Abstract
Mechanical resonators are promising systems for storing and manipulating information. To transfer information between mechanical modes, either direct coupling or an interface between these modes is needed. In previous works, strong coupling between different modes in a single mechanical resonator and direct interaction between neighboring mechanical resonators have been demonstrated. However, coupling between distant mechanical resonators, which is a crucial request for long-distance classical and quantum information processing using mechanical devices, remains an experimental challenge. Here, we report the experimental observation of strong indirect coupling between separated mechanical resonators in a graphene-based electromechanical system. The coupling is mediated by a far-off-resonant phonon cavity through virtual excitations via a Raman-like process. By controlling the resonant frequency of the phonon cavity, the indirect coupling can be tuned in a wide range. Our results may lead to the development of gate-controlled all-mechanical devices and open up the possibility of long-distance quantum mechanical experiments.
Non-neighbouring mechanical resonators can interact via indirect coupling. Here, the authors leverage a resonant phonon cavity in a graphene-based electromechanical system to demonstrate strong indirect coupling between separated mechanical resonators.
Introduction
The rapid development of nanofabrication technology enables the storage and manipulation of phonon states in micro- and nano-mechanical resonators1–5. Mechanical resonators with quality factors6 exceeding 5 million and frequencies7,8 in the sub-gigahertz range have been reported. These advances have paved the route to controllable mechanical devices with ultralong memory time9. To transfer information between different mechanical modes, tunable interactions between these modes are required10. While different modes in a single mechanical resonator can be coupled by parametric pump3,4,11–16 and neighboring mechanical resonators can be coupled via phonon processes through the substrate2 or direct contact interaction17, it is challenging to directly couple distant mechanical resonators.
Here, we observe strong effective coupling between mechanical resonators separated at a distance via a phonon cavity that is significantly detuned from these two resonator modes. The coupling is generated via a Raman-like process through virtual excitations in the phonon cavity and is tunable by varying the frequency of the phonon cavity. Typically, a Raman process can be realized in an atom with three energy levels in the Λ form18,19. The two lower energy levels are each coupled to the third energy level via an optical field with detunings. When these two detunings are tuned to be equal to each other, an effective coupling is formed between the lower two levels. To our knowledge, tunable indirect coupling in electro-mechanical systems has not been demonstrated before. The physical mechanism of this coupling is analogous to the coupling between distant qubits in circuit quantum electrodynamics20,21, where the interaction between qubits is induced by virtual photon exchange via a superconducting microwave resonator.
Results
Sample characterization
The sample structure is shown in Fig. 1a, where a graphene ribbon22,23 with a width of ~1 μm and ~5 layers is suspended over three trenches (2 μm in width, 150 nm in depth) between four metal (Ti/Au) electrodes. This configuration defines three distinct electromechanical resonators: R1, R2 and R3. The metallic contacts S and D3 are each 2 μm wide and D1 and D2 are each 1.5 μm wide, which leads to a 7-μm separation between the centers of R1 and R3 (see Supplementary Methods and Supplementary Fig. 1). All measurements are performed in a dilution refrigerator at a base temperature of approximately 10 mK and at pressures below 10−7 torr. The suspended resonators are biased by a dc gate voltage ( for the ith resonator) and actuated by an ac voltage ( for the ith resonator with driving frequency fgi = ωd/2π) through electrodes (gi for the ith resonator) underneath the respective resonators. To characterize the spectroscopic properties of the resonators, a driving tone is applied to one or more of the bottom gates with frequency ωd, and another microwave tone with frequency ωd + δω is applied to the contact S. A mixing current (Imix = Ix + jIy) can then be obtained at D3 (D1 and D2 are floated during all measurements) by detecting the δω signal with a lock-in amplifier fixed at zero phase during all measurements (see Supplementary Methods and Supplementary Fig. 2).
Figure 1b shows the measured mixing current as a function of the dc gate voltage and the ac driving frequency on R3, where the oblique lines represent the resonant frequencies of the resonator modes. We denote the resonant frequency of the ith resonator as fmi = ωmi/2π. This plot shows that when . The frequencies of the resonators can hence be tuned in a wide range (see Supplementary Note 1 and Supplementary Fig. 3 for results of R1 and R2), which allows us to adjust the mechanical modes to be on or off resonance with each other. The quality factors (Q) of the resonant modes are determined by fitting the measured spectral widths (see Supplementary Fig. 8) at low driving powers (typically −50 dBm). Figure 1c shows the spectral dependence of R3, which gives a linewidth of γ3/2π ~ 28 kHz at a resonant frequency of fm3 ~ 98.05 MHz. The resulting quality factor is Q ~ 3500. The quality factors of the other two resonators are similar, at ~3000.
Strong coupling between neighboring resonators
Neighboring resonators in this system couple strongly with each other, similar to previous studies on gallium arsenide2 and carbon nanotube17. Figure 1d, e shows the spectra of the coupled modes (R1, R2) and (R2, R3), respectively, by plotting the mixed current Ix as a function of gate voltages and driving frequencies. In Fig. 1e, the voltage is fixed at 10.5 V, with a corresponding resonant frequency fm3 = 101.15 MHz, and is scanned over a range with fm2 being near-resonant to fm3. A distinct avoided level crossing appears when fm2 approaches fm3, which is a central feature of two resonators with direct coupling. From the measured data, we extract the coupling rate between these two modes as Ω23/2π ~ 200 kHz, which is the energy splitting when fm2 = fm3. In Supplementary Note 2 and Supplementary Fig. 4, we fit the measured spectrum with a single two-mode model using this coupling rate. Similarly in Fig. 1d, by fixing at 10.5 V and scanning the voltage , we obtain the coupling rate between R1 and R2 as Ω12/2π ~ 240 kHz. There are several possible origins for the coupling between two adjacent resonators in this system. One coupling medium is the substrate and the other medium is the graphene ribbon itself. Mechanical energy can be transferred in a solid-state material by phonon propagation, as demonstrated in several experiments2,17,24. Second, because adjacent resonators share lattice bonds, the phonon energy can transfer in the graphene ribbon. The dependence of the coupling strength on the width of the drain contacts is still unknown (see Supplementary Fig. 5 for another sample).
The measured coupling strength satisfies the strong coupling condition with . Defining the cooperativity for this phonon–phonon coupling system as , we find that C = 44. A similar strong coupling condition can be found between modes R1 and R2. By adjusting the gate voltages of these three resonators, R2 can be successively coupled to both R1 and R3 (see Fig. 1f).
For comparison, we study the coupling strength between modes R1 and R3. The frequency fm2 of resonator R2 is set to be detuned from fm1 and fm3 by 700 kHz in Fig. 2a. In the dashed circle, we observe a near-perfect level crossing when fm1 approaches fm3, which indicates a negligible coupling between these two modes, with (also see Supplementary Fig. 6).
Raman-like coupling between well-separated resonators
The three resonator modes in our system are in the classical regime. The Hamiltonian of these three classical resonators can be written as:
1 |
where is a coupling parameter between i- and jth resonators, ppi is the effective momentum and xi is the effective coordinate of the oscillation for the ith resonator, respectively. Let and , with αi and being complex numbers. The Hamiltonian in Eq. (1) can be written as
2 |
Here, we have applied the rotating-wave approximation and neglected the and terms. This approximation is valid when . This Hamiltonian describes the direct couplings between neighboring resonators (R1, R2) and (R2, R3). Through these couplings, the mechanical modes hybridize into three normal modes, and an effective coupling between modes R1 and R3 can be achieved. If the resonators work in the quantum regime, and can be quantized into the annihilation and creation operators of a quantum harmonic oscillator, respectively.
We study the hybridization of this three-mode system by fixing the gate voltages (mode frequencies) of modes R1 and R2, and sweeping the gate voltage of R3 over a wide range. The spectrum of this system depends strongly on the detuning between modes R1 and R2, which is defined as Δ12 = 2π(fm2−fm1). In Fig. 2a, Δ12/2π ~ 70 kHz. Similar to Fig. 1f, modes R3 and R1 show a level crossing. Moreover, we observe a large avoided level crossing between modes R2 and R3 when the frequency fm3 approaches fm2, indicating strong coupling between these two modes. Hence, even with strong couplings between all neighboring resonators, the effective coupling between the distant modes R1 and R3 is still negligible when the frequency of mode R2 is significantly far off resonance from the other two modes. On the contrary, when the detuning Δ12/2π is lowered to ~180 kHz, a distinct avoided level crossing between modes R1 and R3 is observed, as shown inside the dashed circle in Fig. 2b.
With coupling strengths Ω12/2π = 240 kHz and Ω23/2π = 170 kHz extracted from the measured data, we plot the theoretical spectra of the normal modes in this three-mode system given by Eq. (2), for Δ12/2π = 700 and 180 kHz in Fig. 2c, d, respectively. Our result shows good agreement between theoretical and experimental results.
With direct couplings between neighboring resonators, an effective coupling between the two distant resonators R1 and R3 can be obtained via their couplings to mode R2. The effective coupling can be viewed as a Raman process, as illustrated in Fig. 3a. Here mode R2 functions as a phonon cavity that connects the mechanical resonators R1 and R3 via virtual phonon excitations. The physical mechanism of this effective coupling is similar to that of the coupling between distant superconducting qubits via a superconducting microwave cavity20. The detuning between the phonon cavity and the other two modes Δ12 can be used as a control parameter to adjust this effective coupling.
To derive the effective coupling, we consider the case of Δ12 = Δ32 = Δ, where Δ32/2π = fm2−fm3 and . The avoided level crossing between modes R1 and R3 can be extracted at this point. Using a perturbation theory approach, we obtain the effective Hamiltonian between modes R1 and R3 as (see Methods for details)
3 |
Here, an effective coupling is generated between R1 and R3 with magnitude Ω13 = Ω12Ω23/2Δ, and the resonant frequencies of each mode are shifted by a small term. The effective coupling Ω13 in the Hamiltonian depends strongly on the detuning Δ. Thus, the effective coupling between R1 and R3 can be controlled over a wide range by varying the frequency (gate voltage) of resonator R2.
The effective coupling strength Ω13 between R1 and R3 as a function of Δ12 is shown in Fig. 3b. Each data point is obtained by changing the gate voltage of R2 and repeating the measurements in Fig. 2a, b (see Supplementary Fig. 7). Over a large range of detuning, the effective coupling is larger than the linewidths of the resonators γ1,2,3/2π, with Ω13 > 30 kHz. The red line shows the results using perturbation theory. The experimental data indicate good agreement with the theoretical results.
Discussion
In summary, we have demonstrated indirect coupling between separated mechanical resonators in a three-mode electromechanical system constructed from a graphene ribbon. Our study suggests that coupling between well-separated mechanical modes can be created and manipulated via a phonon cavity. These observations hold promise for a wide range of applications in phonon state storage, transmission, and transformation. In the current experiment, the sample works in an environment subjected to noise and microwave heating with typical temperatures as high as 100 mK and phonon numbers reaching ~24. By cooling the mechanical resonators to lower temperatures25–28, quantum states could be manipulated via this indirect coupling29,30. Furthermore, in the quantum limit, by coupling the mechanical modes to solid-state qubits, such as quantum-dots and superconducting qubits17,31,32, this system can be utilized as a quantum data bus to transfer information between qubits33,34. Future work may lead to the development of graphene-based mechanical resonator arrays as phononic waveguides24 and quantum memories35 with high tunabilities.
Methods
Theory of three-mode coupling
We describe this three-mode system with the Hamiltonian (ħ = 1)
4 |
where is the coupling between mechanical resonators i and j. The couplings between the resonators induce hybridization of the three modes. The hybridized normal modes under this Hamiltonian can be obtained by solving the eigenvalues of the matrix
5 |
where Δij/2π = fmi−fmj is the frequency difference between Ri and Rj. The eigenvalues of this matrix correspond to the frequencies of the normal modes, i.e., the peaks in the spectroscopic measurement.
We consider the special case of Δ12 = Δ23 = Δ, with , in the three-mode system. Here, the eigenvalues of the normal modes can be derived analytically. One eigenvalue is ωΔ = Δ, which corresponds to the eigenmode
6 |
This mode is a superposition of the end modes α1 and α3, and does not include the middle mode. The two other eigenvalues are
7 |
with . The corresponding normal modes are
8 |
With , for Δ > 0, . The mode αΔ+ is nearly degenerate with αΔ, and
9 |
The mode αΔ− has frequency , with αΔ− ≈ α2. The normal modes now become separated into two nearly degenerate modes {αΔ,αΔ+}, which are superpositions of modes α1 and α3, and a third mode αΔ− that is significantly off resonance from the other two modes. The nearly degenerate modes can be viewed as a hybridization of α1 and α3 with an effective splitting . A similar result can be derived for Δ < 0, where , with αΔ− given by the expression in Eq. (8), and with αΔ+ ≈ α2.
The effective coupling rate can be derived with a perturbative approach on the matrix M. When |Δ| Ω12, Ω23, the dynamics of α1 and α3 is governed by matrix
10 |
This matrix tells us that because of their interaction with the middle mode α2, the frequency of mode α1 (α3) is shifted by (), which is much smaller than |Δ|. Meanwhile, an effective coupling is generated between these two modes with magnitude . The effective Hamiltonian for α1 and α3 can be written as
11 |
The effective coupling can be controlled over a wide range by varying the frequency of the second mode α2.
Data availability
The remaining data contained within the paper and Supplementary files are available from the author upon request.
Electronic supplementary material
Acknowledgements
This work was supported by the National Key R&D Program of China (Grant No. 2016YFA0301700), the NSFC (Grants Nos. 11625419, 61704164, 61674132, 11674300, 11575172, and 91421303), the SPRP of CAS (Grant No. XDB01030000), and the Fundamental Research Fund for the Central Universities. L.T. is supported by the National Science Foundation under Award No. DMR-0956064 and PHY-1720501 and the UC Multicampus-National Lab Collaborative Research and Training under Award No. LFR-17-477237. This work was partially carried out at the USTC Center for Micro and Nanoscale Research and Fabrication.
Author contributions
G.L. and Z.-Z.Z. fabricated the device. G.-W.D. and Z.-Z.Z. performed the measurements. L.T., G.-W.D, and Z.-Z.Z analyzed the data and developed the theoretical analysis. H.-O.L, G.C., M.X., and G.-C.G. supported the fabrication and measurement. G.-P.G. and G.-W.D. planned the project. All authors participated in writing the manuscript.
Competing interests
The authors declare no competing financial interests.
Footnotes
Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Gang Luo and Zhuo-Zhi Zhang contributed equally to this work.
Change history
3/19/2019
The original version of this Article contained a number of errors. As a result of this, changes have been made to both the PDF and the HTML versions of the Article. A full list of these changes is available online
Contributor Information
Guang-Wei Deng, Email: gwdeng@ustc.edu.cn.
Lin Tian, Email: ltian@ucmerced.edu.
Guo-Ping Guo, Email: gpguo@ustc.edu.cn.
Electronic supplementary material
Supplementary Information accompanies this paper at 10.1038/s41467-018-02854-4.
References
- 1.Metcalfe M. Applications of cavity optomechanics. Appl. Phys. Rev. 2014;1:031105. doi: 10.1063/1.4896029. [DOI] [Google Scholar]
- 2.Hajime O, et al. Coherent phonon manipulation in coupled mechanical resonators. Nat. Phys. 2013;9:480–484. doi: 10.1038/nphys2665. [DOI] [Google Scholar]
- 3.Faust T, Rieger J, Seitner MJ, Kotthaus JP, Weig EM. Coherent control of a classical nanomechanical two-level system. Nat. Phys. 2013;9:485–488. doi: 10.1038/nphys2666. [DOI] [Google Scholar]
- 4.Zhu D, et al. Coherent phonon Rabi oscillations with a high-frequency carbon nanotube phonon cavity. Nano Lett. 2017;17:915–921. doi: 10.1021/acs.nanolett.6b04223. [DOI] [PubMed] [Google Scholar]
- 5.Aspelmeyer M, Meystre P, Schwab K. Quantum optomechanics. Phys. Today. 2012;65:29–35. doi: 10.1063/PT.3.1640. [DOI] [Google Scholar]
- 6.Moser J, Eichler A, Guttinger J, Dykman MI, Bachtold A. Nanotube mechanical resonators with quality factors of up to 5 million. Nat. Nanotechnol. 2014;9:1007–1011. doi: 10.1038/nnano.2014.234. [DOI] [PubMed] [Google Scholar]
- 7.Sazonova V, et al. A tunable carbon nanotube electromechanical oscillator. Nature. 2004;43:284–287. doi: 10.1038/nature02905. [DOI] [PubMed] [Google Scholar]
- 8.Chen CY, et al. Graphene mechanical oscillators with tunable frequency. Nat. Nanotechnol. 2013;8:923–927. doi: 10.1038/nnano.2013.232. [DOI] [PubMed] [Google Scholar]
- 9.Mahboob I, Yamaguchi H. Bit storage and bit flip operations in an electromechanical oscillator. Nat. Nanotechnol. 2008;3:275–279. doi: 10.1038/nnano.2008.84. [DOI] [PubMed] [Google Scholar]
- 10.Cirac JI, Zoller P, Kimble HJ, Mabuchi H. Quantum state transfer and entanglement distribution among distant nodes in a quantum network. Phys. Rev. Lett. 1997;78:3221–3224. doi: 10.1103/PhysRevLett.78.3221. [DOI] [Google Scholar]
- 11.Eichler A, del Álamo Ruiz M, Plaza JA, Bachtold A. Strong coupling between mechanical modes in a nanotube resonator. Phys. Rev. Lett. 2012;109:025503. doi: 10.1103/PhysRevLett.109.025503. [DOI] [PubMed] [Google Scholar]
- 12.Liu CH, Kim IS, Lauhon LJ. Optical control of mechanical mode-coupling within a MoS2 resonator in the strong-coupling regime. Nano Lett. 2015;15:6727–6731. doi: 10.1021/acs.nanolett.5b02586. [DOI] [PubMed] [Google Scholar]
- 13.Li SX, et al. Parametric strong mode-coupling in carbon nanotube mechanical resonators. Nanoscale. 2016;8:14809–14813. doi: 10.1039/C6NR02853E. [DOI] [PubMed] [Google Scholar]
- 14.Mathew JP, Patel RN, Borah A, Vijay R, Deshmukh MM. Dynamical strong coupling and parametric amplification of mechanical modes of graphene drums. Nat. Nanotechnol. 2016;11:747–751. doi: 10.1038/nnano.2016.94. [DOI] [PubMed] [Google Scholar]
- 15.De Alba R, et al. Tunable phonon-cavity coupling in graphene membranes. Nat. Nanotechnol. 2016;11:741–746. doi: 10.1038/nnano.2016.86. [DOI] [PubMed] [Google Scholar]
- 16.Castellanos-Gomez A, Meerwaldt HB, Venstra WJ, van der Zant HSJ, Steele GA. Strong and tunable mode coupling in carbon nanotube resonators. Phys. Rev. B. 2012;86:041402. doi: 10.1103/PhysRevB.86.041402. [DOI] [Google Scholar]
- 17.Deng GW, et al. Strongly coupled nanotube electromechanical resonators. Nano Lett. 2016;16:5456–5462. doi: 10.1021/acs.nanolett.6b01875. [DOI] [PubMed] [Google Scholar]
- 18.Raman CV, Krishnan KS. A new type of secondary radiation. Nature. 1928;121:501–502. doi: 10.1038/121501c0. [DOI] [Google Scholar]
- 19.Gaubatz U, Rudecki P, Schiemann S, Bergmann K. Population transfer between molecular vibrational levels by stimulated raman-scattering with partially overlapping laserfields - a new concept and experimental results. J. Chem. Phys. 1990;92:5363–5376. doi: 10.1063/1.458514. [DOI] [Google Scholar]
- 20.Majer J, et al. Coupling superconducting qubits via a cavity bus. Nature. 2007;449:443–447. doi: 10.1038/nature06184. [DOI] [PubMed] [Google Scholar]
- 21.Deng GW, et al. Coupling two distant double quantum dots with a microwave resonator. Nano Lett. 2015;15:6620–6625. doi: 10.1021/acs.nanolett.5b02400. [DOI] [PubMed] [Google Scholar]
- 22.Bunch JS, et al. Electromechanical resonators from graphene sheets. Science. 2007;315:490–493. doi: 10.1126/science.1136836. [DOI] [PubMed] [Google Scholar]
- 23.Chen CY, et al. Performance of monolayer graphene nanomechanical resonators with electrical readout. Nat. Nanotechnol. 2009;4:861–867. doi: 10.1038/nnano.2009.267. [DOI] [PubMed] [Google Scholar]
- 24.Hatanaka D, Mahboob I, Onomitsu K, Yamaguchi H. Phonon waveguides for electromechanical circuits. Nat. Nanotechnol. 2014;9:520–524. doi: 10.1038/nnano.2014.107. [DOI] [PubMed] [Google Scholar]
- 25.Teufel JD, et al. Sideband cooling of micromechanical motion to the quantum ground state. Nature. 2011;475:359–363. doi: 10.1038/nature10261. [DOI] [PubMed] [Google Scholar]
- 26.Chan J, et al. Laser cooling of a nanomechanical oscillator into its quantum ground state. Nature. 2011;478:89–92. doi: 10.1038/nature10461. [DOI] [PubMed] [Google Scholar]
- 27.Kepesidis KV, Bennett SD, Portolan S, Lukin MD, Rabl P. Phonon cooling and lasing with nitrogen-vacancy centers in diamond. Phys. Rev. B. 2013;88:064105. doi: 10.1103/PhysRevB.88.064105. [DOI] [Google Scholar]
- 28.Stadler P, Belzig W, Rastelli G. Ground-state cooling of a carbon nanomechanical resonator by spin-polarized current. Phys. Rev. Lett. 2014;113:047201. doi: 10.1103/PhysRevLett.113.047201. [DOI] [PubMed] [Google Scholar]
- 29.Jähne K, et al. Cavity-assisted squeezing of a mechanical oscillator. Phys. Rev. A. 2009;79:063819. doi: 10.1103/PhysRevA.79.063819. [DOI] [Google Scholar]
- 30.Palomaki TA, Teufel JD, Simmonds RW, Lehnert KW. Entangling mechanical motion with microwave fields. Science. 2013;342:710–713. doi: 10.1126/science.1244563. [DOI] [PubMed] [Google Scholar]
- 31.Steele GA, et al. Strong coupling between single-electron tunneling and nanomechanical motion. Science. 2009;325:1103–1107. doi: 10.1126/science.1176076. [DOI] [PubMed] [Google Scholar]
- 32.Lassagne B, Tarakanov Y, Kinaret J, Garcia-Sanchez D, Bachtold A. Coupling mechanics to charge transport in carbon nanotube mechanical resonators. Science. 2009;325:1107–1110. doi: 10.1126/science.1174290. [DOI] [PubMed] [Google Scholar]
- 33.Tian L, Zoller P. Coupled ion-nanomechanical systems. Phys. Rev. Lett. 2004;93:266403. doi: 10.1103/PhysRevLett.93.266403. [DOI] [PubMed] [Google Scholar]
- 34.Schneider BH, Etaki S, van der Zant HSJ, Steele GA. Coupling carbon nanotube mechanics to a superconducting circuit. Sci. Rep. 2012;2:599. doi: 10.1038/srep00599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Zhang XF, et al. Magnon dark modes and gradient memory. Nat. Commun. 2015;6:8914. doi: 10.1038/ncomms9914. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The remaining data contained within the paper and Supplementary files are available from the author upon request.