Classical Stuckelberg interferometry of a nanomechanical ...

10
PHYSICAL REVIEW B 94, 245406 (2016) Classical St ¨ uckelberg interferometry of a nanomechanical two-mode system Maximilian J. Seitner, 1, 2, * Hugo Ribeiro, 3 Johannes K¨ olbl, 1 Thomas Faust, 2 org P. Kotthaus, 2 and Eva M. Weig 1, 2 1 Departement of Physics, University of Konstanz, 78457 Konstanz, Germany 2 Center for NanoScience (CeNS) and Fakult¨ at f ¨ ur Physik, Ludwig-Maximilians-Universit¨ at, Geschwister-Scholl-Platz 1, M¨ unchen 80539, Germany 3 Department of Physics, McGill University, Montreal, Quebec, Canada H3A 2T8 (Received 1 July 2016; revised manuscript received 20 October 2016; published 5 December 2016) St¨ uckelberg interferometry is a phenomenon that has been well established for quantum-mechanical two-level systems. Here, we present classical two-mode interference of a nanomechanical two-mode system, realizing a classical analog of St¨ uckelberg interferometry. Our experiment relies on the coherent energy exchange between two strongly coupled, high-quality factor nanomechanical resonator modes. Furthermore, we discuss an exact theoretical solution for the double-passage St ¨ uckelberg problem by expanding the established finite-time Landau- Zener single-passage solution. For the parameter regime explored in the experiment, we find that the St¨ uckelberg return probability in the classical version of the problem formally coincides with the quantum case, which reveals the analogy of the return probabilities in the quantum-mechanical and the classical version of the problem. This result qualifies classical two-mode systems at large to simulate quantum-mechanical interferometry. DOI: 10.1103/PhysRevB.94.245406 I. INTRODUCTION In 1932, St¨ uckelberg [1] investigated the dynamics of a quantum two-level system undergoing a double passage through an avoided crossing. For a given energy splitting, an interference pattern arises that depends on the transit time and the rate at which the energy of the system is changed. This discovery led to the advent of St¨ uckelberg interferometry, which allows for characterizing the parameters of a two-level system or for achieving quantum control over the system [2]. St¨ uckelberg interferometry has been studied intensively in a variety of quantum systems, e.g., Rydberg atoms [3], ultracold atoms and molecules [4], dopants [5], nanomagnets [6], quan- tum dots [710], and superconducting qubits [1115], as well as theoretically in a semiclassical optomechanical approach [16]. Here, we study experimentally a classical analog of St¨ uckelberg interferometry, i.e., the coherent energy exchange of two strongly coupled classical high-Q nanomechanical resonator modes. We employ the analytical solution [17] of the Landau-Zener problem describing the single passage through the avoided crossing [1,1820] to analyze the St¨ uckelberg problem, demonstrating that the classical coherent exchange of energy follows the same dynamics as the coherent tunneling of a quantum-mechanical two-level system. The past few years have seen the advent of highly versatile nanomechanical systems based on strongly coupled, high- quality factor modes [2123]. The strong coupling generates a pronounced avoided crossing of the classical mechanical modes realizing a nanomechanical two-mode system that can be employed as a testbed for the dynamics at energy level crossings [2124]. In the case of a quantum two-level system, e.g., spin-1/2, a single passage through the avoided crossing results in Landau- Zener dynamics originating from the tunneling of a quantum- mechanical excitation between two quantum states [18]. In the classical case, the exchange of excitation energy between * [email protected] two strongly coupled mechanical modes represents a well- established analogy to this process [25,26]. During a double passage through the avoided crossing within the coherence time of the system, phase is accumu- lated, leading to self-interference. This interference results in oscillations of the return probability, in a quantum-mechanical context well-known as St¨ uckelberg oscillations [1], which have previously been studied in many quantum systems [79,15,27]. In the classical case, the return probability is analogous to the probability that the excitation, namely oscillation energy, returns coherently to the same mechanical mode. II. NANOMECHANICAL TWO-MODE SYSTEM We explore experimentally a purely classical mechanical two-mode system, consisting of two orthogonally polarized fundamental flexural modes of a nanomechanical resonator [Fig. 1(a)]. The flexural modes belong to the in-plane and out-of-plane vibration of a 50-μm-long, 270-nm-wide, and 100-nm-thick doubly clamped, high-stress silicon nitride (SiN) string resonator. Dielectric drive and control via electric gradient fields [28] as well as the microwave cavity en- hanced heterodyne dielectric detection scheme [22,28,29] are provided via two adjacent gold electrodes, as detailed in Appendix A. Applying a dc voltage to the electrodes induces an electric polarization in the silicon nitride string, which, in turn, couples to the electric field gradient, resulting in a quadratic resonance frequency shift with the applied voltage [28]. The electric field gradients along the in- and out-of-plane direction have opposing signs, and hence they have an inverse tuning behavior. Whereas the out-of-plane oscillation shifts to higher mechanical resonant frequencies, the in-plane oscillation decreases in frequency with the applied dc voltage [28]. Hence, the inherent frequency offset of in- plane and out-of-plane oscillation, induced by the rectangular cross section of the string, can be compensated. Furthermore, the applied inhomogeneous electric field induces a strong coupling between the two modes [24]. Near resonance, they hybridize into normal modes [22], diagonally polarized along 2469-9950/2016/94(24)/245406(10) 245406-1 ©2016 American Physical Society

Transcript of Classical Stuckelberg interferometry of a nanomechanical ...

Page 1: Classical Stuckelberg interferometry of a nanomechanical ...

PHYSICAL REVIEW B 94, 245406 (2016)

Classical Stuckelberg interferometry of a nanomechanical two-mode system

Maximilian J. Seitner,1,2,* Hugo Ribeiro,3 Johannes Kolbl,1 Thomas Faust,2 Jorg P. Kotthaus,2 and Eva M. Weig1,2

1Departement of Physics, University of Konstanz, 78457 Konstanz, Germany2Center for NanoScience (CeNS) and Fakultat fur Physik, Ludwig-Maximilians-Universitat,

Geschwister-Scholl-Platz 1, Munchen 80539, Germany3Department of Physics, McGill University, Montreal, Quebec, Canada H3A 2T8

(Received 1 July 2016; revised manuscript received 20 October 2016; published 5 December 2016)

Stuckelberg interferometry is a phenomenon that has been well established for quantum-mechanical two-levelsystems. Here, we present classical two-mode interference of a nanomechanical two-mode system, realizing aclassical analog of Stuckelberg interferometry. Our experiment relies on the coherent energy exchange betweentwo strongly coupled, high-quality factor nanomechanical resonator modes. Furthermore, we discuss an exacttheoretical solution for the double-passage Stuckelberg problem by expanding the established finite-time Landau-Zener single-passage solution. For the parameter regime explored in the experiment, we find that the Stuckelbergreturn probability in the classical version of the problem formally coincides with the quantum case, which revealsthe analogy of the return probabilities in the quantum-mechanical and the classical version of the problem. Thisresult qualifies classical two-mode systems at large to simulate quantum-mechanical interferometry.

DOI: 10.1103/PhysRevB.94.245406

I. INTRODUCTION

In 1932, Stuckelberg [1] investigated the dynamics ofa quantum two-level system undergoing a double passagethrough an avoided crossing. For a given energy splitting, aninterference pattern arises that depends on the transit time andthe rate at which the energy of the system is changed. Thisdiscovery led to the advent of Stuckelberg interferometry,which allows for characterizing the parameters of a two-levelsystem or for achieving quantum control over the system [2].Stuckelberg interferometry has been studied intensively in avariety of quantum systems, e.g., Rydberg atoms [3], ultracoldatoms and molecules [4], dopants [5], nanomagnets [6], quan-tum dots [7–10], and superconducting qubits [11–15], as wellas theoretically in a semiclassical optomechanical approach[16]. Here, we study experimentally a classical analog ofStuckelberg interferometry, i.e., the coherent energy exchangeof two strongly coupled classical high-Q nanomechanicalresonator modes. We employ the analytical solution [17] of theLandau-Zener problem describing the single passage throughthe avoided crossing [1,18–20] to analyze the Stuckelbergproblem, demonstrating that the classical coherent exchangeof energy follows the same dynamics as the coherent tunnelingof a quantum-mechanical two-level system.

The past few years have seen the advent of highly versatilenanomechanical systems based on strongly coupled, high-quality factor modes [21–23]. The strong coupling generatesa pronounced avoided crossing of the classical mechanicalmodes realizing a nanomechanical two-mode system that canbe employed as a testbed for the dynamics at energy levelcrossings [21–24].

In the case of a quantum two-level system, e.g., spin-1/2, asingle passage through the avoided crossing results in Landau-Zener dynamics originating from the tunneling of a quantum-mechanical excitation between two quantum states [18]. Inthe classical case, the exchange of excitation energy between

*[email protected]

two strongly coupled mechanical modes represents a well-established analogy to this process [25,26].

During a double passage through the avoided crossingwithin the coherence time of the system, phase is accumu-lated, leading to self-interference. This interference results inoscillations of the return probability, in a quantum-mechanicalcontext well-known as Stuckelberg oscillations [1], which havepreviously been studied in many quantum systems [7–9,15,27].In the classical case, the return probability is analogous tothe probability that the excitation, namely oscillation energy,returns coherently to the same mechanical mode.

II. NANOMECHANICAL TWO-MODE SYSTEM

We explore experimentally a purely classical mechanicaltwo-mode system, consisting of two orthogonally polarizedfundamental flexural modes of a nanomechanical resonator[Fig. 1(a)]. The flexural modes belong to the in-plane andout-of-plane vibration of a 50-μm-long, 270-nm-wide, and100-nm-thick doubly clamped, high-stress silicon nitride (SiN)string resonator. Dielectric drive and control via electricgradient fields [28] as well as the microwave cavity en-hanced heterodyne dielectric detection scheme [22,28,29]are provided via two adjacent gold electrodes, as detailedin Appendix A. Applying a dc voltage to the electrodesinduces an electric polarization in the silicon nitride string,which, in turn, couples to the electric field gradient, resultingin a quadratic resonance frequency shift with the appliedvoltage [28]. The electric field gradients along the in- andout-of-plane direction have opposing signs, and hence theyhave an inverse tuning behavior. Whereas the out-of-planeoscillation shifts to higher mechanical resonant frequencies,the in-plane oscillation decreases in frequency with the applieddc voltage [28]. Hence, the inherent frequency offset of in-plane and out-of-plane oscillation, induced by the rectangularcross section of the string, can be compensated. Furthermore,the applied inhomogeneous electric field induces a strongcoupling between the two modes [24]. Near resonance, theyhybridize into normal modes [22], diagonally polarized along

2469-9950/2016/94(24)/245406(10) 245406-1 ©2016 American Physical Society

Page 2: Classical Stuckelberg interferometry of a nanomechanical ...

MAXIMILIAN J. SEITNER et al. PHYSICAL REVIEW B 94, 245406 (2016)

(b)

(c)

(a)

time t0

UiUf

Up

0 1 2 3 4 5

7.567.587.607.627.647.667.687.70

Freq

uenc

y (M

Hz)

Sweep voltage (V)

Ui

UaUfOut c (t)

In c (t)

1μm

Up=Ui+Up~

~

Up

FIG. 1. Nanomechanical resonator and measurement scheme.(a) False color scanning electron micrograph of the 50-μm-long,270-nm-wide, and 100-nm-thick silicon nitride string (green) flankedby two adjacent gold electrodes in the oblique view. Arrows indicatethe flexural mode polarizations out-of-plane (Out) and in-plane (In).The normalized amplitude of the respective mode is denoted c1(t)for out-of-plane polarization and c2(t) for in-plane polarization,as explained in the text. (b) Avoided mode crossing of sample Aexhibiting a frequency splitting of �/2π = 22.6 kHz at the avoidedcrossing voltage Ua = Ui + 1.47 V = 9.37 V. The lower (out-of-plane) mode is excited at frequency ω1(Ui)/2π = 7.560 MHz definedby the initialization voltage Ui. An additional sweep voltage appliesa triangular voltage ramp rising to a maximum of Up and back tothe readout voltage Uf = Ui + 0.2 V = 8.1 V, thus transgressing theavoided crossing twice. The sweep voltage also decouples the modefrom the fixed-frequency drive, consequently inducing an exponentialdecay of the amplitude. (c) Time evolution of the sweep voltagebeginning at t = 0, increasing to Up and returning to Uf after intervalϑ . The mechanical signal power (green dashed line) is measuredafter a delay ε, and a fit (black dotted line) is used to extract itsmagnitude at time t = ϑ . The measured return signal is normalizedto the mechanical signal power at t = 0.

±45◦. A pronounced avoided crossing with level splitting�/2π reflects the strong mutual coupling of the flexuralmechanical modes as depicted in Fig. 1(b).

To study Stuckelberg interferometry, we perform a doublepassage through the avoided crossing using a fast triangularvoltage ramp. We initialize the system at voltage Ui in the lowerbranch of the avoided crossing via a resonant sinusoidal drivetone at the resonance frequency ω1(Ui)/2π of the out-of-planeoscillation [cf. Fig. 1(b)]. As illustrated in Fig. 1(c), at timet = 0, a fast triangular voltage ramp with voltage sweep rate β

up to the peak voltage Up, and back to the readout voltage Uf ,is applied to tune the system through the avoided crossing.Note that the ramp detunes the system from the resonantdrive, and the mechanical energy starts to decay exponentially.At Uf , we measure the exponential decay of the mechanicaloscillation in the lower branch after time t = ϑ + ε, whereϑ is the duration of the ramp, i.e., the propagation time, andε serves as a temporal offset to avoid transient effects. Thesignal is extrapolated and evaluated at time ϑ by an exponentialfit and normalized to the signal intensity at the initializationpoint (t = 0), consequently yielding a normalized squaredreturn amplitude. The return signal has to be measuredat the readout voltage Uf at ω1(Uf)/2π since the fixed rf

drive tone at ω1(Ui)/2π cannot be turned off during themeasurement. The presented voltage sequence is analogousto the one employed in Ref. [27], and it differs from thefrequently performed periodic driving scheme in Stuckelberginterferometry experiments [2].

III. FINITE-TIME STUCKELBERG THEORY

We follow the work of Novotny [30] to derive the classicalflow (Hamiltonian flow [31]) describing the dynamics of thesystem in the vicinity of the avoided crossing. We start withNewton’s equation of motion for the displacement,

mu1(t) = −k1u1(t) − κ[u1(t) − u2(t)],

mu2(t) = −k2u2(t) + κ[u1(t) − u2(t)],(1)

with uj (t) (j = 1,2) describing, respectively, the out-of-plane(j = 1) and in-plane (j = 2) displacement of the center ofmass of the oscillator, kj is the spring constant of mode j ,κ is the coupling constant between the two modes, and m

is the effective mass of the oscillator. We look for solutionsof the form uj (t) = cj (t) exp(iω1t), with cj (t) a normalizedamplitude, i.e., |c1(t)|2 + |c2(t)|2 = 1, and we have definedωj = √

(kj + κ)/m as the dressed resonance frequency ofmode j in units of 2π . In the experimentally relevant limitwhere κ/k1 � 1, the amplitudes cj (t) are slowly varying intime as compared to the oscillatory function exp(iω1t). Asa consequence, it is possible to neglect the second derivatescj (t) in the equations describing the motion of cj (t), whichare obtained by replacing the ansatz for uj (t) in Eq. (1). Thus,the system of coupled differential equations describing theevolution of the normalized amplitudes is

ic1 = κ

2ω1mc2,

(2)

ic2 = κ

2ω1mc1 − ω2

2 − ω21

2ω1c2.

In the vicinity of the avoided crossing, where the modes can ex-change energy, we have ω2 � ω1 such that (ω2

2 − ω21)/2ω1 �

ω2 − ω1. If we further assume ω2 − ω1 � αt , with α thefrequency sweep rate, and we define � = |λ| = κ/(mω1),Eq. (2) reduces to

ic(t) = H (t)c(t), (3)

with c(t) = [c1(t) c2(t)]T and

H (t) =(

0 λ2

λ2 −αt

). (4)

Since we are interested in multiple passages through theavoided crossing, we look for the classical flow ϕ(t,ti) definingthe state of the system at time t given that we know itsstate at some prior time ti, c(t) = ϕ(t,ti)c(ti). Typically, c(ti)is the initial condition of the system. One can show thatthe classical flow obeys the same differential equation asc(t), iϕ(t,ti) = H (t)ϕ(t,ti). By applying the time-dependentunitary transformation S(t) = exp(iαt2/4)12 to the classicalflow, i.e., ϕ(t,ti) = S(t)ϕ(t,ti)S†(ti), we find that ϕ(t,ti) obeysthe differential equation,

i ˙ϕ(t,ti) = (S†(t)H (t)S(t) − iS†(t)S(t))ϕ(t,ti)

= H (t)ϕ(t,ti), (5)

245406-2

Page 3: Classical Stuckelberg interferometry of a nanomechanical ...

CLASSICAL STUCKELBERG INTERFEROMETRY OF A . . . PHYSICAL REVIEW B 94, 245406 (2016)

0.20.40.60.81.01.2

0.2

0.4

0.6

0.8

1.0

0 10 20 300.00.20.40.60.81.01.2

0 10 20 30 400.0

0.2

0.4

0.6

0.8

1.0

Inverse sweep rate 1/β (μs/V)

Nor

mal

ized

squ

ared

retu

rn a

mpl

itude

Theo

retic

al re

turn

pro

babi

lity

(a)

(d)(c)

(b)

FIG. 2. Classical Stuckelberg oscillations. Normalized squared return amplitude (left axis, blue dots) and theoretically calculated returnprobability (right axis, red line) vs inverse sweep rate for fixed peak voltages of Up = 2.5 V (a), Up = 3.5 V (b), Up = 4.5 V (c), andUp = 5.0 V (d) measured on sample A.

with

H (t) =(

αt2

λ2

λ2 −αt

2

), (6)

where 12 denotes the unity operator in two dimensions.Equation (5) coincides with the Schrodinger equation for theunitary evolution operator of the Landau-Zener problem, forwhich an exact finite-time solution is known [17] (see alsoAppendix B). With the help of the classical flow, one caneasily calculate the state of the system after a double passagethrough the avoided crossing (Stuckelberg problem). We find

c(t) = ϕb(t, − tp)ϕ(tp,ti)c(ti), (7)

with ϕb(t,ti) = σxϕ(t,ti)σx describing the evolution of thesystem during the back sweep (see Appendix B) where σx

denotes the Pauli matrix in the x direction, and tp labels thetime at which the forward (backward) sweep stops (starts).From Eq. (7), one can obtain the return probability to mode 1,

P1→1 = |ϕ11(tp,ti)ϕ∗11(t, − tp) + ϕ∗

12(tp,ti)ϕ∗12(t, − tp)|2, (8)

with ϕij (t,ti) the matrix elements of ϕ(t,ti). Note that we usethe frequency sweep rate α in the theory, which is convertedto the experimentally accessible voltage sweep rate β via aconversion factor ζ = 55 kHz/V as elucidated in Appendix C.

The analogy between the unitary evolution operator and theclassical flow, both expressed in the basis of uncoupled states(modes), allows one to draw the analogy to the quantum-mechanical return probability in Stuckelberg interferometry.The normalized amplitudes are associated with the normalizedenergy in each resonator mode, and they differ conceptuallyfrom the probability that a quantum-mechanical two-levelsystem is found in either of the two quantum states. Neverthe-less, the dynamics of the normalized amplitudes in classicalStuckelberg interferometry is analogous to the dynamics of thequantum-mechanical probabilities since they follow the sameequations. In this sense, the coherent exchange of oscillationenergy between two coupled modes can be associated with the

transfer of population between two quantum states. A moredetailed discussion and comparison of our theoretical approachto previous models [2,17,25,26] reveals previously unchartedparameter regimes in Stuckelberg interferometry, and it willbe presented elsewhere [32].

IV. CLASSICAL STUCKELBERG INTERFEROMETRY

Experimentally, we investigate classical Stuckelberg os-cillations with two different samples in a vacuum of� 10−4 mbar. Sample A is investigated at 10 K in atemperature-stabilized pulse tube cryostat, which offers agreatly enhanced stability of the electromechanical systemagainst temperature fluctuations. Sample B is explored at roomtemperature in order to confirm the results and to check theirstability under ambient temperature fluctuations. Note that inboth experiments, the system operates deeply in the classicalregime [22], and it does not exhibit any quantum-mechanicalproperties. Sample A exhibits a mechanical quality factor Q =ω/� ≈ 2 × 105 and linewidth �/2π ≈ 40 Hz at resonancefrequency ω1(Ui)/2π = 7.560 MHz of the 50-μm-long stringresonator ensuring classical coherence times in the millisecondregime [22]. The level splitting �/2π = 22.6 kHz exceedsthe mechanical linewidth by almost three orders of magnitude,which puts the system deep into the strong-coupling regime.

We initialize the system at Ui = 7.9 V and apply triangularvoltage ramps with different voltage sweep rates β for a setof peak voltages Up. Figure 2 depicts the normalized squaredreturn amplitude for different peak voltages and the theoreticalreturn probabilities calculated without any free parameters.The normalized squared return amplitude may exceed a valueof unity due to normalization artefacts that arise from thedifferent signal magnitudes at the initialization and readoutvoltages in addition to measurement errors. We observe clearoscillations in the return signal, in good agreement withthe theoretical predictions for lower peak voltages. As thenumber of oscillations increases for higher peak voltages, the

245406-3

Page 4: Classical Stuckelberg interferometry of a nanomechanical ...

MAXIMILIAN J. SEITNER et al. PHYSICAL REVIEW B 94, 245406 (2016)

Inverse sweep rate 1/β (μs/V) Nor

mal

ized

squ

ared

retu

rn a

mpl

itude

20 40 60 80 100

Pea

k vo

ltage

Up (

V)

3.0

3.5

4.0

4.5

5.0

0

0.5

1.0

1.5

Inverse sweep rate 1/β (μs/V)

Theo

retic

al re

turn

pro

babi

lity

20 40 60 80 100

Pea

k vo

ltage

Up (

V)

3.0

3.5

4.0

4.5

5.0

0

0.25

0.5

0.75

(a) (b)

1.0

FIG. 3. Comparison of experimental data and the theoretical model. (a) Color-coded normalized squared return amplitude vs inverse sweeprate and peak voltage measured on sample B. The dataset is not interpolated. (b) Color-coded theoretical return probability given by Eq. (9) vsinverse sweep rate and peak voltage for the equivalent data range. The theory is calculated with a single set of parameters, extracted from theavoided crossing illustrated in Appendix C (Fig. 7), and it contains no free parameters.

deviation from the theoretical prediction is more pronounced.We attribute this to uncertainties and fluctuations of thecharacteristic sweep parameters of the system, which changeunder application of the voltage ramp and over time asdiscussed in Appendixes D and E. A further deviation arisesfrom the assumption that a linear change of the voltage leadsto a linear change of the difference in frequency. This is onlyan approximation since the mechanical resonance frequenciestune quadratically with the applied voltage [28]. However,since most of the energy exchange happens in the vicinityof the avoided crossing, where the difference in frequencyis linearized, one expects to see noticeable deviations fromtheory only for higher peak voltages.

To reproduce the experimental data and to test the stabilityof classical Stuckelberg interferometry against fluctuations,we repeat the experiment on a second sample of the samedesign at room temperature (sample B, denoted by index “B”).The now 55-μm-long resonator has a mechanical linewidth of�B/2π ≈ 25 Hz at frequency ωB,1(UB,i)/2π = 6.561 MHz,which results in a quality factor of QB ≈ 2.6 × 105 at theinitialization voltage UB,i = 10.4 V and hence an improvedmechanical lifetime of 6.21 ms. Furthermore, the sample ex-hibits a mode splitting of �B/2π = 6.3 kHz and a conversionfactor of ζB = 19 kHz/V.

Figure 3 depicts a color-coded two-dimensional map ofthe normalized squared return amplitude as a function of theinverse voltage sweep rate β and the peak voltage Up alongsidethe theoretical return probability of the classical Stuckelbergoscillations, again calculated with no free parameters. We in-vestigate double passages up to a total propagation time of ϑ =1.0 ms in the experiments conducted on sample B. To accountfor the decay of both modes when tuned away from the drive forthe considerably longer ramps applied to sample B, we modelthe mechanical damping by an exponential decay with anaveraged decay time t0 = 5.7 ms. After a measurement time tm,the probability to measure an excitation of mode j is given by

|cj (tm)|2 = exp(−tm/t0)P1→j , (9)

with P1→1 given by Eq. (8) and P1→2 = 1 − P1→1. Theexperimental data show remarkably good agreement with thetheoretical predictions, despite temperature fluctuations of

several degrees kelvin per hour, which shift the mechanicalresonance frequency up to 40 linewidths. To initialize thesystem at the same resonance frequency in each measure-ment, a feedback loop regulates the initialization voltage Ui

(see Appendix D). Consequently, the recording of a singlehorizontal scan at a fixed peak voltage in Fig. 3(a) takesup to 16 hours, incorporating a non-negligible amount offluctuations of the system parameters, such as, e.g., thecenter voltage of the avoided crossing Ua, which imposesconsiderable uncertainties on the parameters used for thetheoretical calculations. To further illustrate the influence offluctuations, Figs. 4(a) and 4(b) depict horizontal and vertical

0 20 40 60 80 1000.0

0.5

1.0

0.0

0.2

0.4

0.6

0.8

1.0

Inverse sweep rate 1/β (μs/V)

Nor

mal

ized

squ

ared

retu

rn a

mpl

itude

Theo

retic

al re

turn

pro

babi

lity

3.0 3.5 4.0 4.5 5.00.0

0.5

1.0

0.0

0.2

0.4

0.6

0.8

1.0

Peak voltage Up (V)

(a)

(b)

FIG. 4. Exemplary classical Stuckelberg oscillations of sampleB. Line cuts of Figs. 3(a) and 3(b). (a) Normalized squared returnamplitude (left axis, blue dots) and theoretically calculated returnprobability given by Eq. (9) (right axis, red line) vs inverse sweeprate for a fixed peak voltage Up = 3.3 V. (b) The same quantities asabove but plotted as a function of peak voltage for a fixed inversesweep rate 1/β = 51.6 μs/V. Blue dots are joined by blue dashedlines for illustration reasons.

245406-4

Page 5: Classical Stuckelberg interferometry of a nanomechanical ...

CLASSICAL STUCKELBERG INTERFEROMETRY OF A . . . PHYSICAL REVIEW B 94, 245406 (2016)

line cuts of the two-dimensional map in Fig. 3 at Up = 3.3 Vand at inverse sweep rate 1/β = 51.6 μs/V, respectively.For small inverse sweep rates, i.e., very fast sweeps, theexperimental data in Fig. 4(a) deviate from the theoreticalmodel due to a flattening of the voltage ramps in the room-temperature experiment (see Appendix E). For sweeps with1/β � 50 μs/V, the experimentally observed Stuckelbergoscillations exhibit good agreement with the theoretical predic-tions even for the line cut along the vertical peak voltage axis[cf. Fig. 4(b)]. Note that Fig. 4 depicts the best results from alldatasets at room temperature. Further exemplary line cuts areprovided in Appendix E, also exhibiting a clear oscillatorybehavior in the normalized squared return amplitude, butincorporating larger deviations from theory in certain regionsand therefore revealing fluctuations of system parameters overtime, predominately induced by temperature drifts.

V. CONCLUSION AND OUTLOOK

In conclusion, we have demonstrated classical Stuckelbergoscillations that have previously been experimentally ob-served exclusively in the framework of quantum mechan-ics [2,7,8,27]. An exact solution of the Stuckelberg prob-lem [1] based on the finite-time Landau-Zener single-passagesolution [17] describes the return probability in the classicalversion of the double-passage problem, which is shown tofollow the same equation as in the case of a quantum two-level system. In this way, we have demonstrated that in ourexperimental parameter limit, the coherent exchange of en-ergy between two strongly coupled classical nanomechanicalresonator modes follows the same dynamics as the exchangeof excitations in a quantum-mechanical two-level system inthe framework of Stuckelberg interferometry.

This analogy might be exploited in future experimentsin order to determine whether a system of coupled me-chanical modes operates in the classical regime or in thequantum-mechanical limit. In the general case of two coupledquantum harmonic oscillators, the effective model describingthe dynamics would resemble that of the multiple-crossingLandau-Zener problem [33], which leads to a much morecomplex dynamics than the provided solution for the classicalcase. Hence, if one would be able to design an experimentin which two strongly coupled mechanical oscillators can becooled to their quantum-mechanical limit, it should be possibleto determine if the system operates in the classical or thequantum regime by means of a simple Stuckelberg returnamplitude measurement rather than by probing the Wignerdistribution function [34]. Whereas the presented experimentalsetup is still far from operation in the quantum-mechanicallimit, we could imagine this technique to be applied to differentmechanical systems that already demonstrated the quantumlimit of mechanical oscillators [35,36].

Overall, we have found good agreement between ex-periment and theory. However, parameter regimes yieldinglarger deviations are reminiscent of the sensitivity of theexact Stuckelberg solution to the initial system parameters,such as the position of the avoided crossing, and hence tofluctuations in the system. This circumstance, in turn, mightbe exploited for future investigations in resonator metrologyof decoherence and noise, adapting the approach to employ

Stuckelberg interferometry to characterize the coherence of aqubit [10].

Furthermore, the possibility to create a superposition stateof two mechanical modes may allow for future applicationas highly sensitive nanomechanical interferometers [37–39]analogous to the applications with cold atom and moleculematter-wave interferometers [4,40–42], whereas the presenceand implications of, e.g., phase noise [43] can be resolved bya change in resonator population and interference pattern.

Finally, classical Stuckelberg interferometry should not belimited to the presented strongly coupled, high-quality factornanomechanical string resonator modes [22], but it can beobserved in principle in every classical two-mode systemexhibiting the possibility of a double passage through anavoided crossing within the classical coherence time.

ACKNOWLEDGMENTS

Financial support by the Deutsche Forschungsgemeinschaftvia the collaborative research center SFB 767 and via projectKo 416-18 is gratefully acknowledged. H. R. acknowledgesfunding from the Swiss SNF. We thank Aashish A. Clerk fora critical reading of the manuscript.

APPENDIX A: THE NANOELECTROMECHANICALSYSTEM

The nanomechanical device and experimental setup aredepicted in Fig. 5. The sample investigated at a temperatureof 10 K (sample A) consists of a 50-μm-long, 270-nm-wide, and 100-nm-thick doubly clamped silicon nitride (SiN)string resonator. The room-temperature measurements were

(a)

OutIn

1μm

Microwavecavity

(b)

10K

+

~

Radiofrequency drive

AFG

±dc tuning

300Kμw read-out

Summationamplifier

FIG. 5. Nanoelectromechanical system. (a) False color scanningelectron micrograph of the 50-μm-long, 270-nm-wide, and 100-nm-thick silicon nitride string (green) in the oblique view. The adjacent1-μm-wide gold electrodes (yellow) are processed on top of thesilicon nitride layer. Arrows indicate the flexural mode polarizationsout-of-plane (Out) and in-plane (In). (b) Electrical transductionsetup. The arbitrary function generator (AFG) ramp voltage and thedc tuning voltage are added via a summation amplifier and thencombined with the rf drive using a bias tee. The microwave readoutis bypassed by the second capacitor, acting as a ground path for themicrowave cavity.

245406-5

Page 6: Classical Stuckelberg interferometry of a nanomechanical ...

MAXIMILIAN J. SEITNER et al. PHYSICAL REVIEW B 94, 245406 (2016)

dc Voltage (V)

Freq

uenc

y (M

Hz)

6.55

6.60

6.65

6.70

6.75

Sig

nal (

dBm

)

-130-125-120-115-110-105-100-95

-10-20 100 20

In

Out

FIG. 6. Dielectric frequency tuning. Color-coded frequency spec-trum of sample B as a function of applied dc tuning voltage. Theresonance frequency of the 55-μm-long resonator’s fundamentalout-of-plane oscillation (Out) increases quadratically as a function ofdc voltage. The resonance frequency of the corresponding in-planemode (In) decreases quadratically. Tuning both modes into resonance,they exhibit a pronounced avoided crossing indicated by the blackdashed rectangle. This particular region is displayed in Fig. 7(a).The additional resonances in the spectrum originate from a differentmechanical resonator, which is coupled to the same microwave cavity.

conducted on a similar sample (sample B), differing only inits resonator length of 55 μm. As stated in the main text, thetemperature does not affect the purely classical character of thesystem. The string resonators exhibit a high intrinsic tensilepre-stress of σSiN = 1.46 GPa resulting from the LPCVDdeposition of the SiN film on the fused silica substrate. Thishigh stress translates into large intrinsic mechanical qualityfactors of up to Q ≈ 500 000, which reduce quadraticallywith the applied dc tuning voltage in the experiment as aresult of dielectric damping [28]. Dielectric drive, detection,and control are provided via two adjacent gold electrodesin an all integrated microwave cavity enhanced transductionscheme [22,24,28,29]. In the experiment, we consider thetwo orthogonally polarized fundamental flexural modes ofthe nanomechanical string resonator, namely the oscillationperpendicular to the sample plane (out-of-plane) and theoscillation parallel to the sample plane (in-plane). Applying adc voltage to one of the two gold electrodes induces an electricpolarization in the silicon nitride string resonator, whichcouples to the field gradient of the inhomogeneous electricfield. Consequently, the mechanical resonance frequenciestune quadratically with the applied dc voltage as depictedin Fig. 6. Whereas the out-of-plane resonance (Out) tunestoward higher resonance frequencies as a function of dcvoltage, the resonance frequency of the in-plane mode (In)decreases [28]. Dielectric tuning of both modes into resonancereveals a pronounced avoided crossing originating from thestrong mutual coupling induced by the inhomogeneous electricfield. In the coupling region, the mechanical modes hybridizeinto diagonally (±45◦) polarized eigenmodes of the stronglycoupled system.

APPENDIX B: THEORETICAL MODEL

In this Appendix, we derive an exact expression forthe classical return probability. A detailed discussion and

comparison of our theoretical approach to previous modelswill be published elsewhere [32].

We start by solving the system of first-order differentialequations defined in Eq. (5) of the main text. Since theseequations are formally identical to the Schrodinger equationfor the (quantum) Landau-Zener problem, we can followthe work of Vitanov et al. [17] to derive the classical flow,c(τ ) = ϕ(τ,τi)c(τi) with c(τ ) = [c1(τ ) c2(τ )]T. Here, τ = √

αt

is a dimensionless time and τi is the initial dimensionlesstime. Note that we use dimensionless times in this appendixin order to provide a derivation that is consistent with thework of Vitanov et al. [17]. The equations in dependenceof times in the main text can be recovered by replacementof the dimensionless times following the above definition.In Appendix C, we provide the explicit conversion fromexperimentally accessible parameters to the dimensionlesstimes. We find(

c1(τ )

c2(τ )

)=

(ϕ11(τ,τi) ϕ12(τ,τi)

−ϕ∗12(τ,τi) ϕ∗

11(τ,τi)

)(c1(τi)

c2(τi)

)(B1)

with

ϕ11(τ,τi) = �(1 + i

η2

4

)√

[D−1−i

η2

4(e−i 3π

4 τi)D−iη2

4(ei π

4 τ )

+D−1−iη2

4(ei π

4 τi)D−iη2

4(e−i 3π

4 τ )]

(B2)

and

ϕ12(τ,τi) = �(1 + i

η2

4

)√

2

ηe−i π

4[D−i

η2

4(e−i 3π

4 τi)D−iη2

4(ei π

4 τ )

−D−iη2

4(ei π

4 τi)D−iη2

4(e−i 3π

4 τ )]. (B3)

Here, η = λ/√

α is the dimensionless coupling, �(z) is theGamma function, and Dν(z) is the parabolic cylinder function.To find the flow describing the evolution of the amplitudesdefined in Eq. (2), we apply the unitary transformation definedin the main text, ϕ(τ,τi) = S(τ )ϕ(τ,τi)S†(τi), with

S(τ ) = exp

[i

4τ 2

]12. (B4)

We find

ϕ(τ,τi) = exp

[i

4

(τ 2 − τ 2

i

)]ϕ(τ,τi). (B5)

The flow ϕ(τ,τi) describes the evolution of the normalizedamplitudes for a forward sweep; the frequency of mode 1(2) increases (decreases) with time. This implies that theback sweep cannot be described by ϕ(τ,τi) since during theevolution the frequency of mode 1 (2) decreases (increases).Hence, the system of coupled differential equations describingthe dynamics during the backward sweep (denoted by index“b”) is given by

i

(˙c1,b

˙c2,b

)=

(−αt2

λ2

λ2

αt2

)(c1,b

c2,b

). (B6)

The solutions of Eq. (B6) can be obtained analogously to theforward flow since the matrices appearing in Eq. (6) of the

245406-6

Page 7: Classical Stuckelberg interferometry of a nanomechanical ...

CLASSICAL STUCKELBERG INTERFEROMETRY OF A . . . PHYSICAL REVIEW B 94, 245406 (2016)

Freq

uenc

y di

ffere

nce

(kH

z)

Sweep voltage (V)

(a) (b)

0 1 2 3 4 5

6.56

6.58

6.60

Sweep voltage (V)

Freq

uenc

y (M

Hz)

ΔB/2�

0 1 2 3 4 5-60

-40

-20

0

20

40

FIG. 7. Calibration of the conversion factor. (a) Avoided crossing region of sample B. A sweep voltage equal to zero corresponds tothe initialization point Ui = 10.4 V. The two modes exhibit a frequency splitting of �B/2π = 6.3 kHz at the avoided crossing voltageUa = Ui + 1.96 V = 12.36 V. The gap in the upper branch (red) results from a signal detection efficiency of the particular mode polarizationbelow the noise level. (b) Frequency difference of the two modes (black) and averaged slope of the linearized frequency tuning illustrated by agreen dash-dotted line.

main text and Eq. (B6) are related by a unitary transformation.We find

ϕb(τ,τi) = σxϕ(τ,τi)σx =(

ϕ∗11(τ,τi) −ϕ∗

12(τ,τi)

ϕ12(τ,τi) ϕ11(τ,τi)

), (B7)

where σx denotes the Pauli matrix in the x direction. Theflow describing the evolution of the amplitudes c1(t) and c2(t)during the back sweep is obtained as previously, and we have

ϕb(τ,τi) = exp

[i

4

(τ 2 − τ 2

i

)]ϕb(τ,τi). (B8)

The state of the system after a double sweep is given by

c(τ ) = ϕb(τ, − τp)ϕ(τp,τi)c(τi), (B9)

where τp labels the time at which the first sweep stops and −τp

corresponds to the initial time of the back sweep [cf. Eq. (7)of the main text]. As stated in Eq. (8) of the main text, theprobability to return to mode 1 is then given by

P1→1(τ,τp,τi)

= |ϕ11(τp,τi)ϕ∗11(τ,−τp) + ϕ∗

12(τp,τi)ϕ∗12(τ,−τp)|2. (B10)

APPENDIX C: CONVERSION FACTOR CALIBRATION

In the theoretical model, the state of the system aftera double passage through the avoided crossing depends onthe characteristic sweep times. Experimentally, we realizethis double passage by the application of fast triangularvoltage ramps, tuning the resonant frequency of the mechanicalmodes [28]. In the following, we focus on sample B to illustratehow the different times are obtained.

We initialize the resonance in the lower-frequency branchat the voltage Ui = 10.4 V, where we apply a continuous sinu-soidal drive tone at ω1(Ui)/2π = 6.561 MHz. We then rampthe sweep voltage up to the peak voltage Up across the avoidedcrossing at voltage Ua = Ui + 1.96 V = 12.36 V and thenback to the readout voltage Uf = Ui + 0.5 V = 10.9 V, wherethe oscillation energy is read out again in the lower-frequencybranch. The offset of the readout voltage with respect to theinitialization voltage is necessary since we cannot stop thesinusoidal drive tone at ω1(Ui)/2π during the experiment. For

a fixed peak voltage Up, the voltage sweep is performed fordifferent voltage sweep rates β, given in the experimentalunits [β] = V/s. In the theoretical model, the frequencydifference of the two modes in units of 2π is approximatedby ω2 − ω1 � αt , where the sweep rate α has the dimensions[α] = 2π × Hz/s. Consequently, we introduce the conversionfactor ζ from voltage to frequency, defined via the relation

α = 2π × ζβ. (C1)

Figure 7 illustrates the calibration of the conversion factor. Asis conventional in experiments on Stuckelberg interferometry,the frequency difference of the two mechanical modes isapproximated to be linear in time, i.e., linear in sweep voltage.In our particular system, the resonance frequencies of themechanical flexural modes tune quadratically with voltageoutside of the avoided crossing (see Fig. 6). Nevertheless,for the designated region around the avoided crossing, thetwo frequency branches can be linearized as follows. We takethe frequency difference of both modes before and after theavoided crossing [cf. Fig. 7(b)], respectively, and we extractthe slopes via a linear fit. The two different slopes on the leftand the right-hand side of the avoided crossing are averaged,yielding an effective conversion factor (dash-dotted green line)

ζ = 19kHz

V. (C2)

To estimate the error of the conversion factor, we apply aquadratic fit to the frequency difference of the two modes,and we find a fit residual of approximately 2 kHz/V betweenthe linear and the quadratic fit.

Depending on the specific peak voltage Up, one could takeinto account a weighted average of the two slopes in orderto mitigate the deviation of the quadratic frequency tuningfrom the linear approximation. Here, one has to point outdeliberately that we neglect any weighted average, but wetake solely the above conversion factor for the calculation ofthe theoretical return probabilities. We are well aware of thefact that this linearization translates into a direct discrepancybetween the theoretical model and the experimental results.Nevertheless, in our opinion, these discrepancies are prevailedby the benefits of a closed theoretical calculation using a singleset of parameters that is supported by the remarkably good

245406-7

Page 8: Classical Stuckelberg interferometry of a nanomechanical ...

MAXIMILIAN J. SEITNER et al. PHYSICAL REVIEW B 94, 245406 (2016)

agreement between experiment and theory. Hence, we expressthe characteristic sweep times in the theoretical model by thefollowing parameters extracted from the avoided crossing inFig. 7(a):

ti = − 1

β(Ua − Ui) = τi√

α,

tp = 1

β(Up − Ua) = τp√

α,

tf = 1

β(Ua − Uf) = τf√

α,

(C3)

where Up = Ui + Up. As explained above, the return probabil-ity is measured at the readout voltage Uf = Ui. Consequently,we replace τ by τf in the back sweep of the theory, whichmodifies Eq. (B10) to

P1→1(τf,τp,τi)

= |ϕ11(τp,τi)ϕ∗11(τf,−τp) + ϕ∗

12(τp,τi)ϕ∗12(τf,−τp)|2. (C4)

APPENDIX D: TEMPERATURE FLUCTUATIONS

As stated in the main text, the measurement of thenormalized squared return amplitude for various voltage sweeprates β at a particular peak voltage Up takes up to 16 h. Duringthis time, the ambient temperature undergoes fluctuations of±2 K per hour due to insufficient air conditioning. Sincethe mechanical resonance frequency shifts due to thermalexpansion of the silicon nitride by approximately 500 Hz/K,both resonances shift by approximately 40 linewidths. Toinitialize the system at the same resonance frequency for everyparticular measurement, we implement a feedback loop thatregulates the initialization voltage. Therefore, the initializationvoltage shifts slightly from measurement to measurement,reflecting the temperature fluctuations. Figure 8 depicts the

0 20 40 60 80 100-50

0

50

100

Inverse sweep rate 1/β (μs/V)

Initi

aliz

atio

n vo

ltage

shi

ft (m

V)

FIG. 8. Temperature fluctuations. Initialization voltage shift vsinverse sweep rate for the dataset depicted in Fig. 4(a) (Up = 3.3 V)of the main text. Each point corresponds to the measurement of thenormalized squared return amplitude for a given inverse sweep rate.The first measurement is performed at 1/β = 100 μs/V, representingthe initialization at resonance frequency ωi(Ui)/2π for voltageUi = 10.4 V. The implemented feedback loop regulates the initializa-tion voltage in order to compensate for the temperature fluctuations ofthe mechanical resonance. Consequently, the voltage shift illustratesthe fluctuations of the ambient temperature.

initialization voltage shift versus inverse sweep rate for thedataset of peak voltage Up = 3.3 V, which corresponds tothe measurement depicted in Fig. 4(a) of the main text. Eachpoint represents a single measurement for a particular sweeprate. The first measurement is performed at an inverse sweeprate of 100 μs/V at the initialization voltage Ui = 10.4 Vand therefore corresponds to a shift of zero volts. Clearly,the temperature fluctuations not only affect the initializationvoltage required to obtain the desired resonance frequency,but they will also alter other system parameters, such asthe position of the avoided crossing Ua, that greatly affectthe theory (cf. Appendix C). Consequently, the temperaturefluctuations lead to deviations between experiment and theory,since we calculate the return probability with a single set ofparameters. In turn, these deviations might be used to inferfluctuations of the system in future applications of Stuckelberginterferometry.

APPENDIX E: EXPERIMENTAL UNCERTAINTIES

In Fig. 9 we provide additional horizontal and vertical linecuts from Fig. 3 of the main text. We observe pronouncedoscillations in the normalized squared return amplitude (bluedots) as well as in the theoretically calculated return probability(red line). Nevertheless, the deviations between experimentand theory are more apparent, especially for Fig. 9(b), whichdepicts a vertical line cut for a fixed inverse sweep rate of1/β = 60 μs/V, i.e., within the “plateau” in Fig. 3(b) of themain text. Whereas the normalized squared return amplitudeexhibits destructive interference, with the signal dropping

0 20 40 60 80 1000.0

0.5

1.0

0.0

0.2

0.4

0.6

0.8

1.0

3.0 3.5 4.0 4.5 5.00.0

0.5

1.0

0.0

0.2

0.4

0.6

0.8

1.0

Inverse sweep rate 1/β (μs/V)

Nor

mal

ized

squ

ared

retu

rn a

mpl

itude

Theo

retic

al re

turn

pro

babi

lity

Peak voltage Up (V)

(a)

(b)

FIG. 9. Classical Stuckelberg oscillations. (a) Normalizedsquared return amplitude (left axis, blue dots) and theoreticallycalculated return probability given by Eq. (9) of the main text(right axis, red line) vs inverse sweep rate for a fixed peak voltageUp = 3.85 V. (b) Same quantities as above but plotted as a functionof peak voltage for a fixed inverse sweep rate 1/β = 60 μs/V. Bluedots are joined by blue dashed lines for illustration reasons.

245406-8

Page 9: Classical Stuckelberg interferometry of a nanomechanical ...

CLASSICAL STUCKELBERG INTERFEROMETRY OF A . . . PHYSICAL REVIEW B 94, 245406 (2016)

close to zero, the minima in the return probability saturateat a value of approximately 0.3. This discrepancy is supposedto originate from the high sensitivity of the theoretical model tothe input parameters. Experimental uncertainties and fluctua-tions deter the system from interference with the same constantparameters throughout all individual measurements. Since the“plateau” in the theory is characteristic for a particular set ofexact and constant parameters, it cannot be recovered underthe given experimental conditions.

The experimental uncertainties arise not solely from thetemperature fluctuations. The voltage ramp also affects thecharacteristic parameters, such as the exact position ofthe avoided crossing Ua. As previously stated, the dc volt-age induces dipoles in the silicon nitride string resonator,which couple to the electric field gradient. A variation indc voltage changes the inhomogeneous electric field at thesame time, to which the nanoelectromechanical system needsto equilibrate. Consequently, the resonance frequencies ofthe mechanical modes drift toward the equilibrium positionof the system. This drift, in turn, alters the characteristicsystem parameters, i.e., the characteristic voltages used forthe theoretical calculations, and it depends on the magnitudeof the peak voltage Up. Concerning the initialization voltage,we simultaneously account for this effect via the initializationfeedback loop (see Sec. IV). Nevertheless, the exact position ofthe avoided crossing Ua varies slightly due to this retardationeffect. Experimentally, we mitigate the influence of this driftby means of a “thermalization” break of 10 s after each voltageramp.

Another possible uncertainty arises from the imprecision inthe value of the peak voltage Up at the sample. The outputamplitude uncertainty of the arbitrary function generatorused in the room-temperature experiments is classified bythe manufacturer as ±1% of the nominal output voltage.Consequently, the maximum uncertainty in the peak voltagecorresponds to ±0.05 V for a maximum peak voltage ofUp = 5.0 V, which is equal to the voltage step size betweentwo horizontal lines of Fig. 3(a) in the main text.

As stated in the main text, we observed additional deviationsin the experimental data of sample B from the theory for veryfast voltage sweeps (1/β � 50 μs/V). These deviations origi-nate from a flattening of the triangular voltage ramps. Recordsof the triangular voltage pulse taken by an oscilloscoperevealed a flattening of the voltage apex depending on the peakvoltage Up, which becomes significant for very fast sweeps.This flattening translates into a peak voltage cutoff and hencea different value of Up, which is transduced to the sample.We attribute this to the limited bandwidth of the summationamplifier, which reduces the pulse fidelity for very short ramptimes. In the experiments conducted on sample A, a high-performance summation amplifier has been employed togetherwith a different arbitrary function generator. The latter exhibitsa greatly enhanced bandwidth and sampling rate (nearly oneorder of magnitude) compared to the device employed inthe room-temperature experiment. As a consequence, theflattening of the voltage pulse apex is less pronounced, andwe find good agreement between the experimental data andthe theory for inverse voltage sweep rates 1/β � 50 μs/V.

[1] E. C. G. Stuckelberg, Theorie der unelastischen Stosse zwischenAtomen, Helv. Phys. Acta 5, 369 (1932).

[2] S. N. Shevchenko, S. Ashhab, and F. Nori, Landau-Zener-Stuckelberg interferometry, Phys. Rep. 492, 1 (2010).

[3] S. Yoakum, L. Sirko, and P. M. Koch, Stuckelberg Oscillationsin the Multiphoton Excitation of Helium Rydberg Atoms:Observation with a Pulse of Coherent Field and Suppressionby Additive Noise, Phys. Rev. Lett. 69, 1919 (1992).

[4] M. Mark, T. Kraemer, P. Waldburger, J. Herbig, C. Chin,H.-C. Nagerl, and R. Grimm, Stuckelberg Interferometry withUltracold Molecules, Phys. Rev. Lett. 99, 113201 (2007).

[5] E. Dupont-Ferrier, B. Roche, B. Voisin, X. Jehl, R. Wacquez,M. Vinet, M. Sanquer, and S. De Franceschi, Coherent Couplingof two Dopants in a Silicon Nanowire Probed by Landau-Zener-Stuckelberg Interferometry, Phys. Rev. Lett. 110, 136802(2013).

[6] W. Wernsdorfer, R. Sessoli, A. Caneschi, D. Gatteschi, and A.Cornia, Nonadiabatic Landau-Zener tunneling in Fe8 molecularnanomagnets, Europhys. Lett. 50, 552 (2000).

[7] J. R. Petta, H. Lu, and A. C. Gossard, A coherent beam splitterfor electronic spin states, Science 327, 669 (2010).

[8] L. Gaudreau, G. Granger, A. Kam, G. C. Aers, S. A. Stu-denikin, P. Zawadzki, M. Pioro-Ladriere, Z. R. Wasilewski, andA. S. Sachrajda, Coherent control of three-spin states in a triplequantum dot, Nat. Phys. 8, 54 (2012).

[9] H. Ribeiro, G. Burkard, J. R. Petta, H. Lu, and A. C. Gossard,Coherent Adiabatic Spin Control in the Presence of ChargeNoise Using Tailored Pulses, Phys. Rev. Lett. 110, 086804(2013).

[10] F. Forster, G. Petersen, S. Manus, P. Hanggi, D. Schuh,W. Wegscheider, S. Kohler, and S. Ludwig, Characterizationof Qubit Dephasing by Landau-Zener-Stuckelberg-MajoranaInterferometry, Phys. Rev. Lett. 112, 116803 (2014).

[11] W. D. Oliver, Y. Yu, J. C. Lee, K. K. Berggren, L. S.Levitov, and T. P. Orlando, Mach-Zehnder interferometry ina strongly driven superconducting qubit, Science 310, 1653(2005).

[12] M. Sillanpaa, T. Lehtinen, A. Paila, Y. Makhlin, and P. Hakonen,Continuous-Time Monitoring of Landau-Zener Interference ina Cooper-Pair Box, Phys. Rev. Lett. 96, 187002 (2006).

[13] M. D. Lahaye, J. Suh, P. M. Echternach, K. C. Schwab, and M. L.Roukes, Nanomechanical measurements of a superconductingqubit, Nature (London) 459, 960 (2009).

[14] S. N. Shevchenko, S. Ashhab, and F. Nori, Inverse Landau-Zener-Stuckelberg problem for qubit-resonator systems, Phys.Rev. B 85, 094502 (2012).

[15] M. Gong, Y. Zhou, D. Lan, Y. Fan, J. Pan, H. Yu, J. Chen, G. Sun,Y. Yu, S. Han, and P. Wu, Landau-Zener-Stuckelberg-Majoranainterference in a 3D transmon driven by a chirped microwave,Appl. Phys. Lett. 108, 112602 (2016).

245406-9

Page 10: Classical Stuckelberg interferometry of a nanomechanical ...

MAXIMILIAN J. SEITNER et al. PHYSICAL REVIEW B 94, 245406 (2016)

[16] G. Heinrich, J. G. E. Harris, and F. Marquardt, Photon shut-tle: Landau-Zener-Stuckelberg dynamics in an optomechanicalsystem, Phys. Rev. A 81, 011801 (2010).

[17] N. V. Vitanov and B. M. Garraway, Landau-Zener model: Effectsof finite coupling duration, Phys. Rev. A 53, 4288 (1996).

[18] L. D. Landau, Zur theorie der energieubertragung. II, Physics ofthe Soviet Union 2, 46 (1932).

[19] C. Zener, Non-adiabatic crossing of energy levels, Proc. R. Soc.London, Ser. A 137, 696 (1932).

[20] E. Majorana, Atomi orientati in campo magnetico variabile,Nuovo Cimento 9, 43 (1932).

[21] H. Okamoto, A. Gourgout, C.-Y. Chang, K. Onomitsu, I.Mahboob, E. Y. Chang, and H. Yamaguchi, Coherent phononmanipulation in coupled mechanical resonators, Nat. Phys. 9,480 (2013).

[22] T. Faust, J. Rieger, M. J. Seitner, J. P. Kotthaus, and E. M.Weig, Coherent control of a classical nanomechanical two-levelsystem, Nat. Phys. 9, 485 (2013).

[23] A. B. Shkarin, N. E. Flowers-Jacobs, S. W. Hoch, A. D.Kashkanova, C. Deutsch, J. Reichel, and J. G. E. Harris,Optically Mediated Hybridization Between Two MechanicalModes, Phys. Rev. Lett. 112, 013602 (2014).

[24] T. Faust, J. Rieger, M. J. Seitner, P. Krenn, J. P. Kotthaus,and E. M. Weig, Nonadiabatic Dynamics of two StronglyCoupled Nanomechanical Resonator Modes, Phys. Rev. Lett.109, 037205 (2012).

[25] H. J. Maris and Q. Xiong, Adiabatic and nonadiabatic processesin classical and quantum mechanics, Am. J. Phys. 56, 1114(1988).

[26] B. W. Shore, M. V. Gromovyy, L. P. Yatsenko, and V. I.Romanenko, Simple mechanical analogs of rapid adiabaticpassage in atomic physics, Am. J. Phys. 77, 1183 (2009).

[27] G. Sun, X. Wen, B. Mao, Y. Yu, J. Chen, W. Xu, L. Kang,P. Wu, and S. Han, Landau-Zener-Stuckelberg interference ofmicrowave-dressed states of a superconducting phase qubit,Phys. Rev. B 83, 180507(R) (2011).

[28] J. Rieger, T. Faust, M. J. Seitner, J. P. Kotthaus, and E. M. Weig,Frequency and Q factor control of nanomechanical resonators,Appl. Phys. Lett. 101, 103110 (2012).

[29] T. Faust, P. Krenn, S. Manus, J. P. Kotthaus, and E. M. Weig,Microwave cavity-enhanced transduction for plug and playnanomechanics at room temperature, Nat. Commun. 3, 728(2012).

[30] L. Novotny, Strong coupling, energy splitting, and level cross-ings: A classical perspective, Am. J. Phys. 78, 1199 (2010).

[31] V.I. Arnold, Mathematical Methods of Classical Mechanics,2nd ed. (Springer, New York, 1989), Pt. III, Chap. 8, Sec. 38.

[32] M. J. Seitner, H. Ribeiro, J. Kolbl, T. Faust, and E. M. Weig,Finite time Stuckelberg interferometry with nanomechanicalmodes, arXiv:1609.05023 [cond-mat.mes-hall].

[33] T. Usuki, Theoretical study of Landau-Zener tunneling at theM+N level crossing, Phys. Rev. B 56, 13360 (1997).

[34] M. O. Scully and M. Suhail Zubairy, Quantum Optics(Cambridge University Press, Cambridge, 1997).

[35] A. D. O’Connell, M. Hofheinz, M. Ansmann, R. C. Bialczak, M.Lenander, E. Lucero, M. Neeley, D. Sank, H. Wang, M. Weides,J. Wenner, J. M. Martinis, and A. N. Cleland, Quantum groundstate and single-phonon control of a mechanical resonator,Nature (London) 464, 697 (2010).

[36] J. D. Teufel, T. Donner, D. Li, J. W. Harlow, M. S. Allman, K.Cicak, A. J. Sirois, J. D. Whittaker, K. W. Lehnert, and R. W.Simmonds, Sideband cooling of micromechanical motion to thequantum ground state, Nature (London) 475, 359 (2011).

[37] H. Xu, D. Mason, L. Jiang, and J. G. E. Harris, Topologicalenergy transfer in an optomechanical system with exceptionalpoints, Nature (London) 537, 80 (2016).

[38] N. Rossi, F. R. Braakman, D. Cadeddu, D. Vasyukov, G.Tutuncuoglu, A. F. i Morral, and M. Poggio, Vectorial scanningforce microscopy using a nanowire sensor, Nat. Nanotechnol.(2016), doi:10.1038/nnano.2016.189.

[39] L. Mercier de Lepinay, B. Pigeau, B. Besga, P. Vincent,P. Poncharal, and O. Arcizet, A universal and ultrasensitivevectorial nanomechanical sensor for imaging 2D force fields,Nat. Nanotechnol. (2016), doi:10.1038/nnano.2016.193.

[40] M. R. Andrews, C. G. Townsend, H.-J. Miesner, D. S. Durfee,D. M. Kurn, and W. Ketterle, Observation of interferencebetween two Bose condensates, Science 275, 637 (1997).

[41] Y. Castin and J. Dalibard, Relative phase of two Bose-Einsteincondensates, Phys. Rev. A 55, 4330 (1997).

[42] C. Kohstall, S. Riedl, E. R. Sanchez Guajardo, L. A. Sidorenkov,J. H. Denschlag, and R. Grimm, Observation of interferencebetween two molecular Bose-Einstein condensates, New J. Phys.13, 065027 (2011).

[43] O. Maillet, F. Vavrek, A. D. Fefferman, O. Bourgeois, and E.Collin, Classical decoherence in a nanomechanical resonator,New J. Phys. 18, 073022 (2016).

245406-10