric 1.8 GHz SH1mode is weakly transduced due to symmetry.

A 3D FEM model of the transducer, IDT, and taper,is used to compute both the linear response of the systemand analyze the waves emitted. A solution on the full do-main is shown in Figure 1b. A perfectly-matched layer

Page 4: arXiv:1907.13058v1 [physics.app-ph] 30 Jul 2019the promise of on-chip phononics. The modes of small waveguides have been selectively transduced with op-tical forces in cavity optomechanics

4

10 20 30 40N

0

20

40

60

80 c

0

40

60

80

100a

20

(%)

0

20

40

60

80

100 b

(%)

(%)

FIG. 3. By computing the linear response Y (ω) and decom-position am we study the N -dependence of the transmissiontbµ from a 50 Ω transmission line to the SH0 mode. a. As Nincreases so too does the peak conductance. This decreasesthe microwave reflections S11. b. Increasing N also decreasesγe and when N is large such that γi γe, material dampingdominates the response. This is seen in the fractions of thetotal power dissipated due to loss and due to transmissioninto SH0, η. c. These competing effects lead to an optimalN for maximizing |tbµ|2 which increases with Qi. Numericalresults (points) are smoothed with a moving window average(curve) for clarity.

bounds the domain. Reflections at the interface betweenthe IDT and taper give rise to resonances which pro-portionally decrease the full-width-half-maximum and in-crease the peak height of G (ω). In order to match theresonant response to our measurements we incorporate auniform material loss tangent corresponding to Qi = 300and scale the piezoelectric tensor from its bulk values by0.67 [33]. The simulated and measured conductance ofthe SH0 mode is plotted in Figure 4c for comparison.For the SH0 mode, the peak conductance of 6.9 mS andbandwidth (full-width-half-max) of 7.3 MHz are in goodagreement with measurements, 6.5 mS and 9.7 MHz.

The modes of the 1 µm waveguide studied in Sec-tion III form a basis (see Appendix B) in which we candecompose the power radiated by the IDT and check thatthe transducer excites a single mode. Given a solution|ψ〉, the coefficients am are computed using the innerproduct defined in Equation 3

am = 〈ψm| ψ〉 (4)

such that

|ψ〉 =∑m

am |ψm〉 (5)

where |am|2 is the power in mode m. For each modem there’s an associated backwards propagating mode−m, the pair of which form a piezoelectric port (see Ap-pendix C).

From the band diagram in Figure 2a, we conclude thata horn consisting of a 10 µm linear taper functions ap-proximately adiabatically. The power transmitted intothe waveguide is transmitted into the SH0 mode with−10 dB total power in spurious modes (labeled isola-tion in Figure 2c). Excluding nodes in the conductance(Figure 4c), power in the next most strongly transducedmode (SH1) remains below −15 dB over 200 MHz. Thisindicates that the 10 µm long linear taper efficiently scat-ters the SH0 mode of the 3.4 µm waveguide into the SH0mode of the 1 µm waveguide.

Of the total power dissipated by an N = 40 transducer,2G (ω) |V (ω)|2, 11% is emitted into the waveguide at thecenter frequency, 96% of which is in the SH0 mode. Only5% is lost to clamping while the other 84% is lost tomaterial damping.

We are interested in how this transducer behaves whencoupled to a 50 Ω transmission line. In particular, we areinterested in the transmission coefficient tbµ from mi-crowaves in the transmission line to phonons in the SH0mode and how it varies withN . The relationship betweenthe FEM analysis and components of the S-matrix liketbµ is detailed in Appendix C. In our previous work [33],we’ve shown that as N is increased, the extrinsic couplingγe drops as N−3 for strongly resonant IDTs like thosereported here. In the high N -limit where γe γi, thebandwidth γ approaches the material limit γi = ω0/Qi.And while, in this limit, increasing N continues to in-crease the conductanceG thus decreasing reflections fromimpedance mismatch (Figure 3a), this power is lost todamping. In this regime, increasing N seems to improveS11, but actually reduces the microwave-to-mechanicaltransmission. For this reason, tbµ cannot be extractedfrom S11. In Section V, tbµ is instead determined from ameasurement of S21. In Figure 3b we sweep N and com-pute the percentages of the total power dissipated by theIDT transmitted into the SH0 mode and lost to dampingand see in Figure 3c that for Qi = 300, tbµ reaches 12%at N = 29.

There are a few approaches to improve |tbµ|2 beyond12%. The most obvious is to improve the material pa-rameters k2

eff and Qi (see Figure 3). For applicationsin quantum science, operating at cryogenic temperatureswill likely increase Qi by suppressing thermally inducedmechanical loss and ohmic dissipation in the electrodes.Another strategy is to reduce the reflection coefficientat the IDT-waveguide interface thus increasing γe for agiven N and allowing us to make longer transducers be-fore reaching loss-limits. Lastly we could diverge from thelow width, low density of states design and employ wider

Page 5: arXiv:1907.13058v1 [physics.app-ph] 30 Jul 2019the promise of on-chip phononics. The modes of small waveguides have been selectively transduced with op-tical forces in cavity optomechanics

5

waveguides, embracing the challenges of multi-mode de-sign [31].

Time (ns)0 200 400 600 800 1000 1200 1400 1600

-140

-120

-100

-80

-60

1L 3L 5L 7LRound Trip

-100

-80

|h| (

dB H

z)

=11.6 dB/RT

|h| (

dB H

z)

e

1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0-120

-100

-80

-60

-40

-20

0

S-pa

ram

eter

s (dB

)

(GHz)

d

1.68 1.7 1.72-80

-40

c

0 1 2 3 4 5 6 7 8 9 10

-6

-4

-2

0

|S 11| (

dB)

(GHz)

1.4 1.6 1.8 2.0 2.2 2.4

-6

-4

-2

0

1.6 1.64 1.68 1.72 1.76 1.8

-2

0

2

4

6

|S 11| (

dB)

(GHz) (GHz)

Adm

ittan

ce (m

S)

a

b

SH0 SH2

FIG. 4. a. The |S11| of an N = 40 transducer. b. The|S11| restricted to the SH0 and SH2 responses. c. The mea-sured conductance G and susceptance χ of the SH0 modeare overlaid on FEM results. The SH2 mode is also stronglytransduced with a peak conductance of 5.6 mS and full-width-half-max of 8.2 MHz. d. The |S21| of a delay line with aL = 200 µm waveguide shows large SH0 and SH2 responsesat 1.7 and 2.2 GHz. For the SH2, of the −21.1 dB insertionloss, we attribute −1.7 dB to impedance mismatch. The |S21|of an ideal delay line with no insertion loss would equal thetwo-port mismatch, 1−|S11|2/2−|S22|2/2 (see Appendix E).e. For L = 800 µm, the heights of the echoes in the impulseresponse are fit (inset) to extract a round trip loss of 11.6 dB.The S21 is filtered in the time-domain with rectangular en-velopes (shaded blue, red, and green) to compute the single,triple, and quintuple-transit S21 plotted with correspondingcolors (inset). The peak single-transit response is used as an

estimate of |tbµ|2 e−αL/2.

V. MEASUREMENTS

Starting with a 500 nm-thick film of LN on a 500 µm-thick silicon substrate, the film is thinned to 300 nm byargon milling before patterning an HSQ mask with e-beam lithography to define the waveguides. The mask istransferred to the LN by angled argon milling [30]. Wethen perform an acid clean to remove resputtered, amor-phous LN. We deposit 100 nm of Al for electrodes and200 nm Al for contact pads by e-beam lithography andphotolithography, respectively; metal evaporation; andliftoff. Finally we release the structures with a maskedXeF2 dry etch.

The S-parameters of the transducers are measuredwith a vector network analyzer (Rhode & SchwarzZNB20) on a probe station calibrated to move the ref-erence plane to the tips of the probes (GGB nickel 40A).Several modes below 10 GHz are strongly transduced asseen in the S11 plotted in Figure 4. The conductanceG ≡ ReY and susceptance χ ≡ ImY for the SH0 modeplotted in Figure 4c match well with the overlaid simu-lated curve and Γ-point frequency of the IDT unit cellbands shown in Figure 1a. The peak conductance andfull-width-half-max for the SH0 is 6.5 mS and 9.7 MHz,inferred by Lorentzian fit. We infer a static capacitanceof 31 fF by fit to the DC response of χ and use it alongwith the conductance fit by Equation 2 to calculate a k2

effof 15%. This is decreased by the feedthrough capacitanceof the contact pads.

For an L = 200 µm long waveguide, the minimumloss through the structure containing a transducer, ta-per, waveguide, taper, and another transducer is |S21| =−15.7 dB at ω = 2π × 1.7 GHz (Figure 4d). We at-tribute −1.5 dB to impedance mismatch from the mea-sured S11 and S22. Reflections at the IDT-waveguideinterface resonantly enhance transmission through thewaveguide which is why the peak |S21| is significantlylarger than |tbµ|4.

The peak of |S21| encompasses all loss mechanismsin the device including clamping, damping, excitationof spurious modes, and propagation loss and is affectedby reflections and resonance. We can isolate the prop-agation loss α and tbµ by analyzing the time-domainimpulse response h (t), the inverse Fourier transform ofS21(ω), plotted for a device with L = 800 µm (Figure 4e).The first pulse takes the shortest path through the de-vice and is attenuated by |tbµ|2e−αL/2. Each subsequentecho takes an additional round trip, is attenuated by|r|2e−αL, and delayed by 2L/vg = 4.0 × 102 ns. Wefit |r|2e−αL = −11.6 dB from the peaks in Figure 4e andtransform the first pulse (blue) back to the frequencydomain (inset) to find |tbµ|2e−αL/2 = −28.6 dB. Moredetail is provided in the Appendix D.

The single-transit and round-trip loss are two con-straints on three unknown quantities: |tbµ|2, |r|2, andα. By sweeping the length of the device, all three pa-rameters can be determined independently. In lieu of alength sweep, we ignore scattering into other modes and

Page 6: arXiv:1907.13058v1 [physics.app-ph] 30 Jul 2019the promise of on-chip phononics. The modes of small waveguides have been selectively transduced with op-tical forces in cavity optomechanics

6

assume |tbµ|2 + |r|2 = 1 at the IDT-waveguide interfaceto find a |tbµ|2 of 7.0% (comparable to the simulatedvalue of 8.9% for N = 40), an |r|2 of 93%, and an α of6.8 dB/mm. If an additional loss channel of 10% (50%) isadded, |tbµ|2 decreases to 6.6% (4.9%) while α decreasesto 6.3 dB/mm (2.9 dB/mm). Given the measured groupvelocity of vg = 4.0 × 103 m/s, this α corresponds to aquality factor Q = ω0/αvg of 1700 in the waveguide andan f0Q of 2.9×1012 which is comparable to our previouswork in multi-moded, high frequency delay lines with anf0Q of 4.6× 1012 [33]. We see an order of magnitude im-provement over delay lines in suspended LN employingthe S0 mode at 350 MHz where f0Q = 0.45 × 1012 [36].Resonators using antisymmetric thickness modes exhibitf0Q products over twice as large (9.15× 1012) [32].

VI. CONCLUSIONS

In the fields of optomechanics and Brillouin scatter-ing, interaction rates increase as the mechanical modedecreases in size. Whether photon-phonon interactionsare being used to make compact acousto-optic mod-ulators [10, 11, 13] or as an approach to quantummicrowave-to-optical links [17–20, 30], there’s a commonneed to efficiently and selectively transduce the modes ofwavelength-scale mechanical structures.

Our hope is that insights from our approach to address-ing this problem, here in suspended LN, can be generally

applied to exciting wavelength-scale mechanical devices.The explicit connection between footprint and bandwidthconstrains the design space. We show that in the presenceof loss, long IDTs have limited transmission coefficientswhich may improve in cryogenic environments. In thepursuit of efficient, broadband transducers at room tem-perature, our results motivate mitigating reflections atthe waveguide-IDT interface and engineering horns forwider, multi-moded transducers. Efficient, single-modetransducers that are ready to incorporate into complexphononic networks fundamentally advance our controlover mechanical degrees of freedom.

ACKNOWLEDGEMENTS

The authors would like to thank Rishi N. Patel, Patri-cio Arrangoiz-Arriola, and Timothy P. McKenna for use-ful discussions. This work was supported by a MURIgrant from the U. S. Air Force Office of Scientific Re-search (Grant No. FA9550-17-1-0002), by a fellowshipfrom the David and Lucille Packard foundation, and bythe National Science Foundation through ECCS-1808100and PHY-1820938. Part of this work was performed atthe Stanford Nano Shared Facilities (SNSF), supportedby the National Science Foundation under Grant No.ECCS-1542152, and the Stanford Nanofabrication Facil-ity (SNF).

[1] K. Lakin, G. Kline, and K. McCarron, IEEE Trans-actions on Microwave Theory and Techniques 41, 2139(1993).

[2] T. Manzaneque, R. Lu, Y. Yang, and S. Gong, in 2017IEEE 30th International Conference on Micro ElectroMechanical Systems (MEMS) (IEEE, 2017) pp. 155–158.

[3] A. Y. Cleland, M. Pechal, P.-J. C. Stas, C. J. Sarabalis,and A. H. Safavi-Naeini, (2019), arXiv:1905.08403.

[4] K. J. Satzinger, Y. P. Zhong, H.-S. Chang, G. A. Peairs,A. Bienfait, M.-H. Chou, A. Y. Cleland, C. R. Conner,E. Dumur, J. Grebel, I. Gutierrez, B. H. November, R. G.Povey, S. J. Whiteley, D. D. Awschalom, D. I. Schuster,and A. N. Cleland, Nature 563, 661 (2018).

[5] P. Arrangoiz-Arriola, E. A. Wollack, Z. Wang, M. Pechal,W. Jiang, T. P. McKenna, J. D. Witmer, R. Van Laer,and A. H. Safavi-Naeini, Nature 571, 537 (2019).

[6] M. V. Gustafsson, T. Aref, A. F. Kockum, M. K. Ek-strom, G. Johansson, and P. Delsing, Science 346, 207(2014).

[7] K. Fang, M. H. Matheny, X. Luan, and O. Painter, Na-ture Photonics 10, 489 (2016).

[8] R. N. Patel, Z. Wang, W. Jiang, C. J. Sarabalis, J. T.Hill, and A. H. Safavi-Naeini, Physical Review Letters121, 040501 (2018).

[9] T. Manzaneque, R. Lu, Y. Yang, and S. Gong, IEEETransactions on Microwave Theory and Techniques 67,1379 (2019).

[10] L. Fan, C.-L. Zou, N. Zhu, and H. X. Tang, NaturePhotonics 13, 323 (2019).

[11] Q. Liu, H. Li, and M. Li, Optica 6, 778 (2019).[12] S. A. Tadesse and M. Li, Nature Communications 5, 5402

(2014).[13] D. B. Sohn, S. Kim, and G. Bahl, Nature Photonics 12,

91 (2018).[14] M. Merklein, B. Stiller, K. Vu, S. J. Madden, and B. J.

Eggleton, Nature Communications 8, 574 (2017).[15] C. G. Poulton, R. Pant, and B. J. Eggleton, Journal of

the Optical Society of America B 30, 2657 (2013).[16] C. J. Sarabalis, R. Van Laer, and A. H. Safavi-Naeini,

Optics Express 26, 22075 (2018).[17] A. Vainsencher, K. J. Satzinger, G. A. Peairs, and A. N.

Cleland, Applied Physics Letters 109, 033107 (2016).[18] J. Bochmann, A. Vainsencher, D. D. Awschalom, and

A. N. Cleland, Nature Physics 9, 712 (2013).[19] K. C. Balram, M. I. Davanco, B. R. Ilic, J.-H. Kyhm,

J. D. Song, and K. Srinivasan, Physical Review Applied7, 024008 (2017).

[20] M. Forsch, R. Stockill, A. Wallucks, I. Marinkovic,C. Gartner, R. A. Norte, F. van Otten, A. Fiore, K. Srini-vasan, and S. Groblacher, (2018), arXiv:1812.07588.

[21] R. H. Olsson III and I. El-Kady, Measurement Scienceand Technology 20, 012002 (2009).

[22] S. Mohammadi and A. Adibi, AIP Advances 1, 041903(2011).

Page 7: arXiv:1907.13058v1 [physics.app-ph] 30 Jul 2019the promise of on-chip phononics. The modes of small waveguides have been selectively transduced with op-tical forces in cavity optomechanics

7

[23] W. Fu, Z. Shen, Y. Xu, C.-L. Zou, R. Cheng, X. Han, andH. X. Tang, Nature Communications 10, 2743 (2019).

[24] R. Van Laer, B. Kuyken, D. Van Thourhout, andR. Baets, Nature Photonics 9, 199 (2015).

[25] E. A. Kittlaus, H. Shin, and P. T. Rakich, Nature Pho-tonics 10, 463 (2016).

[26] M. S. Kang, A. Brenn, and P. St.J. Russell, PhysicalReview Letters 105, 153901 (2010).

[27] A. H. Safavi-Naeini, D. Van Thourhout, R. Baets, andR. Van Laer, Optica 6, 213 (2019).

[28] E. Romero, R. Kalra, N. Mauranyapin, C. Baker,C. Meng, and W. Bowen, Physical Review Applied 11,064035 (2019).

[29] R. Van Laer, R. N. Patel, T. P. McKenna, J. D. Wit-mer, and A. H. Safavi-Naeini, APL Photonics 3, 086102(2018).

[30] W. Jiang, R. N. Patel, F. M. Mayor, T. P. McKenna,P. Arrangoiz-Arriola, C. J. Sarabalis, J. D. Witmer, R. V.Laer, and A. H. Safavi-Naeini, Optica 6, 845 (2019).

[31] A. Siddiqui, R. H. Olsson, and M. Eichenfield, Journalof Microelectromechanical Systems 27, 1054 (2018).

[32] Y. Yang, R. Lu, T. Manzaneque, and S. Gong, in2018 IEEE/MTT-S International Microwave Symposium- IMS (IEEE, 2018) pp. 563–566.

[33] C. J. Sarabalis, Y. D. Dahmani, A. Y. Cleland, and A. H.Safavi-Naeini, (2019), arXiv:1904.04981.

[34] M. Wu, E. Zeuthen, K. C. Balram, and K. Srinivasan,1, 1 (2019), arXiv:1907.04830.

[35] “Comsol Multiphysics 5.4,” https://www.comsol.com.

[36] G. Vidal-Alvarez, A. Kochhar, and G. Piazza, in2017 IEEE International Ultrasonics Symposium (IUS)(IEEE, 2017) pp. 1–4.

[37] C. J. Sarabalis and A. H. Safavi-Naeini, Manuscript inpreparation (2019).

[38] I. Kuznetsova, B. Zaitsev, S. Joshi, and I. Borodina,IEEE Transactions on Ultrasonics, Ferroelectrics andFrequency Control 48, 322 (2001).

[39] B. A. Auld, Acoustic fields and waves in solids, VolumeII, 2nd ed. (Robert E. Krieger Publishing Company, Mal-abar, Florida, 1990).

[40] B. A. Auld, Acoustic fields and waves in solids, VolumeI, 2nd ed. (Robert E. Krieger Publishing Company, Mal-abar, Florida, 1990).

[41] R. S. Weis and T. K. Gaylord, Applied Physics A Solidsand Surfaces 37, 191 (1985).

[42] A. S. Andrushchak, B. G. Mytsyk, H. P. Laba, O. V.Yurkevych, I. M. Solskii, A. V. Kityk, and B. Sahraoui,Journal of Applied Physics 106, 073510 (2009).

[43] K.-y. Hashimoto, Surface Acoustic Wave Devices inTelecommunications (Springer Berlin Heidelberg, Berlin,Heidelberg, 2000).

[44] D. M. Pozar, Microwave Engineering, 4th ed. (John Wi-ley & Sons, Inc., Hoboken, New Jersey, 2012).

Appendix A: Power Dissipation in the FourierDomain

The voltage V (t) can be expressed in the frequencydomain

V (ω) =1

∫ ∞−∞

dωV (t)eiωt (A1)

and similarly for current I and admittance Y . With ourchoice of Fourier convention, the power dissipated by anelectrical element P(t) = V (t)I(t) is the convolution ofV (ω) and I(ω) in the frequency domain. Averaging thisquantity in time extracts the DC component of the spec-trum thus reducing the convolution to

〈P〉 =

∫ ∞−∞

dωV (ω)I(−ω) (A2)

Since the voltage is a real-valued quantity, V ∗(ω) isequal to V (−ω) (the same argument holds for I(ω) andY (ω)). Changing our limits of integration and usingOhm’s law, I(ω) = Y (ω)V (ω), we find

〈P〉 =

∫ ∞0

dωY ∗(ω) |V (ω)|2 + Y (ω) |V (ω)|2 (A3)

Since 2Re(Y )(ω) = 2G(ω) = Y (ω) + Y ∗(ω), the timeaverage power dissipated by the electrodes is

P0 = 2

∫ ∞0

dωG(ω) |V (ω)|2 . (A4)

To determine the time-average power for a piezoelectricwave, we repeat the previous analysis starting from theinstantaneous piezoelectric Poynting vector

Ppiezo(t) = −σ(t)v(t) + Φ(t)∂tD(t) (A5)

to find the time-average piezoelectric power

Ppiezo = −∫ ∞

0

∫d~S ·(σ∗v+σv∗+iωD∗Φ−iωDΦ∗).

(A6)We compare this expression to the inner product in Equa-tion 3 to confirm |am|2 is the time-average power in modem. We note that our time-average power differs from thatof Auld’s by a factor of 1/4 [39] resulting from differingFourier conventions. All values reported are power ratiosand thus factors of 2 from choices of convention drop out.

Appendix B: Basis

Decomposition of the mechanical energy radiated intoa waveguide is necessary for calculating transmission co-efficients like tbµ needed to characterize a phononic com-ponent. For completeness we briefly describe the basisof propagating modes in a 300 nm thick, 1 µm-wide,X-cut LN, rectangular waveguide. We categorize thefive 1.7 GHz modes as Lamb (A), horizontal shear (SH),and longitudinal (S) modes which differ in their principalstrains Sxz, Syz, and Szz, respectively. These modes areplotted in Figure 5 along with their reflection symme-tries (σz, σx) where (+,−), for example, means symmet-ric and antisymmetric with respect to reflection acrossthe xy and yz-planes, respectively.

Page 8: arXiv:1907.13058v1 [physics.app-ph] 30 Jul 2019the promise of on-chip phononics. The modes of small waveguides have been selectively transduced with op-tical forces in cavity optomechanics

8

zx

FIG. 5. Modes of an LN waveguide at 1.7 GHz. Color inthese plots visualize the dominant displacement field. Lightblue arrows show the direction of displacement.

Appendix C: Computing the S-matrix

In our FEM analysis, we are solving a set of inhomoge-nous equations describing the behavior of our piezoelec-tric device. The drive term of these equations is a vec-

tor (V,a−)>

where V is the voltage across the leads ofour transducer and a− is a vector of coefficients for thepiezoelectric waves incident on the domain as defined byEquation 3.

The solutions of these equations can be represented inmatrix form (

Ia+

)=

(Y x>1x2 X

)(Va−

). (C1)

The scalar Y is the admittance of the transducer. For Mmodes, x1 and x2 are vectors with M components and Xis an M ×M matrix. For the simulations reported here,the coefficients of a− are set to 0 and we solve for thefirst column of the matrix in Equation C1 in terms of theinput voltage V .

In order to study how these transducers behave inphononic networks—for example, the two-port transmis-sion devices we use to measure tbµ—we want to transformthe matrix in Equation C1 into a scattering matrix S. Todo so, we reexpress V and I in terms of the microwaveamplitudes a±µ which we abbreviate to a± in this section

V =

√Z0

2(a+ + a−) (C2)

I =1√2Z0

(a− − a+) . (C3)

Here Z0 is impedance of the transmission line. Like am,the squares of the amplitudes a2

+ and a2− are the outward

and inward-going, time-averaged power in the transmis-sion line. This is easily checked by computing the powerinto the microwave port

V ∗I + V I∗ = |a−|2 − |a+|2 . (C4)

Substituting Equation C3 into Equation C1 and collect-ing terms we find the S matrix(

a+

a+

)=

(Y0−YY0+Y −

√2Y0

Y0+Y x>1√2Y0

Y0+Y x2 X + 1Y0+Y x2x

>1

)︸ ︷︷ ︸

S

(a−a−

)(C5)

where Y0 = Z−10 . From reciprocity S = S> we find

x ≡ x2 = −x1 and therefore

S =

rµµ t1µ t2µ . . .t1µ r11 t21

t2µ t21 r22

.... . .

=

(Y0−YY0+Y

√2Y0

Y0+Y x>√

2Y0

Y0+Y x X + 1Y0+Y xx>

).

(C6)The first component of S connecting a− and a+ is thereflection S-parameter S11 in the absence of reflectionsin the network (such as off a second transducer). Thecomponent of S connecting a− to the SH0 coefficient ab

is tbµ. Its magnitude is conveniently expressed

|tbµ| =√

1− |S211||xb|√

2G(C7)

where xb is the SH0 component of x and is the coefficientab for a 1 V drive. 2G |V |2 is the total power dissipatedby the transducer.

Appendix D: De-embedding tbµ, α, and rbb from S21

When making a transducer, especially one embeddedin a network, e.g., a transducer coupled to a resonator, itis tempting to be satisfied with a well-matched S11. Sup-pression of |S11|, i.e., low microwave reflections, seems toimply that microwaves are being converted to mechani-cal waves and that the device is efficient. Under thisprescription, one would choose an IDT’s width and sim-ply tune its length until it is matched. This approachdoes not lead to efficient devices.

A strong S11 dip is a necessary but insufficient con-dition for efficiency (|tbµ|2 → 1). In a microwave orphononic network, reflections can strongly modify theresponse of a component. Resonance can enhance thetransmission through the device. If network perfor-mance is the prime and only concern, measuring a res-onator’s intracavity phonon number against microwaveinput power, for example, will suffice. But if the goal isto make a transducer which can serve as a general compo-nent, one that can be embedded in an arbitrary networkand the response accurately predicted, we need to de-embed the transducer’s response from the larger networkresponse.

In Appendix C, we describe how the full scatteringmatrix S can be computed by the FEM. In Section IV,we show that transmission into the SH0 mode exceedsthe total transmission into all other modes by 10 dB.This allows us to reduce the S matrix of Equation C6 totwo-ports

S =

(rµµ tbµtbµ rbb

). (D1)

The S-matrix for the waveguide is

Swg =

(e−αL/2−iωτ 0

0 e−αL/2−iωτ

)(D2)

Page 9: arXiv:1907.13058v1 [physics.app-ph] 30 Jul 2019the promise of on-chip phononics. The modes of small waveguides have been selectively transduced with op-tical forces in cavity optomechanics

9

1.4 1.5 1.6 1.7 1.8 1.9 2.0 (GHz)

-120

-100

-80

-60

-40

-20

|S 21| (

dB)

a

b

c

FIG. 6. a. Signal-flow graph for a two-port device. b. Pathsfor the single, triple, and quintuple-transit (colored blue, red,and green respectively) corresponding to the S21 curves ofmatching color in c. c. S21 filtered by path.

where τ = L/vg is the transit time of the waveguide. Thedevices measured in Section V consist of a transducer,waveguide, and transducer. These components are cas-caded in the signal flow graph in Figure 6a which can bereduced by standard methods [44] to find

S11 = rµµ +t2bµrbbe

−αL−2iωτ

1− r2bbe−αL−2iωτ

(D3)

and

S21 =t2bµe

−αL/2−iωτ

1− r2bbe−αL−2iωτ

(D4)

where ports 1 and 2 are the electrical port of the first andsecond transducer. The second term in our expression forS11 comes from reflections rbb and gives rise to the Fabry-Perot peaks found on the blue side of ω0 in Figure 4c.

The impulse response h (t) is computed byinverse Fourier transforming S21. Expanding(1− r2

bbe−αL−2iωτ

)−1to

∑n r

2nbbe−nαL−2inωτ , each

term represents an echo in the impulse response inFigure 4e. These echoes and the paths they take arediagrammed in Figure 6b.

Since the L = 800 µm device is long enough to resolvethe echoes, the amplitudes of the echoes can be analyzeddirectly in the frequency domain by filtering out eachecho in Figure 4e associated with a path in Figure 6b andtaking the Fourier transform. The results of this proce-dure are inset to Figure 4e but are reproduced larger be-low for clarity. The transmission factor t2bµ exp (−αL/2)is extracted from the first transit plotted in blue.

Appendix E: Insertion loss from impedancemismatch

In section V we attribute a fraction of the insertionloss to impedance mismatch between the transducers andtransmission lines. This mismatch in Figure 4d is la-beled the two-port mismatch. Derived below, this quan-tity is the average of the ratios of the power dissipatedby each transducer over the incident microwave power1− |S11|2/2− |S22|2/2.

For any lossy, passive system

|S11|2 + |S12|2 < 1 (E1)

and

|S22|2 + |S21|2 < 1. (E2)

Summing these conditions and assuming reciprocity, i.e.S21 = S12 we have

|S11|2 + |S22|2 + 2|S21|2 < 2 (E3)

Rearranging the above expression we get

|S21|2 < 1− |S11|2/2− |S22|2/2 (E4)

The right-hand side which sets an upperbound on S21 isthe two-port mismatch.