arXiv:2007.14851v1 [quant-ph] 29 Jul 2020 hnclsses[ systems chanical btcei h xsec ftedr-oeefc [ effect dark-mode the of existence the is obstacle oeculdt w ehnclmds hsi realized inter- is phonon-exchange This phase-dependent a modes. introducing mechanical by two to coupled mode by effect modes mechanical multiple the of realize to method mds ope oacmo aiyfil,a demonstrated [ as field, theoretically cavity common a to coupled (modes) cavity 42 in challenge outstanding [ an optomechanics [ remains resonator mechanical resonators multiple mechanical of ground-state cooling single ground-state in task. simultaneous a made important of been have and cooling advances obligatory great an Though becomes mechanical these resonators of cooling ground-state simultaneous [ the sensors high-performance 28 as used be 25 can also but only macro- [ investigating not effects for resonators [ coherence platform These mechanical scopic promising to a coupling devices. provide tunable physical and diverse compatibility, wide resonance, Introduction. nti ai omncto,w rps reliable a propose we Communication, Rapid this In – applications, those in noise thermal suppress To [ ]. transducers ], 45 na poehnclsse ossigo cavity a of consisting system optomechanical an in nue ytemlil ehnclresonators mechanical multiple the by induced ] 16 – 1 e aoaoyo o-iesoa unu tutrsadQ and Structures Quantum Low-Dimensional of Laboratory Key 21 h w ehnclmds eas eeaieti ehdt c physics. in to effects method valid dark-state general and this has dark-mode generalize scheme other cooling also this We in mechanism physical modes. mechanical lead which two mechanism, the transfer energy nonreciprocal mechanical the two fo on the to for performance used me cooling is asymmetrical nondegenerate dark-mode-breakin an which or reliable interaction, degenerate phonon-exchange and th two dependent universal through of mod modes cooling a mechanical dark ground-state these propose the by we from formed energy Here modes extracting dark outstandi prevent an the and become because has is mode cavity-field This common a to coupled 42 ,adtplgcleeg rnfr[ transfer energy topological and ], h iutnosgon-tt oln fmlil degener multiple of cooling ground-state simultaneous The Mcaia eoaosi aiyoptome- cavity in resonators —Mechanical eateto hsc n yegtcInvto etrfor Center Innovation Synergetic and Physics of Department orcpoa rudsaecoigo utpemechanica multiple of cooling ground-state Nonreciprocal , 42 43 7 – 26 hsc eatet h nvriyo ihgn n Arbor, Ann Michigan, of University The Department, Physics n xeietly[ experimentally and ] 1 44 – ,admcaia optr [ computers mechanical and ], 3 .Tepyia rgnbhn this behind origin physical The ]. iutnosgon-tt cooling ground-state simultaneous 3 aeteavnae feasy of advantages the have ] olg fPyisadEetoi niern,Isiueo Institute Engineering, Electronic and Physics of College 2 e aoaoyfrMte irsrcueadFnto fHu of Function and Microstructure Matter for Laboratory Key hoeia unu hsc aoaoy IE,Siaa35 Saitama RIKEN, Laboratory, Physics Quantum Theoretical elnLi, Wenlin egGoLai, Deng-Gao 4 – 15 6 raigtedark-mode the breaking nvriyo aeio -23 aeio(C,Italy (MC), Camerino I-62032 Camerino, of University N-N,Lg nioFri6 -02 iez,Italy Firenze, I-50125 6, Fermi Enrico L.go CNR-INO, 4 ,qatmmany-body quantum ], iha omlUiest,Cegu606,China 610068, Chengdu University, Normal Sichuan ua omlUiest,Cagh 101 China 410081, Changsha University, Normal Hunan colo cec n ehooy hsc Division, Physics Technology, and Science of School 5 4 NN ein iPrga -62 eui,Italy Perugia, I-06123 Perugia, di Sezione INFN, ai Vitali, David ,2 1, 44 , i-egHuang, Jin-Feng 45 29 ]. – 41 ,5 6 5, 4, ,the ], 22 23 27 8 ], rnoNori, Franco – , , 1 where cinbtentetomcaia oe [ modes mechanical two the between action n hsavnewl ehlflfrteminiaturization the for [ helpful devices be quantum will of advance this and tutr [Fig. structure ope ihec te i hs-eedn phonon- frequency phase-dependent with a [ are via interaction which other exchange modes, each mechanical with two coupled to coupled chanically two the between We transfer [ modes excitation caused effect. mechanical is nonreciprocal interference performance cooling an by asymmetrical the via that find realized mechanical two is the of resonators and cooling anymore, state mode ground dark asymmetrical no is this there in system, loop-coupled interaction phonon-exchange phase-dependent the h aksaeo akmd fet nohrphysical other in effects break dark-mode to [ or generalized systems be dark-state can and the universal is mechanism of cooling simultaneous the to ytmHmloinras( reads Hamiltonian system aiy narttn rm endby defined frame rotating a In cavity. inL Yin, Xian-Li System. H otedrcinlflwo hnn between phonons of flow directional the to s t n hsmto a eetne obreak to extended be can method this and ity oe n xon hspeoeo based phenomenon this expound and modes malo-ope ofiuain efind We configuration. loop-coupled a rm I t rna-eeeaemcaia modes mechanical near-degenerate or ate ,7 2, hnclmdsb nrdcn phase- a introducing by modes chanical a ∆ = atmCnrlo iityo Education, of Ministry of Control uantum ehdt elz h simultaneous the realize to method g 46 oln hne ftecvt mode. cavity the of channel cooling e W osdratremd optomechanical three-mode a consider —We unu fet n Applications, and Effects Quantum gcalnei aiyoptomechanics. cavity in challenge ng o utpemcaia oe.The modes. mechanical multiple ool n i-ioLiao Jie-Qiao and ( +(Ω a ]. sdcul rmtecvt mode cavity the from decouple es † c 1 ihgn41914,USA 48109-1040, Michigan and ) a 1 ω oi tt Physics, State Solid f † a]cnitn facvt edoptome- field cavity a of consisting (a)] agPnHou, Bang-Pin a a L Ω + + n amplitude and 51 -18 Japan 1-0198, a Province, nan l X =1 b 46 66 ∗ – l a =1 65 , .Amncrmtcdiigfield driving monochromatic A ]. , 2 † [ + ) , 67 ω .W loetn hsmethod this extend also We ]. 2 resonators l ~ l .Ti dark-mode-breaking This ]. b 1, 1 = ( l † η b b ∗ 3 l † ( l N r,rsetvl,the respectively, are, ) e Ω [ ) + iθ b sapidt h optical the to applied is 46 ehnclresonators mechanical g 1 † l exp( b a ] 2 † a + ( − b e l 46 − iω + iθ .Oigto Owing ]. L b b ta 2 † l † b )] † 1 a ) , ) the , (1) =0

2

a b 3 3 ωL Ω 2 (a) Dark-mode-unbreaking 2 (b) =0=0.2, =0 † † † † Dark-mode-breaking c Phonon numbers 0.8 1.2 0.8 1.2 † c 1 1 H.c. D/w γ γ2 m γ γ 1 3 1 2 (b) f (e) =0.05 (c) n1 2 2 0.1 10 FIG. 1. (a) A loop-coupled optomechanical system consists 10 of one cavity-field mode a optomechanically coupled to two (d) n f mechanical modes b and b , which are coupled with each 2 1 2 0.1 other via a phase-dependent phonon-exchange coupling (with 1 the coupling strength η and phase θ). (b) The reduced two- 0.2 mechanical-mode system with the effective phonon-exchange channel (χl=1,2), the common optomechanical-cooling channel (γ , nopt), and the mechanical dissipations (γ , n¯ ). l,opt l=1,2 l FIG. 2. (Color online) (a) The final average phonon numbers f f n1 (blue curves) and n2 (red curves) in the two mechanical resonators versus the effective driving detuning ∆ in the dark- annihilation (creation) operators of the cavity mode (ωc) and the lth mechanical mode (ω ). The g terms mode-unbreaking (η = 0, solid curves) and -breaking (η = l l=1,2 0.05ω and θ = π/2, dashed curves) cases when ω = ω = describe the optomechanical couplings. The Ω term m 1 2 ω . (b) nf and nf as functions of ω /ω in both the dark- denotes the cavity-field driving with detuning m 1 2 2 1 ∆c = mode-unbreaking (solid curves) and -breaking (dashed curves) ωc ωL, and the η term describes a phase-dependent f f − cases when ∆ = ω1. (c) n1 and (d) n2 vs η and θ under phonon-exchange interaction between the two mechanical f the optimal driving ∆ = ωm and ω1 = ω2 = ωm. (e) n1 resonators, with the real coupling strength η and phase θ. f Note that this model can be implemented with either circuit and n2 vs θ at η = 0.05ωm. Other used parameters are given −5 by G1/ωm = G2/ωm = 0.1, γ1/ωm = γ2/ωm = 10 , electromechanical systems [8, 23] or photonic crystal 3 optomechanical cavity systems [51]. The phase-dependent κ/ωm = 0.2, and n¯1 =n ¯2 = 10 . phonon-hopping coupling in the electromechanical system can be indirectly induced by coupling to a charge dark mode B defined by [46] qubit [46]. In the photonic crystal optomechanical setup, − the phase-dependent phonon-hopping coupling has been 2 2 B± = (G1(2)δb1 G2(1)δb2)/ G + G . (3) suggested by using two assistant cavity fields [51]. In ± 1 2 † † q † addition, we mention that the two mechanical modes Then HRWA = ∆δa δa + ω+B+B+ + ω−B−B− + could be either bare mechanical modes in individual † † 2 2 G+(δaB+ + B+δa ) with G+ = G + G . Here the mechanical resonators or supermodes of coupled me- 1 2 dark mode B− decouples from the cavity mode and the chanical resonators [68, 69]. For the latter case, the ground-state cooling of the two resonatorsp is unaccessible. phase-dependent phonon-exchange coupling should be implemented between these supermodes accordingly. Ground-state cooling by breaking the dark mode.—To † † By expressing the operators o a, bl=1,2, a , b analyze the action of the phonon-exchange interaction, we ∈{ l=1,2} with their steady-state average values and fluctuations o = introduce two bosonic modes B˜ = fδb eiθhδb + 1 − 2 o ss + δo, the system can be linearized in the strong- and B˜ = e−iθhδb + fδb , where the coefficients drivingh i regime, and the linearized Hamiltonian in the − 1 2 are given by f = ω˜ ω / (˜ω ω )2 + η2 and rotating-wave approximation (RWA) reads | − − 1| − − 1 h = ηf/(˜ω− ω1), with the resonance frequencies † † † † 1 − p 2 2 HRWA =∆δa δa + [ωlδbl δbl + Gl(δaδbl + δblδa )] ω˜± = 2 (ω1 + ω2 (ω1 ω2) + 4η ) and the ˜± − −iθ lX=1,2 coupling strengths G+ p= fG1 e hG2 and iθ † −iθ † ˜ iθ − +η(e δb1δb2 + e δb2δb1), (2) G− = e hG1 + fG2. The linearized optomechanical Hamiltonian becomes where ∆ is the normalized driving detuning and Gl=1,2 = † ˜† ˜ ˜† ˜ ˜∗ ˜† glα are the linearized optomechanical-coupling strengths. HRWA =∆δa δa +ω ˜+B+B+ +ω ˜−B−B− + (G+δaB+ ∗ The displacement α a ss = iΩ /(κ + i∆) is +G˜ B˜ δa†) + (G˜∗ δaB˜† + G˜ B˜ δa†). (4) assumed to be real by choosing≡ h i a proper− driving amplitude + + − − − − Ω, where κ is the decay rate of the cavity field. When In the degenerate-resonator (ω1 = ω2 = ωm) and ω1 = ω2 and η = 0, there exists a bright mode B+ and a symmetric-coupling (G1 = G2 = G) cases, the coupling 3

˜ −iθ ˜ 1 1 strengths become G+ = √2G(1 + e )/2 and G− = (a) L (b) L L √2G(1 eiθ)/2. When θ = nπ for an integer n, − the cavity field is decoupled from one of the two hybrid 0 0 mechanical modes B˜− (for even n) and B˜+ (for odd n). However, in the general case θ = nπ, the dark-mode L 6 -1 -1 effect is broken [46], and then the simultaneous ground- rates Phonon-scattering 1 3 4 0.5 1.5 state cooling becomes accessible under proper parameter Π conditions. We emphasize that the dark-mode-breaking mechanism is universal and it can be proved by analyzing FIG. 3. (Color online) The relative resonant-phonon-scattering the eigenstates of a 3 3 matrix, which is used to describe rate Λ (blue solid curves) and Λ (red dashed curves) × b2b1 b1b2 either a three-mode system or a three-level system [46]. versus (a) the ratio Π of the optomechanical cooperativities when To study the cooling performance of the two mechanical θ = π/2 and (b) the phase θ when Π = 1. Here ∆ = ωm and resonators, we calculate the final average phonon numbers ω1 = ω2 = ωm. Other parameters used are the same as those in f f Fig. 2. n1 and n2 by solving the steady-state covariance matrix governed by the Lyapunov equation [46]. Figure 2(a) f f shows the phonon numbers n1 and n2 as functions of the θ = nπ, the two mechanical resonators cannot be cooled driving detuning ∆ when the system works in both the to their ground states, which correspondsto the dark-mode- dark-mode-unbreaking (η = 0) and -breaking (η/ωm = unbreaking case, as shown in Figs. 2(c-e). 0.05 and θ = π/2) regimes. The results indicate that Nonreciprocal phonon transfer.—To explain the asym- ground-state cooling of the two mechanical resonators is metrical cooling phenomenon in Fig. 2(e), we introduce a unfeasible when the system possesses the dark mode [the relative resonant-phonon-scattering rate Λ = (T upper solid curves in Fig. 2(a)]. When the dark mode vw vw − T )/(T )max corresponding to the transfer of a phonon is broken by adding the phonon-exchange coupling [the wv vw with frequency ωm from modes w to v, where Tvw dashed curves in Fig. 2(a)], the emergence of the valley denotes the transmittance from modes w to v [v, w corresponds to ground-state cooling (nf 1). The ∈ 1,2 ≪ b1, b2 ]. The relative resonant-phonon-scattering rates phonon-exchange coupling provides the physical origin can{ be} expressed as [46] for breaking the dark mode and builds the channel to transfer the excitation energy between the two mechanical −1 2 resonators. The optimal driving detuning is located at ∆= 4√Πsin θ 4Π cos θ Λb b = 1+ , (5) 2 1 2  2  ωm, which is consistent with a typical resolved-sideband (1 + √Π) C1+C2+1 C C + Π cooling [29–31, 35, 36], because the phonons exactly  1 2  compensate the energy mismatch between the scattered     and Λb1b2 = Λb2b1 , where Π= 3/( 1 2) with l=1,2 = photons and the driving light. 2 − 2 C C C C Gl /γlκ and 3 = η /γ1γ2 being the cooperativities When the phonon-exchange coupling is absent, though associated withC the optomechanical couplings and the the dark mode exists theoretically only in the degenerate- phonon-exchange coupling (γl=1,2 denoting the decay rate resonator case (i.e., ω1 = ω2), the dark-mode effect of the lth resonator), respectively. The dependence of the actually works for a wider detuning range in the near- relative resonant-phonon-scattering rates Λvw on the ratio degenerate-resonator case [as marked by the shadow area Π of the optomechanical cooperativities and the phase θ is in Fig. 2(b)] [46]. The width of the shadow area can shown in Fig. 3. In panel (a), we find that in the region be characterized by the effective mechanical linewidth Γl 0 < Π < 1 (Π > 1), Λb2b1 increases (decreases) with (∆ω = ω2 ω1 Γl). The cooling of the individual increasing , and the optimal nonreciprocity ( | − | ≤ Π Λb2b1 = mechanical resonators is suppressed in this region, i.e., the 1) emerges at Π = 1, which indicates directional flow individual mechanical resonators have significant spectral of phonons between the two mechanical resonators. As overlap and become effectively degenerate. When the shown in Fig. 3(b), when 0 <θ<π, Λb2b1 > 0, i.e., phonon-exchangecoupling is applied, the dark-mode effect T > T , the phonon transmission from mechanical is broken and the ground-state cooling for the degenerate b2b1 b1b2 mode b1 to b2 is enhanced, while the transmission in the and near-degenerate resonators becomes feasible [the backward direction is suppressed (see blue solid curves); dashed curves in Fig. 2(b)]. In the range π<θ< 2π, it exhibits Λb b > 0, The dependence of the final average phonon numbers 1 2 i.e., Tb b > Tb b (see red dashed curves). Meanwhile, f f 1 2 2 1 n1 and n2 on the phonon-exchange parameters η and θ is the phonon transmission satisfies the Lorentz reciprocal displayed in Figs. 2(c) and 2(d). The ground-state cooling theorem [Λb2b1 = Λb1b2 = 0, i.e., Tb1b2 = Tb2b1 ] at of the two mechanical resonators is achievable in the region θ = nπ. Moreover, the transmittance is optimal for the 0 <θ<π (π<θ< 2π) for a wide range of η, and the process from b1 (b2) to b2 (b1) and is zero for the opposite cooling performance of the first (second) resonator is better process when θ = π/2 (θ = 3π/2). We see from Eq. (5) f f f f than the other one n1 < n2 (n1 > n2 ). In particular, at that, when Π = 1 and θ = π/2, an excellent nonreciprocal 4

3 3 (a) (b) (solid lines) and the cooling limits (symbols) given by 2 Exact 2 Eq. (6) as functions of the phonon-exchange parameters η Approximate and θ. Figure 4 shows asymmetrical ground-state cooling and excellent agreement between numerical and analytical results. Phonon numbers The first term in Eq. (6) is caused by the thermal 0.05 bath and the effective optical bath connected by the lth 3 3 mechanical mode, while the phonon extraction by the (c) = /2 (d) =3 /2 2 2 phonon-exchange channel is described by the last term. Physically, the nonreciprocity of the phonon transfer is determined by the phonon-exchangerate χl which depends on the phase θ. For the case: n¯1 n¯2 and γ1 γ2, we

Phonon numbers ≈ ≈ have nχ nχ = nχ and thus (√χ1nχ √χ2nχ ) 1 ≈ 2 1 − 2 ≈ (√χ1 √χ2)nχ [see Eq. (6)]. In the range 0 <θ<π 0.2 0.4 0.6 0.2 0.4 0.6 − κ /wm κ /wm (π<θ< 2π), we obtain √χ1 < √χ2 (√χ1 > √χ2). This means that the phonon-transfer efficiency from b1 FIG. 4. (Color online) The exact and approximate final average (b2) to b2 (b1) is larger than that for the opposite case, f f f f f f phonon numbers n1 (blue) and n2 (red) versus (a) the phonon- i.e., n1 < n2 (n1 > n2 ) [see Fig. 4(b)]. When θ = exchange coupling strength η when θ = π/2 and κ/ωm = 0.2, π/2 (3π/2) and √ 1 2 = √ 3, the unidirectional flow (b) the phase θ when η/ωm = 0.05 and κ/ωm = 0.2, and (c,d) of the phonons betweenC C the twoC mechanical resonators the cavity-field decay rate κ when η/ωm = 0.05 for (c) θ = π/2 is obtained [χ1 0 (χ2 0)]. For θ = nπ, the and (d) θ = 3π/2. The solid curves and the symbols correspond phonon transfer between≈ the≈ two mechanical resonators is to the exact (numerical) and approximate (analytical) results, reciprocal (√χ1 = √χ2), due to the emergenceof the dark respectively. Here ∆/ωm = 1 and G1/ωm = G2/ωm = 0.05. Other parameters used are the same as those in Fig. 2. mode. In the absence of the phonon-transfer interaction (η = 0), the ground-state cooling is unfeasible due to the invalid effective cooling channel (Γ + χ γ ) l + → l phonon transfer (Λb2b1 = 1) is realized. [see Fig. 4(a)]. In the absence of the optomechanical f Cooling limits.—The cooling limits can be analytically cooling channels (G1,2 = 0), Eq. (6) becomes nl=1,2 obtained in the large cavity-field-decay regime, in which l−1 ≈ n¯l + ( 1) (nχ1 nχ2 )/2, which indicates quantum the cavity field is eliminated adiabatically such that the thermalization− in the− coupled mechanical system. three-mode optomechanical system is reduced to a two- Cooling N mechanical resonators.—Our proposal can ˜ mode system described by the Hamiltonian Heff = be extended to the cooling of a net-coupled system: a 2 † † † l=1(Ωl iΓl)bl bl +iξ1b1b2 +iξ2b2b1 [Fig. 1(b)], where cavity mode coupled to N 3 mechanical modes via the − ≥ N † Γl = γl + γl,opt and Ωl = ωl ωl,opt are, respectively, optomechanical couplings H = g a†a(b + b ), P − opc j=1 j j j the effective decay rate and resonance frequency for the lth and the nearest-neighboring mechanical modes are coupled mechanical resonator, with the optical induced decay rates through the phase-dependent phonon-exchangeP couplings 2 N−1 † γl,opt = Gl /κ and mechanical frequency shifts ωl,opt = iθj 2 Hpec = j=1 ηj (e bj bj+1 + H.c.). We find that the Gl /2ωl. In addition, iξl is the effective phonon-exchange function of these phases in the optomechanical interactions coupling strength between the two mechanical modes j−1 is determinedP by the term θ [46] and hence, for with ξ = [G G /κ + i(ηe±iθ G G /2ω )]. ν=1 ν 1(2) 1 2 1 2 2(1) convenience, we assume θ = π/2 and θ = 0 for The mechanical− mode b is contacted− to an effective 1 j l j = 2 - (N 1) in our simulations.P In the dark-mode- optomechanical cooling bath (γ and n ) and a heat l,opt opt unbreaking case− (η = 0), the ground-state cooling of the bath (γ and n¯ ). Considering the parameter relations j l l mechanical resonators is unfeasible, with the final average ω κ G γ γ γ , the 1,2 1,2 1,opt 2,opt 1,2 phonon numbers n¯(N 1)/N in the case of n¯ =n ¯ [46]. final average≫ ≫ phonon occupations≫ { can≈ be obtained} ≫ as [46] j When the dark modes− are broken, simultaneous ground- l−1 f f γln¯l + γl,optnopt ( 1) √χl state cooling can be realized in this system (nj < 1). nl=1,2 + − ≈ Γl + χ+ Γl + χ− Conclusions.—We proposed a dark-mode-breaking (√χ n √χ n ), (6) method to realize simultaneous ground-state cooling of × 1 χ1 − 2 χ2 multiple mechanical modes coupled to a common cavity 2 2 where nopt = 4κ /(ω1 + ω2 + 2∆) , nχ1(2) = mode by constructing a loop-coupled optomechanical 2(γ2(1)n¯2(1) + γ2(1),optnopt)/(Γ1 +Γ2 + 2χ+), and χ± = system with a phase drop. We found an asymmetric cooling 2 √χ1χ2 Re[ξ1ξ2/(Γ1+Γ2)], with χl=1,2 = ξl /(Γ1+ phenomenon and expounded it using the nonreciprocal ∓ − | | Γ2) being the effective phonon-transfer rate from b2 (b1) to phonon exchange mechanism. The present physical lim b1 (b2). The cooling limits (nl ) are obtained at ∆ = ωl. mechanism is universal and hence it will motivate the In Fig. 4, we plot the exact final average phonon numbers manipulation of various dark-state related physical effects. 5

Acknowledgments.—D.-G.L. thanks Yue-Hui Zhou and Fazio, Measures of Quantum Synchronization in Continuous Dr. Wei Qin for valuable discussions. J.-Q.L. is Variable Systems, Phys. Rev. Lett. 111, 103605 (2013). supported in part by National Natural Science Founda- [11] M. H. Matheny, M. Grau, L. G. Villanueva, R. B. Karabalin, tion of China (Grants No. 11822501, No. 11774087, M. C. Cross, and M. L. Roukes, Phase Synchronization of and No. 11935006), Natural Science Foundation of Two Anharmonic Nanomechanical Oscillators, Phys. Rev. Hunan Province, China (Grant No. 2017JJ1021), and Lett. 112, 014101 (2014). [12] M. Zhang, S. Shah, J. Cardenas, and M. Lipson, Syn- Hunan Science and Technology Plan Project (Grant chronization and Phase Noise Reduction in Micromechanical No. 2017XK2018). D.-G.L. is supported in part by Hunan Oscillator Arrays Coupled through Light, Phys. Rev. Lett. Provincial Postgraduate Research and Innovation project 115, 163902 (2015). (Grant No. CX2018B290). J.-F.H. is supported in part by [13] R. Riedinger, A. Wallucks, I. Marinkovi´c, C. L¨oschnauer, the National Natural Science Foundation of China (Grant M. Aspelmeyer, S. Hong, and S. Gr¨oblacher, Remote quan- No. 11505055) and Scientific Research Fund of Hunan tum entanglement between two micromechanical oscillators, Provincial Education Department (Grant No. 18A007). B.- Nature (London) 556, 473 (2018). P.H. is supported in part by NNSFC (Grant No. 11974009). [14] C. F. Ockeloen-Korppi, E. Damsk¨agg, J.-M. Pirkkalainen, W.L. and D.V. are supported by the European Union M. Asjad, A. A. Clerk, F. Massel, M. J. Woolley, and M. Horizon 2020 Programme for Research and Innovation A. Sillanp¨a¨a, Stabilized entanglement of massive mechanical oscillators, Nature (London) 556, 478 (2018). through the Project No. 732894 (FET Proactive HOT) and [15] O. D. Stefano, A. Settineri, V. Macr`ı, A. Ridolfo, R. the Project QuaSeRT funded by the QuantERA ERA-NET Stassi, A. F. Kockum, S. Savasta, and F. Nori, Interaction of Cofund in Quantum Technologies. F.N. is supported in part Mechanical Oscillators Mediated by the Exchange of Virtual by: NTT Research, Army Research Office (ARO) (Grant Photon Pairs, Phys. Rev. Lett. 122, 030402 (2019). No. W911NF-18-1-0358), Japan Science and Technology [16] G. Heinrich, M. Ludwig, J. Qian, B. Kubala, and F. Agency (JST) (via the CREST Grant No. JPMJCR1676), Marquardt, Collective Dynamics in Optomechanical Arrays, Japan Society for the Promotion of Science (JSPS) (via the Phys. Rev. Lett. 107, 043603 (2011). KAKENHI Grant No. JP20H00134, and the JSPS-RFBR [17] A. Xuereb, C. Genes, and A. Dantan, Strong Coupling Grant No. JPJSBP120194828), and the Foundational and Long-Range Collective Interactions in Optomechanical Questions Institute Fund (FQXi) (Grant No. FQXi-IAF19- Arrays, Phys. Rev. Lett. 109, 223601 (2012). 06), a donor advised fund of the Silicon Valley Community [18] M. Ludwig and F. Marquardt, Quantum Many-Body Dynamics in Optomechanical Arrays, Phys. Rev. Lett. 111, Foundation. 073603 (2013). [19] A. Xuereb, C. Genes, G. Pupillo, M. Paternostro, and A. Dantan, Reconfigurable Long-Range Phonon Dynamics in Optomechanical Arrays, Phys. Rev. Lett. 112, 133604

∗ (2014). [email protected] [20] A. Xuereb, A. Imparato, and A. Dantan, Heat transport in [1] T. J. Kippenberg and K. J. Vahala, Cavity optomechanics: harmonic oscillator systems with thermal baths: Application Back-action at the mesoscale, Science 321, 1172 (2008). to optomechanical arrays, New J. Phys. 17, 055013 (2015). [2] P. Meystre, A short walk through quantum optomechanics, [21] O. Cernot´ık,ˇ S. Mahmoodian, and K. Hammerer, Spatially Ann. Phys. () 525, 215 (2013). Adiabatic Frequency Conversion in Optoelectromechanical [3] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Cavity Arrays, Phys. Rev. Lett. 121, 110506 (2018). optomechanics, Rev. Mod. Phys. 86, 1391 (2014). [22] H. Xu, D. Mason, L. Jiang, and J. G. E. Harris, Topological [4] J.-Q. Liao and L. Tian, Macroscopic Quantum Superposition energy transfer in an optomechanical system with exceptional in Cavity Optomechanics, Phys. Rev. Lett. 116, 163602 points, Nature (London) 537, 80 (2016). (2016). [23] F. Massel, T. T. Heikkil¨a, J.-M. Pirkkalainen, S. U. [5] S. Mancini, V. Giovannetti, D. Vitali, and P. Tombesi, Cho, H. Saloniemi, P. J. Hakonen, and M. A. Sillanp¨a¨a, Entangling Macroscopic Oscillators Exploiting Radiation Microwave amplification with nanomechanical resonators, Pressure, Phys. Rev. Lett. 88, 120401 (2002). Nature (London) 480, 351 (2011). [6] L. Tian and P. Zoller, Coupled Ion-Nanomechanical Systems, [24] P. Huang, P. Wang, J. Zhou, Z. Wang, C. Ju, Z. Wang, Phys. Rev. Lett. 93, 266403 (2004). Y. Shen, C. Duan, and J. Du, Demonstration of Motion [7] M. J. Hartmann and M. B. Plenio, Steady State Entanglement Transduction Based on Parametrically Coupled Mechanical in the Mechanical Vibrations of Two Dielectric Membranes, Resonators, Phys. Rev. Lett. 110, 227202 (2013). Phys. Rev. Lett. 101, 200503 (2008). [25] P. Huang, L. Zhang, J. Zhou, T. Tian, P. Yin, C. Duan, [8] F. Massel, S. U. Cho, J.-M. Pirkkalainen, P. J. Hakonen, and J. Du, Nonreciprocal Radio Frequency Transduction in T. T. Heikkil¨a, and M. A. Sillanp¨a¨a, Multimode circuit a Parametric Mechanical Artificial Lattice, Phys. Rev. Lett. optomechanics near the quantum limit, Nat. Commun. 3, 987 117, 017701 (2016). (2012). [26] P. Rabl, S. J. Kolkowitz, F. H. L. Koppens, J. G. E. Harris, [9] Z. L. Xiang, S. Ashhab, J. Q. You, and F. Nori, “Hybrid P. Zoller, and M. D. Lukin, A quantum spin transducer based quantum circuits: Superconducting circuits interacting with on nanoelectromechanical resonator arrays, Nat. Phys. 6, 602 other quantum systems”, Rev. Mod. Phys. 85, 623 (2013). (2010). [10] A. Mari, A. Farace, N. Didier, V. Giovannetti, and R. [27] S. C. Masmanidis, R. B. Karabalin, I. D. Vlaminck, G. 6

Borghs, M. R. Freeman, and M. L. Roukes, Multifunctional Lett. 123, 203605 (2019). Nanomechanical Systems via Tunably Coupled Piezoelectric [44] C. F. Ockeloen-Korppi, M. F. Gely, E. Damsk¨agg, M. Actuation, Science 317, 780 (2007). Jenkins, G. A. Steele, and M. A. Sillanpa¨a¨, Sideband [28] I. Mahboob and H. Yamaguchi, Bit storage and bit cooling of nearly degenerate micromechanical oscillators in a flip operations in an electromechanical oscillator, Nat. multimode optomechanical system, Phys. Rev. A 99, 023826 Nanotechnol. 3, 275 (2008). (2019). [29] I. Wilson-Rae, N. Nooshi, W. Zwerger, and T. J. Kip- [45] A. B. Shkarin, N. E. Flowers-Jacobs, S. W. Hoch, A. D. penberg, Theory of Ground State Cooling of a Mechanical Kashkanova, C. Deutsch, J. Reichel, and J. G. E. Harris, Oscillator Using Dynamical Backaction, Phys. Rev. Lett. 99, Optically Mediated Hybridization between Two Mechanical 093901 (2007). Modes, Phys. Rev. Lett. 112, 013602 (2014). [30] F. Marquardt, J. P. Chen, A. A. Clerk, and S. M. Girvin, [46] See Supplemental Material, which includes Refs. [47– Quantum Theory of Cavity-Assisted Sideband Cooling of 50], for analyzing and breaking the dark-mode effect, Mechanical Motion, Phys. Rev. Lett. 99, 093902 (2007). ground-state cooling of the two mechanical resonators, [31] C. Genes, D. Vitali, P. Tombesi, S. Gigan, and M. nonreciprocal phonon transfer, derivation of the cooling Aspelmeyer, Ground-state cooling of a micromechanical limits, ground-state cooling of N mechanical resonators, oscillator: Comparing cold damping and cavity-assisted discussing the parameter condition of the rotating-wave cooling schemes, Phys. Rev. A 77, 033804 (2008). Erratum: approximation; simultaneous cooling of the mechanical Ground-state cooling of a micromechanical oscillator: Com- supermodes; breaking the dark state in a Lambda-type three- paring cold damping and cavity-assisted cooling schemes level system, and a possible physical implementation of the [Phys. Rev. A 77, 033804 (2008)]. present model and detailed derivation of a phase-dependent [32] K. Xia and J. Evers, Ground State Cooling of a phonon-hopping interaction. Nanomechanical Resonator in the Nonresolved Regime via [47] Y.-x. Liu, J. Q. You, L. F. Wei, C. P. Sun, and F. Nori, Quantum Interference, Phys. Rev. Lett. 103, 227203 (2009). Optical Selection Rules and Phase-Dependent Adiabatic [33] L. Tian, Ground state cooling of a nanomechanical State Control in a Superconducting Quantum Circuit, Phys. resonator via parametric linear coupling, Phys. Rev. B 79, Rev. Lett. 95, 087001 (2005). 193407 (2009). [48] M. Nakahara and T. Ohmi, Quantum computing: From [34] A. Xuereb, T. Freegarde, P. Horak, and P. Domokos, linear algebra to physical realizations (CRC Press, Boca Optomechanical Cooling with Generalized Interferometers, Raton, 2008). Phys. Rev. Lett. 105, 013602 (2010). [49] H. Fr¨ohlich, Theory of the Superconducting State. I. The [35] J. Chan, T. P. Alegre, A. H. Safavi-Naeini, J. T. Hill, A. Ground State at the Absolute Zero of Temperature, Phys. Rev. Krause, S. Groeblacher, M. Aspelmeyer, and O. Painter, 79, 845 (1950). Laser cooling of a nanomechanical oscillator into its quantum [50] S. Nakajima, Perturbation theory in statistical mechanics, ground state, Nature (London) 478, 89 (2011). Adv. Phys. 4, 363 (1953). [36] J. D. Teufel, T. Donner, D. Li, J. W. Harlow, M. S. Allman, [51] K. Fang, J. Luo, A. Metelmann, M. H. Matheny, F. K. Cicak, A. J. Sirois, J. D. Whittaker, K. W. Lehnert, and R. Marquardt, A. A. Clerk, and O. Painter, Generalized W. Simmonds, Sideband cooling of micromechanical motion non-reciprocity in an optomechanical circuit via synthetic to the quantum ground state, Nature (London) 475, 359 magnetism and reservoir engineering, Nat. Phys. 13, 465 (2011). (2017). [37] Y.-C. Liu, Y.-F. Xiao, X. Luan, and C. W. Wong, Dynamic [52] N. R. Bernier, L. D. T´oth, A. Koottandavida, M. A. Dissipative Cooling of a Mechanical Resonator in Strong Ioannou, D. Malz, A. Nunnenkamp, A. K. Feofanov, and Coupling Optomechanics, Phys. Rev. Lett. 110, 153606 T. J. Kippenberg, Nonreciprocal reconfigurable microwave (2013). optomechanical circuit, Nat. Commun. 8, 604 (2017). [38] J. B. Clark, F. Lecocq, R. W. Simmonds, J. Aumentado, [53] D. Malz, L. D. T´oth, N. R. Bernier, A. K. Feofanov, and J. D. Teufel, Sideband cooling beyond the quantum T. J. Kippenberg, and A. Nunnenkamp, Quantum-Limited backaction limit with squeezed light, Nature (London) 541, Directional Amplifiers with Optomechanics, Phys. Rev. Lett. 191 (2017). 120, 023601 (2018). [39] X. Xu, T. Purdy, and J. M. Taylor, Cooling a Harmonic [54] J. P. Mathew, J. d. Pino, and E. Verhagen, Synthetic gauge Oscillator by Optomechanical Modification of Its Bath, Phys. fields for phonon transport in a nano-optomechanical system, Rev. Lett. 118, 223602 (2017). Nat. Nanotechnol. 15, 198 (2020). [40] M. Rossi, N. Kralj, S. Zippilli, R. Natali, A. Borrielli, G. [55] K. Fang, Z. Yu, and S. Fan, Realizing effective magnetic Pandraud, E. Serra, G. D. Giuseppe, and D. Vitali, Enhancing field for photons by controlling the phase of dynamic Sideband Cooling by Feedback-Controlled Light, Phys. Rev. modulation, Nat. Photonics 6, 782 (2012). Lett. 119, 123603 (2017). [56] K. Fang, Z. Yu, and S. Fan, Photonic Aharonov-Bohm [41] O. Cernot´ık,ˇ A. Dantan, and C. Genes, Cavity Quan- Effect Based on Dynamic Modulation, Phys. Rev. Lett. 108, tum Electrodynamics with Frequency-Dependent Reflectors, 153901 (2012). Phys. Rev. Lett. 122, 243601 (2019). [57] M. Hafezi and P. Rabl, Optomechanically induced non- [42] C. Genes, D. Vitali, and P. Tombesi, Simultaneous cooling reciprocity in microring resonators, Opt. Express 20, 7672 and entanglement of mechanical modes of amicromirror in (2012). an optical cavity, New J. Phys. 10, 095009 (2008). [58] A. Metelmann and A. A. Clerk, Nonreciprocal Photon [43] C. Sommer and C. Genes, Partial Optomechanical Refrig- Transmission and Amplification via Reservoir Engineering, eration via Multimode Cold-Damping Feedback, Phys. Rev. Phys. Rev. X 5, 021025 (2015). 1

[59] Z. Shen, Y.-L. Zhang, Y. Chen, C.-L. Zou, Y.-F. Xiao, X.-B. Thermal management and non-reciprocal control of phonon Zou, F.-W. Sun, G.-C. Guo, and C.-H. Dong, Experimental flow via optomechanics, Nat. Commun. 9, 1207 (2018). realization of optomechanically induced non-reciprocity, [65] H. Xu, L. Jiang, A. A. Clerk, and J. G. E. Harris, Nat. Photonics 10, 657 (2016). Nonreciprocal control and cooling of phonon modes in an [60] G. A. Peterson, F. Lecocq, K. Cicak, R. W. Simmonds, optomechanical system, Nature (London) 568, 65 (2019). J. Aumentado, and J. D. Teufel, Demonstration of Efficient [66] M. Partanen, K. Y. Tan, J. Govenius, R. E. Lake, M. K. Nonreciprocity in a Microwave Optomechanical Circuit, Mˆakel¨a, T. Tanttu, and M. M¨ott¨onen, Quantum-limited heat Phys. Rev. X 7, 031001 (2017). conduction over macroscopic distances, Nat. Phys. 12, 460 [61] S. Kim, X. Xu, J. M. Taylor, and G. Bahl, Dynamically (2016). induced robust phonon transport and chiral cooling in an [67] S. Barzanjeh, M. Aquilina, and A. Xuereb, Manipulating optomechanical system, Nat. Commun. 8, 205 (2017). the Flow of Thermal Noise in Quantum Devices, Phys. Rev. [62] P. Lodahl, S. Mahmoodian, S. Stobbe, A. Rauschenbeutel, Lett. 120, 060601 (2018). P. Schneeweiss, J. Volz, H. Pichler, and P. Zoller, Chiral [68] D. Ramos, I. W. Frank, P. B. Deotare, I. Bulu, and quantum optics, Nature (London) 541, 473 (2017). M. Lonˇcar, Non-linear mixing in coupled photonic crystal [63] Z. Shen, Y.-L. Zhang, Y. Chen, F.-W. Sun, X.-B. Zou, nanobeam cavities due to cross-coupling opto-mechanical G.-C. Guo, C.-L. Zou, and C.-H. Dong, Reconfigurable mechanisms, Appl. Phys. Lett. 105, 181121 (2014). optomechanical circulator and directional amplifier, Nat. [69] S. Barzanjeh and D. Vitali, Phonon Josephson junction with Commun. 9, 1797 (2018). nanomechanical resonators, Phys. Rev. A 93, 033846 (2016). [64] A. Seif, W. DeGottardi, K Esfarjani, and M. Hafezi, 2

Supplementary Material for “Nonreciprocal Ground-State Cooling of Multiple Mechanical Resonators”

Deng-Gao Lai1,2, Jin-Feng Huang1, Xian-Li Yin1, Bang-Pin Hou3, Wenlin Li4, David Vitali4,5,6, Franco Nori2,7, and ∗ Jie-Qiao Liao1,

1Key Laboratory of Low-Dimensional Quantum Structures and Quantum Control of Ministry of Education, Key Laboratory for Matter Microstructure and Function of Hunan Province, Department of Physics and Synergetic Innovation Center for Quantum Effects and Applications, Hunan Normal University, Changsha 410081, China 2Theoretical Quantum Physics Laboratory, RIKEN, Saitama 351-0198, Japan 3College of Physics and Electronic Engineering, Institute of Solid State Physics, Sichuan Normal University, Chengdu 610068, China 4School of Science and Technology, Physics Division, University of Camerino, I-62032 Camerino (MC), Italy 5INFN, Sezione di Perugia, I-06123 Perugia, Italy 6CNR-INO, L.-argo Enrico Fermi 6, I-50125 Firenze, Italy 7Physics Department, The University of Michigan, Ann Arbor, Michigan 48109-1040, USA

This document consists of ten parts: (I) The dark-mode effect and its breaking in a two-mechanical-resonator optomechanical system; (II) Ground-state cooling of the two mechanical resonators; (III) Phonon scattering probability and nonreciprocal phonon transfer; (IV) Derivation of the cooling limits of the two mechanical resonators; (V) Analyzing the dark-mode effect and breaking the dark-mode effect in a multi-mechanical-resonator optomechanical system; (VI) Ground-state cooling of the multiple mechanical resonators; (VII) Discussions on the justification of performing the rotating-wave approximation (RWA); (VIII) Simultaneous cooling of the mechanical supermodes; (IX) Physical mechanism for breaking the dark-state effect in a Lambda-type three-level system; (X) A possible experimental realization and derivation of a phase-dependent phonon-hopping interaction between two mechanical resonators.

S1. THE DARK-MODE EFFECT AND ITS BREAKING IN A TWO-MECHANICAL-RESONATOR OPTOMECHANICAL SYSTEM

In this section, we analyze the dark-mode effect in a two-mechanical-resonator optomechanical system, which is composed of one cavity-field mode and two mechanical resonators. Note that here we only consider one mechanical mode in each mechanical resonator. We also show that the dark-mode effect can be broken by introducing a phase-dependent phonon-exchangeinteraction between the two mechanical resonators. In a rotating frame defined by the transform operator exp( iω ta†a), the total Hamiltonian of the system reads (~ = 1) − L † † † † † † † ∗ † iθ † −iθ † HI =∆ca a + ω1b1b1 + ω2b2b2 + g1a a(b1 + b1)+ g2a a(b2 + b2)+(Ωa +Ω a )+ η(e b1b2 + e b2b1)(), where ∆c = ωc ωL is the detuning of the cavity-field resonance frequency ωc with respect to the cavity-field driving − † † frequency ωL. The operators a (a ) and bl=1,2 (bl ) are, respectively, the annihilation (creation) operators of the cavity-field mode and the lth mechanical resonator, with the corresponding resonance frequencies ωc and ωl. The g1 and g2 terms in Hamiltonian (S1) describe the optomechanical coupling between the cavity mode and the lth mechanical resonator, with gl=1,2 being the single-photon optomechanical-coupling strength. The Ω term denotes the cavity-field driving with the driving amplitude Ω. To control the energy exchange between the two mechanical resonators, we introduce a phase- dependent phonon-exchange interaction between the two mechanical resonators, with the coupling strength η and the phase θ. According to Hamiltonian (S1), the Langevin equations for the annihilation operators of the optical and mechanical modes can be obtained by phenologically adding the dissipation and noise terms into the Heisenberg equations of motion as a˙ = κ + i[∆ + g (b + b†)+ g (b + b†)] a iΩ∗ + √2κa , (S2a) − { c 1 1 1 2 2 2 } − in b˙ = (γ + iω )b ig a†a iηeiθb + 2γ b , (S2b) 1 − 1 1 1 − 1 − 2 1 1,in ˙ † −iθ b2 = (γ2 + iω2)b2 ig2a a iηe b1 +p 2γ2b2,in, (S2c) − − − where κ and γl=1,2 are the decay rates of the cavity-field mode andp the lth mechanical resonator, respectively. The † † operators ain and bl=1,2,in (ain and bl,in) are the noise operators associated with the cavity-field modeand the lth mechanical 3 resonator, respectively. These noise operators have zero mean values and the following correlation functions, a (t)a† (t′) =δ(t t′), (S3a) h in in i − a† (t)a (t′) =0, (S3b) h in in i b (t)b† (t′) =(¯n + 1)δ(t t′), (S3c) h l,in l,in i l − b† (t)b (t′) =¯n δ(t t′), (S3d) h l,in l,in i l − where n¯l=1,2 is the average thermal-phonon occupation number associated with the heat bath of the lth mechanical resonator. In this paper we consider a vacuum bath for the cavity field and a heat bath (with n¯l=1,2) for each mechanical resonator. The vacuum bath of the cavity field provides the cooling reservoir to absorb the thermal excitations extracted from the two mechanical resonators. To cool the mechanical resonators, we consider the strong-driving regime of the cavity such that the average photon number in the cavity is sufficiently large and then the linearization procedure can be used to simplify the physical model. To this end, we expand the quantum fluctuations of the system around their steady-state values and express the operators in Eq. () as a summation of their steady-state mean values and quantum fluctuations, namely o = o ss + δo for operators † † h i o = a, a , bl=1,2, and bl=1,2. By separating the classical motion and quantum fluctuations, the linearized equations of motion for quantum fluctuations can be written as δa˙ = (κ + i∆)δa iG (δb + δb† ) iG (δb + δb† )+ √2κa , (S4a) − − 1 1 1 − 2 2 2 in δb˙ = iG∗δa (γ + iω )δb iηeiθδb iG δa† + 2γ b , (S4b) 1 − 1 − 1 1 1 − 2 − 1 1 1,in ˙ ∗ −iθ † δb2 = iG δa iηe δb1 (γ2 + iω2)δb2 iG2δa +p 2γ2b2,in, (S4c) − 2 − − − where ∆=∆c + 2(g1Re[β1]+ g2Re[β2]) is the normalized driving detuning of thep cavity field with Re[βl] extracting the real part of βl, and Gl=1,2 = glα is the strength of the linearized optomechanical coupling between the cavity field and the lth mechanical resonator. Here, the steady-state solutions of the classical motion (namely the steady-state average values of the operators of the system) can be obtained as iΩ∗ α a = − , (S5a) ≡h iss κ + i∆ 2 iθ i g1 α + ηe β2 β1 b1 ss = − | | , (S5b) ≡h i γ1 + iω1  2 −iθ i g2 α + ηe β1 β2 b2 ss = − | | . (S5c) ≡h i γ2 + iω2  For simplicity, in the following discussions we consider the case where α is real, which is accessible by choosing a proper driving amplitude Ω. Then the linearized optomechanical coupling strengths G1 and G2 are real. A linearized optomechanical Hamiltonian can be inferred according to Eqs. (). For studying quantum cooling of the two mechanical resonators, the beam-splitting-type interactions (i.e., the rotating-wave interaction term) between these bosonic modes are expected to dominate the linearized couplings in this system, and hence we can simplify the Hamiltonian of the system by making the rotating-wave approximation (RWA). The linearized optomechanical Hamiltonian in the RWA takes the following form (discarding the noise terms)

† † † † † † † iθ † −iθ † HRWA =∆δa δa + ω1δb1δb1 + ω2δb2δb2 + G1(δaδb1 + δb1δa )+ G2(δaδb2 + δb2δa )+ η(e δb1δb2 + e δb2δb1), ()

† † where δa (δa ) and δbl=1,2 (δbl ) are the fluctuation operators of the cavity-field mode and the lth mechanical resonator, respectively. To see the dark-mode effect in this two-mechanical-resonator optomechanical system, we first consider the case where the phase-dependent phonon-exchange interaction between the two mechanical resonators is absent, i.e., η = 0. In this case, the coupled two-mechanical-mode system forms two hybrid mechanical modes: a bright mode and a dark mode, which are expressed by the new annihilation operators as 1 B = (G δb + G δb ), (S7a) + 2 2 1 1 2 2 G1 + G2 1 B− =p (G δb G δb ). (S7b) 2 2 2 1 1 2 G1 + G2 − p 4

† † These new operators satisfy the bosonic commutation relations [B+,B+] = 1 and [B−,B−] = 1. In the absence of the phonon-exchange interaction (η = 0), the Hamiltonian in Eq. (S6) can be rewritten with the two hybrid modes as

† † † † † † † Hhyb =∆δa δa + ω+B+B+ + ω−B−B− + ζ(B+B− + B−B+)+ G+(δaB+ + B+δa ), () where we introduce the resonance frequencies ω± and the coupling strengths ζ and G+ 2 2 G1ω1 + G2ω2 ω+ = 2 2 , (S9a) G1 + G2 2 2 G2ω1 + G1ω2 ω− = 2 2 , (S9b) G1 + G2 G1G2(ω1 ω2) ζ = 2 −2 , (S9c) G1 + G2

2 2 G+ = G1 + G2. (S9d) q When ω1 = ω2, the two hybrid modes are decoupled from each other due to ζ = 0, and the mode B− becomes a dark mode in the sense that it is decoupled from both the cavity mode a and the other hybrid mode B+. In order to break the dark-mode effect, we introduce a phase-dependent phonon-exchange interaction (i.e., the η term)

between the two mechanical resonators. By introducing two new bosonic modes B˜+ and B˜− defined by iθ δb1 =fB˜+ + e hB˜−, (S10a) δb = e−iθhB˜ + fB˜ , (S10b) 2 − + − Hamiltonian (S6) becomes † ˜† ˜ ˜† ˜ ˜∗ ˜† ˜ ˜ † ˜∗ ˜† ˜ ˜ † HRWA =∆δa δa +ω ˜+B+B+ +ω ˜−B−B− + (G+δaB+ + G+B+δa ) + (G−δaB− + G−B−δa ), (S11) where we introduce the resonance frequencies ω˜± and the coupling strengths G˜± as 1 ω˜ = (ω + ω (ω ω )2 + 4η2), (S12a) ± 2 1 2 ± 1 − 2 q G˜ =fG e−iθhG , (S12b) + 1 − 2 iθ G˜− =e hG1 + fG2, (S12c) with ω˜ ω f = | − − 1| , (S13a) (˜ω ω )2 + η2 − − 1 ηf h =p . (S13b) ω˜ ω − − 1 In the degenerate-resonator case, namely when the two mechanical resonators have the same resonance frequencies ω1 = ω2 = ωm, the coupling strengths in Eq. (S12) can be simplified as −iθ G˜+ =(G1 + e G2)/√2, (S14a) G˜ =(G eiθG )/√2. (S14b) − 2 − 1 We proceed to analyze the dependence of the dark-mode effect on the coupling strengths G1 and G2. Concretely, we will consider three special cases. (i) In the symmetric-coupling case: G1 = G2 = G, we obtain the relations −iθ G˜+ =G(1 + e )/√2, (S15a) G˜ =G(1 eiθ)/√2. (S15b) − − It can be seen from Eq. (S15) that, when θ = nπ for an integer n, one of the two hybrid mechanical modes (the dark mode) will be decoupled from the cavity-field mode. In this case, the excitation energy stored in the dark mode cannot be extracted through the optomechanical-cooling channel. In general cases of θ = nπ, the dark-mode effect is broken and then ground-state cooling of the two mechanical resonators becomes accessible6 under proper parameter conditions. 5

(ii) In the case θ = nπ for an even number n, Eq. (S14) becomes

G˜+ =(G1 + G2)/√2, (S16a) G˜ =(G G )/√2. (S16b) − 2 − 1 We can see that the dark mode (i.e., the mode B˜− in this case) can be broken when the two optomechanical coupling strengths are different G = G . In this case, our numerical simulation indicates that simultaneous ground-state cooing 1 6 2 of the two mechanical resonators can be realized when G2/G1 1. (iii) In the case θ = nπ for an odd number n, we have ≪ G˜ =(G G )/√2, (S17a) + 1 − 2 G˜− =(G2 + G1)/√2. (S17b)

In this case, the mode B˜+ becomes the dark mode when G1 = G2. The simultaneous ground-state cooing of the two mechanical resonators can be realized when G /G 1, as shown by Fig. S2(d). 2 1 ≪

S2. GROUND-STATE COOLING OF THE TWO MECHANICAL RESONATORS

In this section, we study the cooling performance in this system by evaluating the final average phonon numbers in the two mechanical resonators. To this end, we proceed to rewrite the linearized Langevin equations (S4) as the following compact form ˙u(t)= Au(t)+ N(t), (S18) where the fluctuation operator vector u(t), the noise operator vector N(t), and the coefficient matrix A are defined as

† † † T u(t) = [δa(t), δb1 (t), δb2(t), δa (t), δb1(t), δb2(t)] , (S19)

† † † T N(t) =[√2κain(t), √2γ1b1,in(t), √2γ2b2,in(t), √2κain(t), √2γ1b1,in(t), √2γ2b2,in(t)] , (S20) and

(κ + i∆) iG1 iG2 0 iG1 iG2 − ∗ − − iθ − − iG1 (γ1 + iω1) iηe iG1 0 0  − ∗ − −iθ − −  iG2 iηe (γ2 + iω2) iG2 0 0 A = − − ∗ − ∗ − ∗ ∗ . (S21)  0 iG1 iG2 (κ i∆) iG1 iG2   ∗ − − −iθ   iG1 0 0 iG1 (γ1 iω1) iηe   iG∗ 0 0 iG − iηe−iθ (γ iω )   2 2 − 2 − 2  The formal solution of the linearized Langevin equation (S18) can be written as  t u(t)= M(t)u(0) + M(t s)N(s)ds, (S22) − Z0 where the matrix M(t) is defined by M(t) = exp(At). Based on the solution, we can calculate the steady-state average phonon numbers in the two mechanical resonators by solving the Lyapunov equation. Note that the parameters used in the following calculations satisfy the stability conditions derived from the Routh-Hurwitz criterion. Namely, the real parts of all the eigenvalues of the coefficient matrix A are negative. For studying quantum cooling of the two mechanical resonators, we focus on the final average phonon numbers in the two mechanical resonators by calculating the steady-state value of the covariance matrix V, which is defined by the matrix elements 1 V = [ u ( )u ( ) + u ( )u ( ) ], i, j = 1 6. (S23) ij 2 h i ∞ j ∞ i h j ∞ i ∞ i − In the linearized optomechanical system, the covariance matrix V satisfies the Lyapunov equation AV + VAT = Q, (S24) − where “T ” denotes the matrix transpose operation and the matrix Q is defined by 1 Q = (C + CT ), () 2 6

3 3 (a) (b) 2 2

1.5 1.5 1 1 1.2 1.2 1 0.5 κ 1 0.5 κ 0.8 0.05 0.8 0.05

5 5 (c) (d)

1 1

1.5 1.5 0.1 0.1 1 1 1.2 1.2 1 0.5 κ 1 0.5 κ 0.8 0.05 0.8 0.05

3 3 (e) (f) = Unbreaking 2 Unbreaking 2

Breaking

κ=1.2 Phonon number κ=0.2 Phonon number Breaking

0.8 1 1.2 0.050.5 1 1.5 κ

f f FIG. S1. (Color online) The final average phonon numbers n1 and n2 versus the resonance-frequency ratio ω2/ω1 and the cavity-field decay rate κ scaled by ω1 in both (a,b) the dark-mode-unbreaking case (η/ω1 = 0) and (c,d) the dark-mode-breaking case (η/ω1 = 0.05 f f and θ = π/2). (e) The final average phonon numbers n1 (blue curves) and n2 (red curves) as functions of ω2/ω1 in both the dark-mode- unbreaking case (η/ω1 = 0, solid curves) and the dark-mode-breaking case (η/ω1 = 0.05 and θ = π/2, dashed curves) under either f f κ/ω1 = 0.2 or κ/ω1 = 1.2. (f) The final average phonon numbers n1 (blue curves) and n2 (red curves) versus κ/ω1 in both the dark- mode-unbreaking case (η/ω1 = 0, solid curves) and the dark-mode-breaking case (η/ω1 = 0.05 and θ = π/2, dashed curves) when ω1 = ω2. Here, we consider red-sideband resonance driving ∆ = ω1. Other used parameters are given by G1/ω1 = G2/ω1 = 0.1, −5 3 γ1/ω1 = γ2/ω1 = 10 , and n¯1 =n ¯2 = 10 . with C being the noise correlation matrix defined by the matrix elements N (s)N (s′) = C δ(s s′). () h k l i k,l − For the Markovian baths considered in this work, the constant matrix C is given by 00 02κ 0 0 00 0 02γ1(¯n1 + 1) 0  00 00 0 2γ (¯n + 1)  C = 2 2 . (S27)  0000 0 0     0 2γ1n¯1 00 0 0   0 0 2γ n¯ 0 0 0   2 2    7

Based on the covariance matrix V, the final average phonon numbers in the two mechanical resonators are obtained by 1 nf = δb† δb = V , (S28a) 1 h 1 1i 52 − 2 1 nf = δb† δb = V , (S28b) 2 h 2 2i 63 − 2

where V52 and V63 can be obtained by solving the Lyapunov equation (S24). f f In Figs. S1(a) and S1(b), we plot the final average phonon numbers n1 and n2 as functions of the ratio ω2/ω1 (the resonance frequency of the second mechanical resonator over that of the first mechanical resonator) and the scaled cavity- field decay rate κ/ω1 when the phase-dependent phonon-exchange coupling is absent (η = 0), i.e., in the dark-mode- unbreaking case. Here, we can see that there exists a peak around ω2 = ω1, which means that the two mechanical resonators cannot be cooled in the degenerate and near-degenerate two-resonator cases. This phenomenon can be clearly explained based on the dark-mode effect. When ω1 = ω2, the two mechanical resonators form two hybrid mechanical modes: a bright mode and a dark mode. The dark mode is decoupled from both the cavity-field mode and the bright mechanical mode and hence the excitation energy stored in the dark mode cannot be extracted through the optomechanical- cooling channel. When the two mechanical resonators are far-off-resonant with each other, there is no dark mode, then the ground-state cooling can be realized when this system works in the resolved-sideband regime and under proper driving condition (red-sideband resonance). The dark-mode effect can be broken by introducing a phase-dependent phonon-exchange interaction between the two mechanical resonators, and then the ground-state cooling can be realized in the degenerate and near-degenerate two- f f mechanical-resonator cases. In Figs. S1(c) and S1(d), we plot the final average phonon numbers n1 and n2 in the two mechanical resonators as functions of the ratio ω2/ω1 and the scaled cavity-field decay rate κ/ω1 in the dark-mode- breaking case (η/ω1 = 0.05 and θ = π/2). Different from the results in Figs. S1(a) and S1(b), here we can see that f the simutaneous ground-state cooling can be realized (n1,2 1) in the resolved-sideband regime (κ ω1), which is consistent with the sideband-cooling results in a typical optomechanical≪ system. In addition, simultaneous≪ ground-state cooling of the two mechanical resonators can be reached in a wide parameter range of ω2/ω1. We also see that the cooling f f performance of the first resonator is better than that of the second resonator (n1

3 1 4 (a) (b) θ=π/2 (c) θ=0.9 π Unbreaking

500 0.5 2

0 0 0 1 2 3 4 0 1 2 3 4 0 1 2 3 4 G G G /G G2/G1 2/ 1 2 1 3 1 (d) θ=π 4 (e) θ=1.1 π (f) θ=3π/2

0.5 2

0.1 0 0 0 1 2 3 4 0 1 2 3 4 0 1 2 3 4 G /G G /G G2/G1 2 1 2 1

f f FIG. S2. (Color online) The final average phonon numbers n1 (blue solid curves) and n2 (red dashed curves) as functions of the ratio G2/G1 when the phonon-exchange coupling parameters θ and η take various values: (a) η/ω1 = 0, (b-f) η/ω1 = 0.05 and θ = π/2, 0.9π, π, 1.1π, and 3π/2. Here we choose the optimal driving ∆ = ω1 = ω2 = ωm, κ/ωm = 0.2, G1/ωm = 0.1, −5 3 γ1/ωm = γ2/ωm = 10 , and n¯1 =n ¯2 = 10 .

(i.e., the excitation-energy extraction efficiency through the cavity-field decay channel) and the phonon-sidebandresolution condition. In the above discussions concerning Fig. S1, we only consider the symmetric-coupling case, i.e., G1 = G2. To better f understand quantum cooling in this system, we also investigate the dependence of the final average phonon numbers n1 f and n2 on the linearized optomechanical-coupling strengths G1 and G2. In Fig. S2, we plot the final average phonon f f numbers n1 and n2 as functions of the ratio G2/G1 when the phonon-exchange coupling parameters η and θ take various values: (a) η/ωm = 0, (b-f) η/ωm = 0.05 and θ = π/2, 0.9π, π, 1.1π, and 3π/2. When the phonon-exchange coupling is absent, i.e., η = 0 [Fig. S2(a)], the final average phonon number in the first (second) mechanical resonator increases (decreases) with the increase of G2/G1. However, we point out that, due to the dark-mode effect, the ground-state cooling of the two mechanical resonators are unfeasible for finite values of the ratio G2/G1. When G2/G1 < 1, the bright mechanical mode is dominated by mode b1. When G2/G1 > 1, the bright mechanical mode is dominated by mode b2. As a result, the cooling efficiency of the first mechanical resonator is better (worse) than that of the second one in the parameter range G2/G1 < 1 (G2/G1 > 1). The cooling performance of the two resonators is exchanged when the value of the ratio G2/G1 changes across the point G2/G1 = 1. In the symmetric-coupling case G2/G1 = 1, the same cooling performance is achieved for the two mechanical resonators (nf = nf 500). The physical reason is that the 1 2 ≈ optomechanical-cooling channels for the two mechanical resonators take the same role when G1 = G2. At this point, the superposition amplitudes of the two mechanical modes b1 and b2 in the bright and dark modes are the same, as shown in Eq. (S7b). In the presence of the phonon-exchange coupling, the ground-state cooling can be realized in a wide parameter range of the ansymmetric couplings G = G when θ = nπ for integer n. In addition, we can see a similar intersection 2 6 1 6 phenomenon for the cooling performance of the two resonators with the increase of the ratio G2/G1. However, the location of the intersection point moves to the right (left) from the point G2/G1 = 1 when the phase θ takes the value in the range 0 <θ<π (π<θ< 2π). This shift is caused by the phase-dependent phonon-exchange coupling between the two mechanical resonators. When 0 <θ<π, the phonon-exchange coupling assists the cooling of the first mechanical f f resonator (i.e., decreasing n1 and increasing n2 ). Hence the phonon-exchange coupling pushes the intersection point moving right. When π<θ< 2π, the phonon-exchange coupling assists the cooling of the second mechanical resonator 9

3 3 (a) θ=π (b) /2 η=0.05ωm

κ=0.2 ωm 2 2 κ=ωm

κ=1.5 ωm

0 0.1 0.2 0 1 2

η/ωm /πθ 3 3 (c) = θ π/2 (d) η=0.05ωm

2 2

0 0.1 0.2 0 1 2

η/ωm /πθ

f f FIG. . (Color online) The final average phonon numbers (a,c) n1 and (b,d) n2 versus either (a,b) the coupling strength η at θ = π/2 or (c,d) the phase θ at η = 0.05ωm when the cavity-field decay rate takes various values: κ/ωm = 0.2, 1, and 1.5. Here we choose −5 3 ∆= ω1 = ω2 = ωm, G1/ωm = G2/ωm = 0.1, γ1/ωm = γ2/ωm = 10 , and n¯1 =n ¯2 = 10 .

f f (i.e., decreasing n2 and increasing n1 ). As a result, the phonon-exchange coupling pushes the intersection point moving left. At θ = π [panel (d)], the dark mode appears in this system when G1 = G2, then the two mechanical modes cannot be cooled. In this case, the dark-mode effect can be broken by choosing different values of the coupling strengths G = G , 1 6 2 i.e., simultaneous ground-state cooling of the two mechanical resonators can only be realized when G2/G1 0.5. The phase-dependent phonon-exchange interaction plays a critical role in the ground-state cooling of≤ the multiple mechanical resonators. Below we investigate the dependence of the cooling performance on the coupling parameters η and θ of the phase-dependent phonon-exchange interaction between the two mechanical resonators. In Figs. S3(a) f f and S3(b), we plot the final average phonon numbers n1 and n2 as functions of the coupling strength η and phase θ when the cavity-field decay rate takes various values: κ/ωm = 0.2, 1, and 1.5. Here, we can see that the two mechanical resonators can be cooled efficiently (from the initial phonon number 1000 to the final phonon number below 10) when η/ωm > 0.02. In addition, the cooling performance becomes worse for a larger value of the cavity-field decay rate κ. The ground-state cooling can only be realized in the resolved-sideband regime κ/ωm < 1. We also show the dependence of f f the final average phonon numbers n1 and n2 on the phase θ for several values of κ/ωm, as shown in Figs. S3(c) and S3(d). f f The plots show that the cooling performance depends on the phase θ. The final average phonon numbers n1 and n2 can be largely decreased when 0 <θ<π and π<θ< 2π. When θ = nπ for an integer n, the two mechanical resonators cannot be cooled due to the dark-mode effect. The cooling performance becomes worse with the increase of the cavity- f f f f field decay rate. In addition, the results show that n1 n2 ) in the parameter range 0 <θ<π (π<θ< 2π), which can be explained based on the nonreciprocal phonon transfer induced by quantum interference in the loop-coupled system. f f In Fig. S3, we have investigatedthe dependenceof the final average phonon numbers n1 and n2 on the phonon-exchange coupling parameters η and θ in the degenerate two-mechanical-resonator case, i.e., ω1 = ω2. In the following we also f f consider a nondegenerate mechanical-resonator case. In Fig. S4 we plot the final average phonon numbers n1 and n2 versus the parameters η and θ in the nondegenerate two-resonator cases, i.e., ω2 = 0.8ω1 or ω2 = 1.2ω1. The plots show that the simultaneous ground-state cooling of the two mechanical resonators can be realized in the nondegenerate f f mechanical-resonator case. In both the cases ω2 = 0.8ω1 and ω2 = 1.2ω1, the dependence of n1 and n2 on the phase θ 10

0.8 0.8 (a) ω2=0.8ω1 (b) ω2=0.8ω1 θ=π/2 η=0.05ω1 0.6 0.6

0.4 0.4

0.2 0.2 0 1 2 0 0.1 0.2

/πθ η/ω1 0.8 0.8 ω =1.2ω (c) 2 1 (d) ω2=1.2ω1

η=0.05ω1 θ=π/2 0.6 0.6

0.4 0.4

0.2 0.2 0 1 2 0 0.1 0.2

/πθ η/ω1

f f FIG. S4. (Color online) The final average phonon numbers n1 (blue solid curves) and n2 (red dashed curves) versus the phase θ and the phonon-exchange coupling strength η in the nondegenerate two-resonator cases (a,b) ω2 = 0.8ω1 and (c,d) ω2 = 1.2ω1. Here we −5 3 choose ∆= ω1, G1/ω1 = G2/ω1 = 0.1, κ/ω1 = 0.2, γ1/ω1 = γ2/ω1 = 10 , and n¯1 =n ¯2 = 10 .

f has an inverse tendency, as shown in Figs. S4(a) and S4(c). In addition, the dependence of nl=1,2 on the phase θ in the f f case ω2 = 0.8ω1 is inverse to that in the case of ω2 = 1.2ω1. In Figs. S4(b) and S4(d), we can see n1 < n2 and the f dependence of nl=1,2 on the coupling strength η has a similar tendency for the cases ω2 = 0.8ω1 and ω2 = 1.2ω1. In the nondegenerate-resonator case, the cooling performance can be controlled by choosing proper phonon-exchange coupling f f parameters η and θ. The same value of the final phonon numbers n1 and n2 can be obtained by choosing the intersection points in Figs. S4(a) and S4(c). In quantum cooling of the mechanical resonators, the optomechanical cavity and its vacuum bath provide the cooling channel to extract the excitation energy in the mechanical resonators. Here, the mechanical resonators are thermalized f by their thermal baths through the mechanical dissipation channels. As a result, the final average phonon numbers n1 f and n2 in the two mechanical resonators depend on the mechanical decay rates γ1 and γ2. In Fig. , we show the final f f f f average phonon numbers n1 and n2 as functions of the decay rates γ1 and γ2. We can see that n1 and n2 increase with the increase of the mechanical decay rates. This is because the energy exchange rates between the mechanical resonators and their heat baths are faster for larger values of the decay rates, and then the thermal excitation in the heat baths will f f raise the total phonon numbers in the mechanical resonators. In Figs. S5(a) and S5(b), we have n1 < n2 because the phase angle θ = π/2 is taken, then the cooling performance of the first resonator is better than that of the second resonator. However, an opposite cooling effect compared with the case of θ = π/2 emerges when θ = 3π/2, as shown in Figs. S5(c) and S5(d). These interesting cooling phenomena can be explained according to the phonon scattering process between the two mechanical resonators, which will be studied in the next section.

S3. PHONON SCATTERING PROBABILITY AND NONRECIPROCAL PHONON TRANSFER

In this section, we study the scattering probabilities of the phonon transport between the two mechanical resonators coupled by a phase-dependent phonon-exchange interaction. We calculate the transmission spectrum of the phonon 11

1 1 (a) θ=π/2 (b) θ=π/2

0.5 0.5

0 0 1 1 (c) θ=3 π/2 (d) θ=3 π/2

0.5 0.5

0 0 0 1 2 0 1 2 -5 -5 γ1/wm ×10 γ2/wm ×10

f f FIG. S5. (Color online) The final average phonon numbers n1 and n2 as functions of (a,c) γ1 and (b,d) γ2 when the phase θ takes −5 −5 different values: (a,b) θ = π/2 and (c,d) θ = 3π/2. In panels (a,c) and (b,d), we choose γ2/ω1 = 10 and γ1/ω1 = 10 , respectively. 3 Other used parameters are ∆= ω1 = ω2 = ωm, G1/ωm = G2/ωm = 0.1, η/ωm = 0.05, κ/ωm = 0.2, and n¯1 =n ¯2 = 10 .

transport based on the Langevin equation (S18). To this end, we rewrite the matrix N(t) defined in Eq. (S20) as

N(t)= Γuin(t), (S29) where the damping matrix Γ is defined as

Γ = diag[√2κ, 2γ1, 2γ2, √2κ, 2γ1, 2γ2], (S30) with diag[x] giving a matrix with the elements ofp the listpx on the leadingp diagonal,p and 0 elsewhere. The input noise vector uin(t) in Eq. (S29) is given by † † † T uin(t) = [ain(t), b1,in(t), b2,in(t), ain(t), b1,in(t), b2,in(t)] . (S31) Making use of the Fourier transformation for operator r δa, δb , δb , δa , δb , δb and its conjugate r†, ∈ { 1 2 in 1,in 2,in} 1 ∞ r˜(ω)= eiωtr(t)dt, (S32a) 2π Z−∞ 1 ∞ r˜†(ω)= eiωtr†(t)dt, (S32b) 2π Z−∞ the solutions to the linearized quantum Langevin equation (S18) in the frequency domain can be obtained as ˜u(ω) = ( iωI A)−1Γ˜u (ω), (S33) − − in where ˜u(ω) and ˜uin(ω) are, respectively, the Fourier transformation of the operator vectors u(t) defined in Eq. (S19) and uin(t) defined in Eq. (S31). The matrix I in Eq. (S33) is an identity matrix. Using the input-output relation

oin + oout = 2γoδo (S34) for o a, b1, b2 and γo κ, γ1, γ2 , we obtain the output fieldp in the frequency domain as ∈ { } ∈ { } ˜uout(ω)= U(ω)˜uin(ω), (S35) 12

where the transformation matrix is given by U(ω)= Γ( iωI A)−1Γ I, (S36) − − − and ˜ ˜ † ˜† ˜† T ˜uout(ω)=[˜aout(ω), b1,out(ω), b2,out(ω), a˜out(ω), b1,out(ω), b2,out(ω)] (S37) denotes the Fourier transformation of uout(t). To analyze the excitation energy transfer in this system, we introduce the spectra for the input and output signals as T Sin(ω) =[sa,in(ω),sb1 ,in(ω),sb2,in(ω)] , (S38a) T Sout(ω) =[sa,out(ω),sb1,out(ω),sb2,out(ω)] , (S38b) where the elements are defined by

† ′ ′ o˜out(ω )˜oout(ω) =so,outδ(ω + ω ), (S39a) h ′ i ′ † o˜in(ω )˜oin(ω) =so,inδ(ω + ω ), (S39b) h ′ i ′ o˜ (ω )˜o† (ω) =(1 + s )δ(ω + ω ). (S39c) h in in i o,in We also define the spectrum for the input vacuum noise as

T Svac(ω) = [sa,vac(ω),sb1,vac(ω),sb2,vac(ω)] , (S40) with s (ω)= U (ω) 2 + U (ω) 2 + U (ω) 2, (S41a) a,vac | 14 | | 15 | | 16 | s (ω)= U (ω) 2 + U (ω) 2 + U (ω) 2, (S41b) b1,vac | 24 | | 25 | | 26 | s (ω)= U (ω) 2 + U (ω) 2 + U (ω) 2. (S41c) b2,vac | 34 | | 35 | | 36 | Then the relation between these spectra can be obtained as

Sout(ω)= T(ω)Sin(ω)+ Svac(ω), (S42) where the transmission matrix T(ω) is defined by

Taa(ω) Tab1 (ω) Tab2 (ω) T(ω)= T (ω) T (ω) T (ω) , (S43)  b1a b1b1 b1b2  Tb2a(ω) Tb2b1 (ω) Tb2b2 (ω) with these matrix elements   T (ω)= U (ω) 2 + U (ω) 2, aa | 11 | | 14 | T (ω)= U (ω) 2 + U (ω) 2, ab1 | 12 | | 15 | T (ω)= U (ω) 2 + U (ω) 2, ab2 | 13 | | 16 | T (ω)= U (ω) 2 + U (ω) 2, b1a | 21 | | 24 | T (ω)= U (ω) 2 + U (ω) 2, b1b1 | 22 | | 25 | T (ω)= U (ω) 2 + U (ω) 2, b1b2 | 23 | | 26 | T (ω)= U (ω) 2 + U (ω) 2, b2a | 31 | | 34 | T (ω)= U (ω) 2 + U (ω) 2, b2b1 | 32 | | 35 | T (ω)= U (ω) 2 + U (ω) 2. (S44) b2b2 | 33 | | 36 | The element Tvw(ω) (v, w a, b1, b2 ) denotes the transmittance from the input mode w to the output mode v. To explore the phonon-transfer∈ nonreciprocity { } between the two mechanical modes, we only focus on the transmittance

Tb1b2 (ω) and Tb2b1 (ω) between the two mechanical modes. Then, we numerically evaluate the transmittance between the two mechanical modes to show the nonreciprocal phonon transfer. Physically, the transmittance Tb1b2 (ω) and Tb2b1 (ω) can be used to analyze the thermal excitations extracted from one mechanical mode to the other one. The above results concerning the phonon transmission are exact. Below we derive some approximate analytical results under the RWA and the resonance condition ∆ = ω1 = ω2 = ωm. Note that under the RWA, we have the approximate 13

(a) = /2 (b) =3 /2 L L

L L

- - 0.8 0.91 1.1 1.2 0.8 0.91 1.1 1.2 w/wm w/wm

(c) (d) L L L

Exact Approximate

L

- - 0.050.1 0.15 0.5 1.5

FIG. S6. (Color online) (a,b) The relative phonon-scattering rate Λb2b1 (blue curves) and Λb1b2 (red curves) as functions of ω when the phase θ takes different values: (a) θ = π/2 and (b) θ = 3π/2. In panels (a,b), we choose the phonon-exchange coupling η/ωm = 0.05.

(c,d) The exact (solid/dashed lines) and approximate (symbols) relative resonant-phonon-scattering rates Λb2b1 and Λb1b2 vs (c) the phonon-exchange coupling η when θ = π/2 and (d) the phase θ when η/ωm = 0.05 under the parameter ω = ωm. Here we take −5 3 ∆= ω1 = ω2 = ωm, G1/ωm = G2/ωm = 0.1, κ/ωm = 0.2, γ1/ωm = γ2/ωm = 10 , and n¯1 =n ¯2 = 10 .

relations T (ω) U (ω) 2 and T (ω) U (ω) 2. In particular, we focus on the resonant phonon transmission b1b2 ≈ | 23 | b2b1 ≈ | 32 | at the mechanical frequency ωm, then an analytical transmittance between the two mechanical modes can be obtained as

2 2 2 4γ1γ2[(G1G2) + (κη) 2G1G2κη sin θ] Tb1b2 U23 = 2 2 2−2 2 ≈| | (G2γ1 + G1γ2 + κγ1γ2 + κη ) + 4(G1G2η cos θ) 4( + 2√ sin θ) 1 2 3 1 2 3 (S45a) = C C C − 2 C C C 2 , ( 1 + 2 + 3 + 1) + 4 1 2 3 cos θ C C C C C 2C 2 2 4γ1γ2[(G1G2) + (κη) + 2G1G2κη sin θ] Tb2b1 U32 = 2 2 2 2 2 ≈| | (G2γ1 + G1γ2 + κγ1γ2 + κη ) + 4(G1G2η cos θ) 4( + + 2√ sin θ) = C1C2 C3 C1C2C3 , (S45b) ( + + + 1)2 + 4 cos2 θ C1 C2 C3 C1C2C3 where we introduce the cooperativities between any two subsystems in this two-mechanical-mode optomechanical system as 2 G1 1 = , (S46a) C γ1κ 2 G2 2 = , (S46b) C γ2κ η2 3 = . (S46c) C γ1γ2 14

(a) (b) Π=0.1 Π=0.5 Π=1 L L

- - 0.5 1.5 0.5 1.5

(c) (d) = /2 =3 /2

L L

Optimal nonreciprocity Optimal nonreciprocity

=3 /2 = /2 - - 1 3 4 1 3 4 Π Π

FIG. . (Color online) Dependence of the relative phonon-scattering rates (a) Λb2b1 and (b) Λb1b2 on the phase θ when ω = ωm

and the optomechanical cooperativity takes various values: Π = 0.1, 0.5, and 1. The relative phonon-scattering rates (c) Λb2b1 and (d)

Λb1b2 versus the ratio of the optomechanical cooperativities Π when θ = π/2 and θ = 3π/2. Here we take ∆ = ω1 = ω2 = ωm, −5 3 G1/ωm = G2/ωm = 0.1, κ/ωm = 0.2, γ1/ωm = γ2/ωm = 10 , and n¯1 =n ¯2 = 10 .

According to Eqs. (S45a) and(S45b), the maximum transmittance for either θ = π/2 or θ = 3π/2 can be obtained as 4(√ + √ )2 (T ) = (T ) = C1C2 C3 . () b2b1 max b1 b2 max ( + + + 1)2 C1 C2 C3 By introducing a relative phonon-scattering rate from the mechanical modes w to v as

Tvw Twv Λvw = − , (S48) (Tvw)max we can then obtain the rates between the two mechanical modes b1 and b2 as T T 4√ sin θ Λ = b2b1 − b1b2 = C1C2C3 , (S49a) b2b1 2 4C1C2C3 cos θ (Tb2b1 )max 2 (√ 1 2 + √ 3) 1+ 2 C C C (C1+C2+C3+1)   Tb1b2 Tb2b1 Λb1b2 = − = Λb2b1 . (S49b) (Tb1b2 )max −

In Figs. S6(a) and S6(b), the relative phonon-scattering rates Λb2b1 (blue curves) and Λb1b2 (red curves) are plotted as functions of the scaled frequency ω/ωm when the phase θ takes different values: (a) θ = π/2 and (b) θ = 3π/2. It is obviously shown that the reciprocity of the phonon transfer between the two mechanical resonators is broken (Λb2b1 = 0) in a wide range of ω and the phonon transfer exhibits a perfect nonreciprocal response when θ = π/2 and θ = 36π/2.

When θ = π/2 (θ = 3π/2), we have Tb2b1 > 0 and Tb1b2 < 0 (Tb2b1 < 0 and Tb1b2 > 0). In particular, when ω = ωm

and θ = π/2, we have Λb2b1 = 1, i.e., Tb1b2 = 0. This means that the unidirectional flow of the phonons from b1 to

b2 is achieved. When ω = ωm and θ = 3π/2, we have Λb1b2 = 1, i.e., Tb2b1 = 0. This means the phonons can only be transferred from b2 to b1. Based on the above results, we can see that the phase-dependent phonon-exchange coupling 15

plays an effective role on the relative phonon scattering between the two mechanical resonators. In Figs. S6(c) and S6(d), we show the dependence of the relative resonant-phonon-scattering rates on the phonon-exchange coupling parameters η and θ. The results indicate that a perfect nonreciprocal phonon transfer requires both η 0.05ω and θ = π/2 or 3π/2. ≈ m Moreover, the exact calculations and the approximate analytical results match well with each other. Here, the solid (Λb2b1 )

and dashed lines (Λb1b2 ) are plotted using the exact solutions, while the symbols are based on the analytical calculations

given in Eqs. (S49a) and (S49b). In Fig. S6(d), when 0 <θ<π, it shows Λb2b1 > 0, i.e., Tb2b1 > Tb1b2 . In the region

π<θ< 2π, it exhibits Λb1b2 > 0, i.e., Tb1b2 > Tb2b1 . Meanwhile, the phonon transmission satisfies the reciprocity

[Λb2b1 = Λb1b2 = 0, i.e., Tb1b2 = Tb2b1 ] at θ = nπ. Moreover, the transmittance is optimal for the process from b1 (b2)

to b2 (b1) and is zero for the opposite process when θ = π/2 (θ = 3π/2), namely, Tb1b2 = 0 and Tb2b1 = 0 at θ = π/2 and θ = 3π/2, respectively. In order to analyze the optomechanical cooperativities among the two subsystems in this three-mode optomechanical system, we introduce a new parameter defined by Π= 3/( 1 2), which is the ratio of the optomechanicalcooperativities. Thus, the analytical solutions given in Eqs. (S49a) and(CS49bC )C become

4√Πsin θ Λb2b1 = , (S50a) 2 4Π cos2 θ √ C C 2 (1 + Π) 1+ 1+ 2+1 ( C C +Π)  1 2  Λ = Λ . (S50b) b1b2 − b2b1

It can be seen from Eqs. (S50a) and (S50b) that the relative nonreciprocal phonon transfer Λb2b1 = 1 (Λb1b2 = 1) is obtained at Π = 1 and θ = π/2 (3π/2). In Figs. S7(a) and S7(b), we plot the relative phonon-scattering rates Λb2b1 and Λb1b2 as functions of the phase θ when Π takes various values: Π = 0.1, 0.5, and 1. The results show that the optimal nonreciprocity appears at Π = 1 and either θ = π/2 or 3π/2. When Π = 1, the absolute value of the relative 6 phonon-scattering rate will be decreased at a given phase θ. We also plot the relative phonon-scattering rates Λb2b1 and

Λb1b2 versus the ratio Π when the phase takes θ = π/2 (solid lines) and θ = 3π/2 (dashed lines), as shown in Figs. S7(c)

and S7(d). In the region 0 < Π < 1, the nonreciprocal phonon-transfer rate Λb2b1 increases with the increase of Π. In the region Π > 1, the relative nonreciprocal phonon-transfer rate is suppressed. The optimal nonreciprocity emerges at Π = 1, which indicates directional flow of phonons between the two mechanical resonators.

S4. THE COOLING LIMITS OF THE TWO MECHANICAL RESONATORS

In this section, we present a detailed derivation of the cooling limits of the two mechanical resonators, which are obtained by adiabatically eliminating the cavity-field mode in the large cavity-field decay regime. In this case, the system is reduced to a two-coupled mechanical resonator system. The derivation of the cooling limits is based on the Langevin equations (S4) for the quantum fluctuations of the system operators. To obtain the cooling limits, we consider the case where the linearized optomechanical coupling strengths G1,2 are real and the system works in the parameter regime:

ω κ G γ . (S51) 1,2 ≫ ≫ 1,2 ≫ 1,2 In this case, the cavity field can be eliminated adiabatically, and then the solution of the cavity-field fluctuation operator δa(t) at the time scale t 1/κ can be obtained as ≫ iG iG iG iG δa(t) 1 δb† (t) 1 δb (t) 2 δb† (t) 2 δb (t)+ F (t), ≈−κ + i(∆ + ω ) 1 − κ + i(∆ ω ) 1 − κ + i(∆ + ω ) 2 − κ + i(∆ ω ) 2 a,in 1 − 1 2 − 2 (S52)

where we introduce the new noise operator

t −(κ+i∆)t (κ+i∆)s Fa,in(t)= √2κe ain(s)e ds. (S53) Z0 16

Substitution of Eq. (S52) into Eqs. (S4b) and (S4c) leads to the equations of motion G2 G2 G2 G2 δb˙ (t)= 1 1 δb† (t)+ 1 1 (γ + iω ) δb (t) 1 κ i(∆ ω ) − κ + i(∆ + ω ) 1 κ i(∆ + ω ) − κ + i(∆ ω ) − 1 1 1  − − 1 1   − 1 − 1  G G G G G G G G + 1 2 1 2 δb† (t)+ 1 2 1 2 iηeiθ δb (t) κ i(∆ ω ) − κ + i(∆ + ω ) 2 κ i(∆ + ω ) − κ + i(∆ ω ) − 2  − − 2 2   − 2 − 2  iG F (t) iG F † (t)+ 2γ b (t), (S54a) − 1 a,in − 1 a,in 1 1,in G G G G G G G G δb˙ (t)= 1 2 1 2 p δb† (t)+ 1 2 1 2 iηe−iθ δb (t) 2 κ i(∆ ω ) − κ + i(∆ + ω ) 1 κ i(∆ + ω ) − κ + i(∆ ω ) − 1  − − 1 1   − 1 − 1  G2 G2 G2 G2 + 2 2 δb† (t)+ (γ + iω )+ 2 2 δb (t) κ i(∆ ω ) − κ + i(∆ + ω ) 2 − 2 2 κ i(∆ + ω ) − κ + i(∆ ω ) 2  − − 2 2   − 2 − 2  iG F (t) iG F † (t)+ 2γ b (t). (S54b) − 2 a,in − 2 a,in 2 2,in By making the RWA in Eqs. (S54a) and(S54bp ), we have δb˙ (t)= (Γ + iΩ )δb (t)+ ξ δb (t) iG F (t) iG F † (t)+ 2γ b (t), (S55a) 1 − 1 1 1 1 2 − 1 a,in − 1 a,in 1 1,in δb˙ (t)=ξ δb (t) (Γ + iΩ )δb (t) iG F (t) iG F † (t)+ 2pγ b (t), (S55b) 2 2 1 − 2 2 2 − 2 a,in − 2 a,in 2 2,in where we introduce the effective resonance frequency Ωl and decay rate Γl for the lth mechanicalp resonator Ω =ω ω , (S56a) l l − l,opt Γl =γl + γl,opt, (S56b) with G2(∆ + ω ) G2(∆ ω ) ω = l l + l − l , (S57a) l,opt κ2 +(∆+ ω )2 κ2 + (∆ ω )2 l − l G2κ G2κ γ = l l , l = 1, 2. (S57b) l,opt κ2 + (∆ ω )2 − κ2 +(∆+ ω )2 − l l Here, ωl,opt and γl,opt denote the resonance frequency shift and the additional energy decay rate induced by the optomechanical couplings, respectively. We also introduce the effective coupling strengths between the two mechanical modes b1 and b2 after adiabatically eliminating the cavity mode as G G [κ + i(∆ + ω )] G G [κ i(∆ ω )] ξ = 1 2 2 1 2 − − 2 iηeiθ, (S58a) 1 κ2 +(∆+ ω )2 − κ2 + (∆ ω )2 − 2 − 2 G G [κ + i(∆ + ω )] G G [κ i(∆ ω )] ξ = 1 2 1 1 2 − − 1 iηe−iθ. (S58b) 2 κ2 +(∆+ ω )2 − κ2 + (∆ ω )2 − 1 − 1 Under the parameter condition ω κ G and at resonance ∆= ω = ω , we have 1,2 ≫ ≫ 1,2 1 2 G G G G ξ 1 2 + i ηeiθ 1 2 , (S59a) 1 ≈− κ − 2ω   2  G G G G ξ 1 2 + i ηe−iθ 1 2 , (S59b) 2 ≈− κ − 2ω   1  and G2 γ l , (S60a) l,opt ≈ κ 2 Gl ωl,opt , l = 1, 2. (S60b) ≈2ωl The final average phonon numbers (namely the steady-state values of the phonon numbers) can be obtained by solving Eq. (S55). To be concise, we reexpress Eq. (S55) as ˙v(t)= Mv(t)+ N(t), (S61) − 17

T where v(t) = (δb1(t), δb2(t)) , M is defined by Γ + iΩ ξ M = 1 1 − 1 , (S62) ξ2 Γ2 + iΩ2  −  and N(t) reads iG F (t) iG F † (t)+ √2γ b (t) N(t)= − 1 a,in − 1 a,in 1 in,1 . (S63) iG F (t) iG F † (t)+ √2γ b (t) − 2 a,in − 2 a,in 2 in,2 ! The formal solution of Eq. (S61) can be expressed as t v(t)= e−Mtv(0) + e−Mt eMsN(s)ds. (S64) Z0 The final average phonon numbers can be obtained by calculating the elements of the variance matrix. By a lengthy calculation, we obtain the approximate analytical expressions for the final average phonon numbers as

2 ∗ f γ1n¯1 [u Γ1 +Γ2 i(Ω1 Ω2)] [u Γ1 +Γ2 + i(Ω1 Ω2)][u +Γ1 Γ2 + i(Ω1 Ω2)] n1 = 2 | − ∗ − − | + 2Re − − ∗ − − 2 u λ1 + λ1 λ1 + λ2 | |  h i [u +Γ Γ + i(Ω Ω )] 2 G2 [u Γ +Γ i(Ω Ω )] 2 1 1 +| 1 − 2 1 − 2 | + 1 | − 1 2 − 1 − 2 | + λ∗ + λ 4 u 2 λ∗ + λ κ + λ + i∆ κ + λ∗ i∆ 2 2   1 1 1 1 ∗ | |  −  [u Γ1 +Γ2 + i(Ω1 Ω2)][u +Γ1 Γ2 + i(Ω1 Ω2)] 1 1 +2Re − − ∗ − − + ∗ λ1 + λ2 κ + λ2 + i∆ κ + λ1 i∆ h 2  − i [u +Γ1 Γ2 + i(Ω1 Ω2)] 1 1 +| − ∗ − | + ∗ λ2 + λ2 κ + λ2 + i∆ κ + λ2 i∆  −  2 ξ1 2 1 1 1 1 1 1 +| |2 G2 ∗ + ∗ 2Re ∗ + ∗ u " λ1 + λ1 κ + λ1 + i∆ κ + λ1 i∆ − λ1 + λ2 κ + λ2 + i∆ κ + λ1 i∆ | |   −  h  − i 1 1 1 1 1 1 1 + ∗ + ∗ + 2γ2n¯2 ∗ ∗ ∗ + ∗ , (S65) λ2 + λ2 κ + λ2 + i∆ κ + λ2 i∆ λ1 + λ1 − λ1 + λ2 − λ2 + λ1 λ2 + λ2 #  −    and 2 ∗ f γ2n¯2 u +Γ1 Γ2 + i(Ω1 Ω2) [u +Γ1 Γ2 i(Ω1 Ω2)][u Γ1 +Γ2 i(Ω1 Ω2)] n2 = 2 | − ∗ − | + 2Re − − − ∗ − − − 2 u λ1 + λ1 λ1 + λ2 | |  h i u Γ +Γ i(Ω Ω ) 2 G2 u +Γ Γ + i(Ω Ω ) 2 1 1 +| − 1 2 − 1 − 2 | + 2 | 1 − 2 1 − 2 | + λ∗ + λ 4 u 2 λ∗ + λ κ + λ + i∆ κ + λ∗ i∆ 2 2   1 1 1 1 |∗ |  −  [u Γ1 +Γ2 i(Ω1 Ω2)][u +Γ1 Γ2 i(Ω1 Ω2)] 1 1 +2Re − − − ∗ − − − + ∗ λ1 + λ2 κ + λ2 + i∆ κ + λ1 i∆ h 2  − i u Γ1 +Γ2 i(Ω1 Ω2) 1 1 +| − ∗ − − | + ∗ λ2 + λ2 κ + λ2 + i∆ κ + λ2 i∆  −  2 ξ2 2 1 1 1 1 1 1 +| |2 G1 ∗ + ∗ 2Re ∗ + ∗ u " λ1 + λ1 κ + λ1 + i∆ κ + λ1 i∆ − λ2 + λ1 κ + λ2 + i∆ κ + λ1 i∆ | |   −  h  − i 1 1 1 1 1 1 1 + ∗ + ∗ + 2γ1n¯1 ∗ ∗ ∗ + ∗ , (S66) λ2 + λ2 κ + λ2 + i∆ κ + λ2 i∆ λ1 + λ1 − λ1 + λ2 − λ2 + λ1 λ2 + λ2 #  −    ∗ ∗ where λ1 and λ2 (λ1 and λ2 being complex conjugate) are the eigenvalues of the coefficient matrix M, 1 λ = [Γ +Γ + i(Ω +Ω ) u], (S67a) 1 2 1 2 1 2 − 1 λ = [Γ +Γ + i(Ω +Ω )+ u] (S67b) 2 2 1 2 1 2 18

where u = 4ξ ξ + [Γ Γ + i(Ω Ω )]2. (S68) 1 2 1 − 2 1 − 2 q For the case ω1 ω2 and Γ1 Γ2, the approximate analytical expressions of the final average phonon numbers can be reduced as ≈ ≈ 2 f γ1n¯1 1 1 1 G1 1 1 1 n1 ∗ + 2Re ∗ + ∗ + ∗ + ∗ ≈ 2 λ1 + λ1 λ1 + λ2 λ2 + λ2 4 λ1 + λ1 κ + λ1 + i∆ κ + λ1 i∆  h i    −  1 1 1 1 1 1 +2Re ∗ + ∗ + ∗ + ∗ λ1 + λ2 κ + λ2 + i∆ κ + λ1 i∆ λ2 + λ2 κ + λ2 + i∆ κ + λ2 i∆ h  − i  −  ξ1 2 1 1 1 1 1 1 + | | G2 ∗ + ∗ 2Re ∗ + ∗ 4 ξ2 " λ1 + λ1 κ + λ1 + i∆ κ + λ1 i∆ − λ1 + λ2 κ + λ2 + i∆ κ + λ1 i∆ | |   −  h  − i 1 1 1 1 1 1 + ∗ + ∗ + 2γ2n¯2 ∗ 2Re ∗ + ∗ , (S69) λ2 + λ2 κ + λ2 + i∆ κ + λ2 i∆ λ1 + λ1 − λ1 + λ2 λ2 + λ2 #  −   h i  and 2 f γ2n¯2 1 1 1 G2 1 1 1 n2 ∗ + 2Re ∗ + ∗ + ∗ + ∗ ≈ 2 λ1 + λ1 λ1 + λ2 λ2 + λ2 4 λ1 + λ1 κ + λ1 + i∆ κ + λ1 i∆  h i    −  1 1 1 1 1 1 +2Re ∗ + ∗ + ∗ + ∗ λ1 + λ2 κ + λ2 + i∆ κ + λ1 i∆ λ2 + λ2 κ + λ2 + i∆ κ + λ2 i∆ h  − i  −  ξ2 2 1 1 1 1 1 1 + | | G1 ∗ + ∗ 2Re ∗ + ∗ 4 ξ1 " λ1 + λ1 κ + λ1 + i∆ κ + λ1 i∆ − λ1 + λ2 κ + λ2 + i∆ κ + λ1 i∆ | |   −  h  − i 1 1 1 1 1 1 + ∗ + ∗ + 2γ1n¯1 ∗ 2Re ∗ + ∗ . (S70) λ2 + λ2 κ + λ2 + i∆ κ + λ2 i∆ λ1 + λ1 − λ1 + λ2 λ2 + λ2 #  −   h i  By substituting Eq. (S67) into Eqs. (S69) and (S70) and considering the parameters relations ω1,2 κ G1,2 γ γ γ , the final average phonon numbers can be simplified as ≫ ≫ ≫ { 1,opt ≈ 2,opt}≫ 1,2 l−1 f γln¯l + γl,optnopt ( 1) √χl(√χ1nχ1 √χ2nχ2 ) nl=1,2 + − − , (S71) ≈ Γl + χ+ Γl + χ− where we introduce the following variables 4κ2 (S72a) nopt = 2 , (ω1 + ω2 + 2∆)

2(γ2(1)n¯2(1) + γ2(1)optnopt) nχ1(2) = , (S72b) Γ1 +Γ2 + 2χ+ ξ ξ χ = √χ χ Re 1 2 , (S72c) ± ∓ 1 2 − Γ +Γ  1 2  2 ξl χl=1,2 = | | . (S72d) Γ1 +Γ2

Here, nopt stands for the effective phonon number in the optomechanical cooling bath, and χ1 and χ2 are the effective lim lim phonon-transfer rates from b2 to b1 and from b1 to b2, respectively. The corresponding cooling limits (n1 , n2 ) are obtained by taking the optimal driving detuning ∆= ωl in Eq. (S71). In particular, the first term in Eq. (S71) is contributed by the thermal bath and the effective optical bath connected by the lth mechanical resonator, while the phonon extraction contribution induced by the phonon-exchange channel is presented by the last term. Physically, the nonreciprocity of the phonon transfer is decided by the phonon-exchange rate χ which depends on the phase θ. In the case n¯ n¯ and l 1 ≈ 2 γ1 γ2, we have nχ nχ = nχ and thus (√χ1nχ √χ2nχ ) (√χ1 √χ2)nχ. In the region 0 <θ<π ≈ 1 ≈ 2 1 − 2 ≈ − (π<θ< 2π), we obtain √χ1 < √χ2 (√χ1 > √χ2). This means that the phonon-transfer rate from b1 (b2) to b2 (b1) is f f f f larger than that for the opposite case. According to Eq. (S71), we then have the relation n1 n2 ) in the region 19 2 2 (a) = /2 (b) =3 /2

Exact Approximate

1 1 Phonon number Phonon number

0 0 0 0.4 0.8 0 0.4 0.8

κ /wm κ /wm

f f FIG. S8. (Color online) The final average phonon numbers n1 and n2 are plotted as functions of κ/ωm when the phase θ takes the values: (a) θ = π/2 and (b) θ = 3π/2. The exact results are given by Eq. (S28) (solid curves) and the approximate results obtained by the adiabatic elimination method are given by Eq. (S71) (symbols). Here, the used parameters are ∆ = ω1 = ω2 = ωm, −5 G1/ωm = G2/ωm = 0.08, η/ωm = 0.05, γ1/ωm = γ2/ωm = 10 , and n¯1 =n ¯2 = 1000.

0 <θ<π (π<θ< 2π). When θ = π/2 (3π/2) and √ 1 2 = √ 3, the unidirectional flow of the phonons between the two mechanical resonators is achieved [χ 0 (χ 0)].C ForC θ =Cnπ, the phonon transfer between the two mechanical 1 ≈ 2 ≈ resonators is reciprocal (√χ1 = √χ2), due to the emergenceof the dark mode. Once the phonon-transfer channel is turned off (η = 0), the ground-state cooling is unfeasible owing to the invalid effective cooling channel (Γl + χ+ γl). In the f l−1 → absence of the optomechanical cooling channels (G1,2 = 0), Eq. (S71) becomes nl=1,2 n¯l + ( 1) (nχ1 nχ2 )/2, which indicates quantum thermalization in this coupled mechanical system. ≈ − − f f Moreover, both the exact and approximate final average phonon numbers n1 and n2 are plotted in Fig. S8 as functions of the cavity-field decay rate κ at the optimal driving detuning ∆= ωm when the modulation phase θ takes various values: f f (a) θ = π/2 and (b) θ = 3π/2. Here, the blue solid curves (n1 ) and the red dashed curves (n2 ) are plotted using the exact solutions given in Eq. (S28), while the symbols are based on the analytical calculations given in Eq. (S71). We can see from Fig. S8 that the analytical cooling limits and the exact results match well with each other when κ/ωm < 0.4, and the difference between the numerical simulation and approximate results increases when κ/ωm > 0.4. This means that the cooling performances of the two mechanical resonators are excellent in the resolved-sideband regime (κ ωm). This result is consistent with the sideband cooling results in the typical optomechanical systems. We also see from≪ Fig. S8(a) f f that the cooling performance of the first resonator is better than that of the second resonator (n1 < n2 ) when θ = π/2. f f However, when θ = 3π/2, the opposite cooling performance (n1 >n2 ) has been displayed in comparison with the case of θ = π/2, as shown in Fig. S8(b). Physically, the nonreciprocal phonon-transfer mechanism is more helpful to cool the first (second) resonator when 0 <θ<π (π<θ< 2π). In particular, the optimal cooling performances of the two mechanical resonators require that the working value of cavity-field decay rate is around κ/ωm = 0.1 0.2, as shown in Fig. S8. This is a result of the competition between the efficiency of extraction of the thermal excitations∼ and the phonon- sideband resolution condition. When κ/ωm < 0.1, the cooling performances of the two mechanical resonators become worse. Physically, the vacuum bath of the cavity field extracts the thermal excitations in the two mechanical resonators through a manner of nonequilibrium dynamics, and then the total system reaches a steady state. When the cavity-field decay rate κ is equal to 0, the vacuum bath cannot extract the thermal phonons in these two mechanical resonators, and then this system will be thermalized to a thermal equilibrium state.

S5. THE DARK-MODE EFFECT AND ITS BREAKING IN A MULTIPLE-MECHANICAL-RESONATOR OPTOMECHANICAL SYSTEM

In this section, we study the dark-mode effect in a multiple-mechanical-resonator optomechanical system, which consists of one cavity mode and N (N 3) mechanical resonators [see Figs. (a) and S9(b)]. The Hamiltonian of ≥ this system can be written in a frame rotating at the driving frequency ωL as N N N−1 † † † † ∗ † iθj † HI =∆ca a + ωj bj bj + gj a a(bj + bj )+(Ωa +Ω a )+ ηj (e bj bj+1 + H.c.), (S73) j=1 j=1 j=1 X X X 20

where ∆c = ωc ωL is the detuning of the cavity-field resonance frequency ωc with respect to the driving frequency ωL. − † † The operators a (a ) and bj (bj ) are, respectively, the annihilation (creation) operators of the cavity-field mode and the jth mechanical resonator (with the resonance frequency ωj ). The optomechanical interactions between the cavity mode and the jth mechanical resonator are described by the gj terms (with gj being the single-photon optomechanical-coupling strength). The cavity-field driving is denoted by the Ω term (with Ω being the driving amplitude). To manipulate the energy exchange between the neighboring mechanical resonators, we introduce a phase-dependent phonon-exchange interaction between the neighboring mechanical resonators, with the coupling strength ηj and the phase θj . By phenomenologically adding the damping and noise terms into the Heisenberg equations obtained based on the Hamiltonian in Eq. (S73), the quantum Langevin equations for the operators of the optical and mechanical modes can be obtained as a˙ = ia[∆ + g (b + b†)+ g (b + b†)+ + g (b + b† )] iΩ κa + √2κa , − c 1 1 1 2 2 2 · · · N N N − − in b˙ = (γ + iω )b ig a†a iη eiθ1 b + 2γ b , 1 − 1 1 1 − 1 − 1 2 1 1,in ˙ † −iθ1 iθ2 b2 = (γ2 + iω2)b2 ig2a a iη1e b1 piη2e b3 + 2γ2b2,in, − − − − ˙ † −iθ2 iθ3 b3 = (γ3 + iω3)b3 ig3a a iη2e b2 iη3e b4 + p2γ3b3,in, − − − − ˙ † −iθ3 iθ4 b4 = (γ4 + iω4)b4 ig4a a iη3e b3 iη4e b5 + p2γ4b4,in, − − − − . . p b˙ = (γ + iω )b ig a†a iη e−iθN−2 b iη eiθN−1 b + 2γ b , N−1 − N−1 N−1 N−1 − N−1 − N−2 N−2 − N−1 N N−1 N−1,in ˙ † −iθN−1 bN = (γN + iωN )bN igN a a iηN−1e bN−1 + 2γN bN,in. p (S74) − − − To cool the mechanical resonators, we consider the strong-driving regimep of the cavity field such that the average photon number in the cavity is sufficiently large and then the linearization procedure can be used to simplify the physical model. To this end, we express the operators in Eq. (S74) as the sum of their steady-state mean values and quantum fluctuations, † † namely o = o ss +δo for operators a, a , bj=1−N , and bj . By separating the classical motion and the quantum fluctuation, the linearizedh equationsi of motion for the quantum fluctuations can be written as d δa = (κ + i∆)δa iα[g (δb + δb† )+ g (δb + δb† )+ + g (δb + δb† ) dt − − 1 1 1 2 2 2 · · · N−1 N−1 N−1 † √ +gN (δbN + δbN )] + 2κain, d δb = (γ + iω )δb ig α∗δa ig αδa† iη eiθ1 δb + 2γ b , dt 1 − 1 1 1 − 1 − 1 − 1 2 1 1,in d p δb = (γ + iω )δb ig α∗δa ig αδa† iη e−iθ1 δb iη eiθ2 δb + 2γ b , dt 2 − 2 2 2 − 2 − 2 − 1 1 − 2 3 2 2,in d p δb = (γ + iω )δb ig α∗δa ig αδa† iη e−iθ2 δb iη eiθ3 δb + 2γ b , dt 3 − 3 3 3 − 3 − 3 − 2 2 − 3 4 3 3,in d p δb = (γ + iω )δb ig α∗δa ig αδa† iη e−iθ3 δb iη eiθ4 δb + 2γ b dt 4 − 4 4 4 − 4 − 4 − 3 3 − 4 5 4 4,in . p . d δb = (γ + iω )δb ig α∗δa ig αδa† iη e−iθN−2 δb dt N−1 − N−1 N−1 N−1 − N−1 − N−1 − N−2 N−2 iη eiθN−1 δb + 2γ b , − N−1 N N−1 N−1,in d δb = (γ + iω )δb igp α∗δa ig αδa† iη e−iθN−1 δb + 2γ b . () dt N − N N N − N − N − N−1 N−1 N N,in Based on Eqs. (S75), we adopt the same procedure as that used in the two-mechanical-resonatorp case to infer a linearized optomechanical Hamiltonian governing the evolution of quantum fluctuations. For studying quantum cooling of these mechanical resonators, we focus on the beam-splitting-type interactions (i.e., the rotating-wave interaction term) between these bosonic modes because these terms dominate the linearized couplings in this system, and hence we can simplify the Hamiltonian of the system by making the RWA. The linearized optomechanical Hamiltonian under the RWA is given by

† N † N † † HI =∆δa δa + ωj j=1 δbj δbj + j=1 Gj (δa δbj + δbj δa)+ Hmrc, (S76) N ∗ where ∆=∆c + j=1 gj (βj + βj ) is theP normalized drivingP detuning after the linearization, and Gj = gj α is the linearized optomechanical coupling strength between the jth mechanical resonator and the cavity-field mode.| The| P 21

a b

6 5 4 6 5 4 H5 H4

H3

……

……

N-1 3 N-1 3 =0 =0 H N-1 H2

H1 N 1 2 N 1 2

FIG. S9. (Color online) (a) The N-mechanical-resonator optomechanical system: a cavity-field mode simultaneously couples to N mechanical resonators through the optomechanical interactions. (b) The phonon-exchange interactions between two neighboring mechanical resonators are introduced into the N-mechanical-resonator optomechanical system described by panel (a). Note that there is no direct coupling between the first resonator and the Nth resonator.

interaction Hamiltonians between the neighboring mechanical resonators are given by

N−1 Hmrc = j=1 Hj , (S77) with P

−iθj † iθj † Hj =ηj (e δbj δbj+1 + e δbj+1δbj ), (S78) which describes the phonon-exchange interaction between the jth resonator and the (j + 1)th resonator. In order to investigate the dark-mode effect in the N-mechanical-resonator optomechanical system, we firstly consider the case where the phonon-exchange interaction between the neighbouring mechanical resonators is absent, i.e., Hmrc = 0, as shown in Fig. S9(a). For convenience, we assume that all the mechanical resonators have the same resonance frequencies (ωj = ωm) and optomechanical coupling strengths (Gj = G). In this system, there exists a bright mode N B+ = j=1 δbj /√N and (N 1) dark modes which decouple from the cavity-field mode. As a result, the phonons stored in these dark modes cannot− be extracted though the optomechanical cooling channel, and then these mechanical resonatorsP cannot be cooled to their quantum ground states. Here, we can obtain the cooling limits of the N mechanical resonators, which are given by n¯(N 1)/N. The result shows that in the presence of the dark-mode effect, the final average phonon numbers in these mechanical− resonators depend on the number of the mechanical resonators. In this case, the ground-state cooling cannot be realized in these mechanical resonators. In particular, the final average phonon numbers in these mechanical resonators are approximately equal to the thermal excitations in their heat baths when N 1 and hence n¯(N 1)/N n¯. ≫ To break− the dark-mode≈ effect and realize the simultaneous ground-state cooling in the N-mechanical-resonator optomechanical system, the phase-dependent phonon-exchange interaction Hmrc should be introduced, as shown in Fig. S9(b). Without loss of generality, we assume that all the coupling strengths of the phonon-exchange interactions are same ηj = η. Thus, we can diagonalize the Hamiltonian of these coupled mechanical resonators as

N N−1 N † −iθj † iθj † † Hmrt = ωm δbj δbj + η (e δbj δbj+1 + e δbj+1δbj )= ΩkBkBk, (S79) j=1 j=1 k=1 X X X

where Bk is the kth mechanical normal mode with the resonance frequency Ωk given by kπ Ω = ω + 2η cos , k = 1, 2, 3, ..., N. (S80) k m N + 1  

The relationship between the mechanical modes δbj and the normal modes Bk is given by

1 N kπ A k=1 sin N+1 Bk, j = 1, δbj = j−1 (S81) 1 −i P θν N jkπ e P ν=1  sin Bk, j 2,  A k=1 N+1 ≥ P   22

where we introduce the variable N + 1 A = . (S82) r 2 The Hamiltonian in Eq. (S76) can be rewritten with these mechanical normal modes as † N † HI =∆δa δa + k=1 ΩkBkBk + Hom, (S83) where the optomechanical Hamiltonian Hom reads P N N − G kπ i Pj 1 θ jkπ † H = sin + e ν=1 ν sin aB + H.c.. (S84) om A N + 1 N + 1 k k=1 " j=2 # X   X   It can be seen from Eq. (S84) that the function of these phases in the optomechanical interactions is determined by the j−1 term ν=1 θν . Hence, we can apply a single phase to realize the dark-mode-breaking task. For simplicity, we assume θj = 0 for j = 2-(N 1) in the following discussions. AsP a special case, we− first analyze the case of N = 2. In this case, the multiple-mechanical-resonator optomechanical system is reduced to the two-mechanical-resonator optomechanical system, which has been analyzed before. When N = 2, the optomechanical interaction reads √2G √2G H = (1 + eiθ1 )aB† + (1 eiθ1 )aB† + H.c.. () om 2 1 2 − 2 It is obvious that when θ = nπ for an integer n, the cavity field is decoupled from one of the two hybrid mechanical modes: either B1 or B2. This hybrid mechanical mode decoupled from the cavity mode is the dark mode. However, in a general case θ = nπ, the dark-mode effect is broken, and then the ground-state cooling becomes accessible under proper parameter conditions.6 For the case of N 3, the effective coupling coefficient between the cavity-field mode a and the kth normal mode Bk in Eq. (S84) can be expressed≥ as N − G kπ i Pj 1 θ jkπ sin + e ν=1 ν sin A N + 1 N + 1 " j=2 #   X   G kπ 2kπ 3kπ = sin + eiθ1 sin + ei(θ1+θ2) sin + A N + 1 N + 1 N + 1 · · ·        − − − i PN 3 θ N 2 i PN 2 θ N 1 i PN 1 θ Nkπ +e ν=1 ν sin − kπ + e ν=1 ν sin − kπ + e ν=1 ν sin N + 1 N + 1 N + 1       − − G 1 i PN 1 θ N iθ 2 i PN 2 θ N 1 = sin kπ + e ν=1 ν sin kπ + e 1 sin kπ + e ν=1 ν sin − kπ A N + 1 N + 1 N + 1 N + 1 h     i h     i N−3 i(θ +θ ) 3 i P θν N 2 + e 1 2 sin kπ + e ν=1 sin − kπ + . () N + 1 N + 1 · · · h     i  Below, we consider two cases corresponding to odd and even numbers N, respectively. (i) For an odd number N and θ = 0 (for j = 2-(N 1)), the coefficient becomes j − N j−1 G kπ i P θν jkπ sin + e ν=1 sin A N + 1 N + 1 " j=2 #   X   G kπ Nkπ 2kπ N 1 = sin + eiθ1 sin + eiθ1 sin + sin − kπ A N + 1 N + 1 N + 1 N + 1           3kπ N 2 kπ +eiθ1 sin + sin − kπ + + eiθ1 sin . (S87) N + 1 N + 1 · · · 2         On one hand, if k is an odd number, we have N j−1 G kπ i P θν jkπ sin + e ν=1 sin A N + 1 N + 1 " j=2 #   X   G kπ 2kπ 3kπ kπ = (1 + eiθ1 )sin + 2eiθ1 sin + 2eiθ1 sin + + eiθ1 sin ; (S88) A N + 1 N + 1 N + 1 · · · 2          23

On the other hand, if k is an even number, we have

N − G kπ i Pj 1 θ jkπ G iθ kπ sin + e ν=1 ν sin = 1 e 1 sin . (S89) A N + 1 N + 1 A − N + 1 "   j=2  #   X  (ii) For an even number N and θ = 0 (for j = 2-(N 1)), the coefficient can be simplified as j − N − G kπ i Pj 1 θ jkπ sin + e ν=1 ν sin A " N + 1 j=2 N + 1 #   X   G kπ N 2kπ N 1 = sin + eiθ1 sin kπ + eiθ1 sin + sin − kπ A N + 1 N + 1 N + 1 N + 1           3kπ N 2 +eiθ1 sin + sin − kπ + . (S90) N + 1 N + 1 · · ·       In this case, when k is an odd number, we have

N − G kπ i Pj 1 θ jkπ sin + e ν=1 ν sin A N + 1 N + 1 " j=2 #   X   G kπ 2kπ 3kπ = (1 + eiθ1 )sin + 2eiθ1 sin + 2eiθ1 sin + ; (S91) A N + 1 N + 1 N + 1 · · ·         In addition, when k is an even number, we have

N − G kπ i Pj 1 θ jkπ G iθ kπ sin + e ν=1 ν sin = 1 e 1 sin . (S92) A N + 1 N + 1 A − N + 1 "   j=2  #   X  According to Eqs. (S87-S92), we can see that for odd numbers k, the coupling strength between the cavity-field mode and the kth normal mode Bk is nonzero. However, for even numbers k, the coupling strength between the cavity-field mode and the kth normal mode Bk can be expressed as G kπ H = 1 eiθ1 sin aB† + H.c., k = even number. (S93) ck A − N + 1 k     Obviously, when θ1 = 2nπ, the coupling strength between the kth mechanical normal mode (Bk=even) and the cavity mode (a) is equal to zero. In this case, all the even normal modes are decoupled from the cavity field. Then ground- state cooling cannot be realized in this system due to the dark-mode effect. Nevertheless, we can cool these mechanical resonators by choosing proper parameters to break the dark-mode effect (θ = 2nπ). 1 6

S6. GROUND-STATE COOLING OF THE MULTIPLE MECHANICAL RESONATORS

In this section, we study the simultaneous cooling of multiple mechanical resonators in the N-mechanical-resonator optomechanical system. To evaluate the cooling performance of the multiple mechanical resonators, we calculate the final average phonon numbers in these mechanical resonators. To this end, we re-express the linearized quantum Langevin equations (S75) as ˙u(t)= Au(t)+ N(t), (S94) where we introduce the vectors of the system operators

u(t) = [δa(t), δb (t), δb (t), , δb (t), δa†(t), δb† (t), δb† (t), , δb† (t)]T , (S95) 1 2 · · · N 1 2 · · · N the vector of the noise operators

† † † † T N(t)=√2[√κain(t), √γ1b1,in(t), √γ2b2,in(t), , √γN bN,in(t), √κa (t), √γ1b (t), √γ2b (t), , √γN b (t)] , · · · in 1,in 2,in · · · N,in (S96) 24

3 3 (a) Dark-mode-unbreaking (b) Dark-mode-unbreaking

2 2 Phonon number Phonon number Dark-mode-breaking Dark-mode-breaking

0.8 1 1.2 0.8 1 1.2 D/wm D/wm

f FIG. . (Color online) The final average phonon numbers nj in these mechanical resonators as functions of the effective driving detuning ∆ in the dark-mode-unbreaking case (ηj = η = 0) and the dark-mode-breaking case (ηj /ωm = η/ωm = 0.1, θ1 = π/2, and −5 3 θj=16 = 0) for N = 3 and N = 4. Here we take Gj /ωm = G/ωm = 0.1, κ/ωm = 0.2, γj /ωm = 10 , and n¯j = 10 .

and the coefficient matrix

−(κ + i∆) −iG1 −iG2 ··· −iGN 0 −iG1 −iG2 ··· −iGN ∗ iθ1 −iθN−1 −iG1 −(γ1 + iω1) −iη1e ··· −iηN−1e −iG1 0 0 · · · 0  ∗ −iθ1  −iG2 −iη1e −(γ2 + iω2) · · · 0 −iG2 0 0 · · · 0  ......   ......     −iG∗ −iη eiθN−1 0 ··· −(γ + iω ) −iG 0 0 · · · 0  A =  N N−1 N N 4  .  0 iG∗ iG∗ · · · iG∗ −(κ − i∆) iG∗ iG∗ · · · iG∗   1 2 N 1 2 N   iG∗ 0 0 · · · 0 iG −(γ − iω ) iη e−iθ1 · · · iη eiθN−1   1 1 1 1 1 N−1   iG∗ 0 0 · · · 0 iG iη eiθ1 −(γ − iω ) · · · 0   2 2 1 2 2   ......   ......   ......   ∗ −iθN−1   iGN 0 0 · · · 0 iGN iηN−1e 0 ··· −(γN − iωN )   (S97)  The formal solution of the linearized quantum Langevin equations Eq. (S94) can be obtained as

t u(t)= M(t)u(0) + M(t s)N(s)ds, (S98) − Z0 where the matrix M(t) is given by M(t)= exp(At), and hence the stability conditions derived from the Routh-Hurwitz criterion have satisfied. Note that in our simulations the real part of the eigenvalues of the coefficient matrix A is negative.

For studying the quantum cooling of these mechanical resonators, we calculate the steady-state average phonon numbers in these mechanical resonators. This can be realized by calculating the steady-state values of the covariance matrix V, which is defined by the matrix elements 1 V = [ u ( )u ( ) + u ( )u ( ) ]. (S99) ij 2 h i ∞ j ∞ i h j ∞ i ∞ i In the linearized optomechanical system, the covariance matrix V satisfies the Lyapunov equation AV + VAT = Q, (S100) − where 1 Q = (C + CT ). (S101) 2 Here C is the noise correlation matrix which is defined by the elements

N (s)N (s′) = C δ(s s′). (S102) h k l i k,l − 25

3 3 (a) (b) 2 2 Phonon numbers

1 2 3 1 2 3 4 The jth resonator The jth resonator

f FIG. S11. (Color online) The final average phonon numbers nj in these mechanical resonators are plotted in the dark-mode-unbreaking [ηj = 0 (orange bars)] and -breaking [ηj = 0.1ωm and θ1 = π/2 (blue bars)] cases for (a) N = 3 and (b) N = 4. Here ∆ = ωm, and other used parameters are the same as those given in Fig. S10.

For the Markovian baths as considered in this work, we have C(s,s′)= Cδ(s s′), where the constant matrix C is given by − 0 0 0 0 2κ 0 0 0 · · · · · · 0 0 0 0 0 2γ1 (¯n1 + 1) 0 0  0 0 0 · · · 00 0 2γ (¯n + 1) · · · 0  · · · 2 2 · · ·  ......   ...... 0     0 0 0 00 0 0 2γN (¯nN + 1)  C =  · · · · · ·  . (S103)  0 0 0 00 0 0 0   · · · · · ·   0 2γ1n¯1 0 00 0 0 0   0 0 2γ n¯ · · · 00 0 0 · · · 0   2 2   . . . ·. · · . . . . ·. · · .   ......     0 0 0 2γ n¯ 0 0 0 0   N N   · · · · · ·  Based on the covariance matrix V, the final average phonon number in the jth mechanical resonator can be obtained as 1 δb†δb = V , (S104) h j j i N+j+2,j+1 − 2

where VN+j+2,j+1 can be obtained by solving the Lyapunov equation.

3 3 θ π (a) η=0.1ωm (b) = /2

2 2 Phonon number Phonon number

0 1 2 0 0.1 0.2

/πθ η/ω m

f FIG. S12. (Color online) The final average phonon numbers nj in the mechanical resonators as functions of (a) the phase θ1 when η/ωm = 0.1 and (b) the phonon-exchange coupling η when θ1 = π/2 for N = 4. Here Gj /ωm = G/ωm = 0.1. Other used parameters are the same as those given in Fig. S10. 26

3 3 (a) (b)

2 Full 2

RWA Phonon numbers Phonon numbers

0.5 1.5 0 0.1 0.20.3 0.4 κ /wm

f f FIG. S13. (Color online) The final average phonon numbers n1 and n2 versus (a) the cavity-field decay rate κ when η = 0.05ωm and (b) the phonon-phonon coupling strength η when κ = 0.2ωm. Here the symbols and the solid curves correspond to the Hamiltonian under the RWA and the full Hamiltonian, respectively. Other parameters used are given by ω2 = ω1 = ωm, ∆ = ωm, θ = π/2, −5 3 G1 = G2 = 0.1ωm, γ1 = γ2 = 10 ωm, and n¯1 =n ¯2 = 10 .

Below we simulate the cooling performance of the mechanical resonators for the cases of N = 3 and 4. For convenience, we assume that all the mechanical resonators have the same resonance frequencies (ωj = ωm for j = 1- N), optomechanical coupling strengths (Gj = G for j = 1-N), and phonon-exchange coupling strengths [ηj = η for j = 1-(N 1)]. Moreover, we consider the case of θ1 = π/2 and θj>2 = 0. In Fig. S10, we plot the final average − f phonon numbers nj in these mechanical resonators as functions of the scaled driving detuning ∆/ωm in both the dark- mode-breaking (ηj = 0.1ωm and θ1 = π/2) and -unbreaking (ηj = η = 0) cases. The results show that the ground-state cooling is unfeasible for these mechanical resonators when the phonon-exchange interactions are absent (ηj = η = 0) [the upper curves in Figs. S10(a) and S10(b)]. This is because the phonon excitation energy stored in the dark modes cannot be extracted through the optomechanical cooling channel. When the couplings among these mechanical resonators are introduced, the dark modes are broken and then the ground-state cooling can be realized, as shown in Figs. S10(a) and S10(b). In particular, the optimal driving detuning is located at ∆ ωm, in consistent with the resolved-sideband cooling case. ≈ To see the cooling performance more clearly, we compare the cooling results of these mechanical resonators in the presence of mechanical couplings with the results corresponding to the absence of the mechanical couplings. In Fig. S11, we plot the final average phonon numbers of these mechanical resonators in the two cases. Here we can see that final average phonon numbers could be smaller than 1 when the mechanical couplings are introduced into the system, which means that the simultaneous ground-state cooling of these mechanical resonators can be achieved by breaking the dark- mode effect. We also investigate the dependence of the cooling performance on the mechanical coupling parameters η and θ. In f Fig. S12, we plot the final average phonon numbers nj in these mechanical resonators as functions of the phase θ and the scaled phonon-exchange coupling strength η/ωm. The results show that ground-state cooling can be realized for proper values of the phase θ1 = 2nπ when η = 0.1ωm [Fig. S12(a)]. In addition, the cooling efficiency of the multiple mechanical resonators can be controlled6 by tuning the phonon-exchange interaction strength η when the phase is fixed at θ1 = π/2 [Fig. S12(b)]. Under these parameters, the dark-mode effect is broken and then the thermal occupations can be extracted through the optomechanical-cooling channels.

S7. DISCUSSIONS ON THE JUSTIFICATION OF PERFORMING THE RWA

iθ † −iθ † In our model, we consider an excitation-number-conservation-type phonon-phonon interaction η(e b1b2 + e b2b1), which is obtained by making the rotating-wave approximation (RWA) in the full phonon-exchange interaction Hamiltonian iθ † −iθ † η(e b1 +e b1)(b2 +b2). To evaluate the validity of the RWA, we compare the results obtained based on the approximate Hamiltonian and the full Hamiltonian including the counterrotating term. In Figs. S13(a) and S13(b), we show the final f f average phonon numbers n1 and n2 as functions of the cavity-field decay rate κ and the mechanical coupling strength η. Here, the symbols and the solid curves correspond to the Hamiltonian under the RWA and the full Hamiltonian, respectively. Figure S13(a) shows an excellent agreement between the results obtained with the approximate Hamiltonian and the full Hamiltonian in both the resolved- and unresolved-sideband regimes. We can also see from Fig. S13(b) that 27

the approximate results match well with the exact results when η < 0.2ωm. Physically, the optomechanical cooling and heating are governed by the rotating-wave and the CR terms, respectively. In the weak-coupling regime (η ω ) and ≪ m under the near-resonance condition (ω2 around ω1), the CR term in the phonon-phonon interaction can be safely omitted by applying the RWA. The difference between these two treatments becomes non-negligible when η > 0.2ωm. The reason is that the CR term, which simultaneously creates phonon excitations in the two mechanical resonators, becomes important for a large phonon-phonon coupling strength η. These features indicate that the RWA performed in the phonon-phonon interaction is justified in our simulations, and that the CR interaction can be omitted safely under the condition η ω . ≪ m

S8. SIMULTANEOUS COOLING OF THE MECHANICAL SUPERMODES

In this section, we discuss the simultaneous cooling of the mechanical supermodes in cavity optomechanical systems. Note that the notations used in this section are independent of the those used in other sections. We consider the case where the two mechanical resonators are coupled to each other by a phonon-hopping coupling. Then, two mechanical supermodes are formed and the cavity field is coupled to the two supermodes. In the presence of the phonon-hopping coupling between the two mechanical resonators, the Hamiltonian of this coupled mechanical system reads (~ = 1) † † † † Hc = ωmc1c1 + ωmc2c2 + λ(c1c2 + c2c1), (S105) † where the operators cl=1,2 (cl ) are the annihilation (creation) operators of the lth mechanical resonator, with the corresponding resonance frequencies ωm, and the parameter λ is a coupling constant of the mechanical interaction between the two mechanical resonators. In the weak-coupling regime (λ ωm), the counter-rotating term in the phonon- phonon interaction can be safely omitted by making the rotating-wave approximation.≪ Below, we diagonalize this coupled mechanical system by introducing two mechanical supermodes C±, given by 1 C+ = (c1 + c2), (S106a) √2 1 C− = ( c1 + c2), (S106b) √2 − † † where these new operators satisfy the bosonic commutation relations [C+,C+] = 1 and [C−,C−] = 1. Thus, Hamiltonian (S105) becomes

† † Hc = ωC,+C+C+ + ωC,−C−C−, (S107) where we introduce the resonance frequencies of these supermodes as ω = ω λ. (S108) C,± m ± To cool the two mechanical supermodes, we couple the two mechanical supermodes to a common optical cavity-field mode by the optomechanical interactions [S1]. In the strong-driving regime, the linearized optomechanical Hamiltonian in the RWA takes the form as

† † † † † † † HRWA =∆δa δa + ωC,+δC+δC+ + ωC,−δC−δC− + G+(δaδC+ + δa δC+)+ G−(δaδC− + δa δC−)(S109), where ∆ is the normalized driving detuning of the cavity field, and the parameters G± are the optomechanical couplings between the cavity-field mode and the two mechanical supermodes. It can be seen from Eq. (S109) that the couplings between the cavity field and the two supermodes are the same as the three-mode optomechanical model considered in the main text. Therefore, all the analyses in the three-mode optomechanical system are suitable to the coupled cavity- supermode case. Based on the fact that the mechanical coupling between the two mechanical resonators is much smaller than the resonant frequencies of the two resonators (λ ωm), the frequencies ωC,± of the two mechanical supermodes are close to each other. We proceed to analyze the cooling≪ performance of the two mechanical supermodes. Concretely, we consider two special cases. (i) When the frequency difference between the two mechanical supermodes is larger than the effective mechanical linewidth (∆ω = ωC,+ ωC,− > Γl=+,−), the simultaneous ground-state cooling of the two mechanical supermodes is accessible under| proper− parameter| conditions. Physically, when the two mechanical supermodes are well separated in frequency, there is no dark mode, then the ground-state cooling can be realized when this system works in the resolved- sideband regime and under a proper driving (red-sideband resonance). This cooling situation is similar to the case shown in Fig. 2(b) and Fig. S1(e) [see blank area]. 28

(ii) When the frequency difference between the two mechanical supermodes is smaller than the effective mechanical linewidth (∆ω = ωC,+ ωC,− Γl=+,−), the cooling of the two mechanical supermodes is suppressed. This is because, though the| dark mode− exists| ≤ theoretically only in the degenerate-resonator case, the dark-mode effect actually works for a wider detuning range in the near-degenerate-resonatorcase. The suppression region of the ground-state for the mechanical supermodes is characterized by the effective mechanical linewidth. The cooling of the individual mechanical supermodes is suppressed in this region, i.e., the individual mechanical supermodes have significant spectral overlap and become effectively degenerate. This cooling situation is similar to the case shown in Fig. 2(b) and Fig. S1(e) [see shadow area]. In the case (ii), for achieving quantum ground-state cooling of the two mechanical supermodes, we need to introduce a phase-dependentphonon-hoppingcoupling between the two mechanical supermodes. Thus, the linearized optomechanical Hamiltonian including a phase-dependent coupling between the two supermodes takes the following form

† † † † † † † HRWA =∆δa δa + ωC,+δC+δC+ + ωC,−δC−δC− + G+(δaδC+ + δa δC+)+ G−(δaδC− + δa δC−) ˜ iφ † −iφ † +λ(e δC+δC− + e δC−δC+). (S110)

By introducing two new bosonic modes C˜+ and C˜− defined by C˜ =f˜′C eiφh˜′C , (S111a) + + − − −iφ ˜′ ˜′ C˜− =e h C+ + f C−, (S111b) Hamiltonian (S110) becomes † ˜† ˜ ˜† ˜ ˜ ˜† † ˜ ˜ ˜† † ˜ HRWA =∆δa δa +ω ˜C,+C+C+ +ω ˜C,−C−C− + G+(δaC+ + δa C+)+ G−(δaC− + δa C−), (S112)

˜′ ˜′ where we introduce the resonance frequencies ω˜C,±, the coupling strengths G˜±, and the coefficients f and h as 1 ω˜ = ω + ω + (ω ω )2 + 4λ˜2 , (S113a) C,+ 2 C,+ C,− C,+ − C,−  q  1 ω˜ = ω + ω (ω ω )2 + 4λ˜2 , (S113b) C,− 2 C,+ C,− − C,+ − C,−  q  G˜ =(f˜′G e−iφh˜′G ), (S113c) + + − − iφ ˜′ ˜′ G˜− =(e h G+ + f G−), (S113d) with ω˜ ω f˜′ = | C,− − C,+| , (S114a) (˜ω ω )2 + λ˜2 C,− − C,+ q λ˜ h˜′ = f˜′. (S114b) (˜ω ω ) C,− − C,+ We note that the cooling of the two mechanical supermodes can also be explained by the physical mechanism proposed in this manuscript. By combining this phase-dependent phonon-exchange interaction with the optomechanical couplings, the interference effect works and the dark-mode effect is broken, which can lead to the ground-state cooling of the two mechanical supermodes.

S9. PHYSICAL MECHANISM FOR BREAKING THE DARK-STATE EFFECT IN A LAMBDA-TYPE THREE-LEVEL SYSTEM

In this section, we show the physical mechanism for breaking the dark-state effect in a Lambda-type three-level system by introducing a phase-dependent transition between the two lower levels (as shown in Fig. S14). It is well known that there exists a dark state in the Lambda-type three-level system in the two-photon resonance regime. For the dark state, the superposition coefficient of the excited state is zero. Below, we show that this dark state will be broken by introducing a phase-dependent transition coupling between the two lower states. Note that in a typical natural atom, the direct transition between the two lower states of a Lambda three-level atom is forbidden due to the transition selection rule. However, this 29

E e e D 1 D2

W2 w2 W w 1 1 f E iq f Wbe wb Eg g

FIG. S14. Schematic diagram of the three-level system with these states |gi, |fi, and |ei (with the corresponding energies Eg, Ef , and Ee). A Lambda-type coupling configuration is formed by the transition processes |gi → |ei and |fi → |ei with the coupling iθ strengthes Ω1 and Ω2, and the detunings ∆1 and ∆2. A phase-dependent resonant coupling (with the coupling strength Ωbe ) between the two lower states |gi and |fi is introduced to break the dark-state effect existing in the Lambda-type three-level system working in the two-photon resonance regime ∆1 =∆2 =∆.

transition is accessible either in artificial cycle three-level systems [S2] or induced by indirect transition. The Hamiltonian of the system reads

H = E e e + E f f + E g g +Ω ( e g e−iω1 t + g e eiω1 t)+Ω ( e f e−iω2 t + f e eiω2 t) e| ih | f | ih | g| ih | 1 | ih | | ih | 2 | ih | | ih | +Ω ( f g eiθ e−iωbt + g f e−iθ eiωbt), (S115) b | ih | | ih | where Ee, Ef , and Eg are, respectively, the energies of these three energy levels e , f , and g . Two monochromatic fields with frequencies ω and ω are coupled to the atomic transitions g e| iand| if | ie (forming a Lambda 1 2 | i → | i | i → | i configuration of couplings), respectively, with Ω1 and Ω2 being the corresponding real transition amplitudes. In this system, corresponding to these two transitions g e and f e , we introduce the transition detunings as ∆ = E E ω and ∆ = E E ω|.i We → know | i that| thei Labmda-type → | i couplings support a dark state in this 1 e − g − 1 2 e − f − 2 system when the transitions satisfy the two-photon resonance condition [∆1 =∆2 in Fig. S14]. Below, we will focus on the two-photon resonant transition case, ∆1 = ∆2 = ∆. To exhibit our dark-state-breaking idea, we introduce a field to resonantly couple the two lower states f and g . In particular, this coupling has a phase-dependent coupling strength, which is the critical factor for this dark-state-breaking| i | i approach. In a rotating frame with respect to H = (E + ω ) e e + E f f + E g g , (S116) 0 g 1 | ih | f | ih | g| ih | the Hamiltonian of the system becomes V =∆ e e +Ω ( e g + g e )+Ω ( e f + f e )+Ω ( f g eiθ + g f e−iθ). (S117) I | ih | 1 | ih | | ih | 2 | ih | | ih | b | ih | | ih | By defining these three basis states with the following vectors e = (1, 0, 0)T , f = (0, 1, 0)T , g = (0, 0, 1)T , (S118) | i | i | i where “T ” denotes the matrix transpose, the interaction Hamiltonian VI can be expressed as

∆ Ω2 Ω1 iθ VI = Ω2 0 Ωbe . (S119)  −iθ  Ω1 Ωbe 0   For the sake of simplicity and without loss of generality, we consider the symmetric coupling case Ω1 =Ω2 =Ω and the single- and two-photon resonance case ∆1 =∆2 =∆=0, then the Hamiltonian (S119) becomes 0 1 1 iθ VI =Ω 1 0 ηe , (S120)  1 ηe−iθ 0    where we introduce the ratio η =Ωb/Ω. The dark-state effect can be analyzed by investigating the eigensystem of the matrix VI in Eq. (S120). The eigen- equation can be expressed as 1 V λ = λ λ , s = 1, 2, 3, (S121) Ω I | si s| si 30

0.6 (a) (b) (c)

0.52

0.6

0.50

0.4

0.48 0.4 ]

s e [ P

0.02 0.2

0.2

0 0 0

0 1 2 0 1 2 0 1 2

0.8 0.8

(d) (e) (f)

0.6

0.6 0.6

0.4 ] s 0.4 0.4

[ e P

0.2

0.2 0.2

0 0 0

0 1 2 0 1 2 0 1 2

[s] FIG. S15. The probability Pe of the excited state |ei in these eigenstates |λsi as a function of θ when (a) η = Ωb/Ω = 0.1, (b) 0.3, (c) 0.5, (d) 0.8, (e) 1.2, and (f) 1.5. Here, we can see that one of these three eigenstates has no excited state probability at θ = nπ, which means that there is a dark state at θ = nπ and hence the dark-state effect is broken when θ 6= nπ.

where λs are the eigenvalues, which are determined by the secular (cubic) equation λ3 (2 + η2)λ 2η cos θ = 0. (S122) − − Using the Cardano formula, the solutions of the cubic equation (S122) can be obtained as

1 i√3 1 i√3 λ = s + s , λ = (s + s )+ (s s ) , λ = (s + s ) (s s ) , (S123) 1 1 2 2 −2 1 2 2 1 − 2 3 −2 1 2 − 2 1 − 2 where

1 1 3 3 s = r + q3 + r2 , s = r q3 + r2 , (S124) 1 2 −  p   p  with q = (2 + η2)/3 and r = η cos θ. In general,− the form of these eigenstates defined in Eq. (S121) can be expressed as λ = c[s] g + c[s] f + c[s] e , s = 1, 2, 3. (S125) | si g | i f | i e | i The dark state can be checked by calculating the probability of the excited state e in these eigenstates as follows | i P [s] = e λ 2 = c[s] 2, s = 1, 2, 3. (S126) e |h | si| | e | The case P [s] = 0 implies a dark state of this system. In Fig. S15, we plot the probability P [s] of the excited state e in e e | i these three eigenstates λs as a function of θ when the ratio η = Ωb/Ω takes various values. Here we can see that when θ = nπ for an integer n|, onei of the eigenstates becomesa dark state. In other cases, there are no dark states. Therefore, the phase-dependent resonant transition g f can be used to break the dark-state effect in this Lambda-type three-level system. | i ↔ | i The analytical expressions of these eigenstates can be obtained as λ ηeiθ + 1 ηeiθ + λ λ = Λ g + s f + s e , s = 1, 2, 3, (S127) | si s | i λ2 1 | i λ2 1 | i  s − s −  31

where the corresponding eigenvalue λs is given by Eq. (S123), and the normalization constant is

2 (1 λs) Λs = − , s = 1, 2, 3. (S128) 4 2 2 2 2 λ + η λ + 4λsη cos θ + λ η + 2 s − s s

p When one of these eigenstates is a dark state, then the probability amplitude of the excited state e in this eigenstate (S127) is zero, and we have the relation | i

λ = ηeiθ. (S129) − By substituting the above relation into the secular equation Eq. (S122), we have

η3 ei3θ + eiθ + i2η sin θ = 0, (S130) − which leads to these two equations 

[cos θ cos (3θ)] η3 = 0, η3 [sin θ sin (3θ)] + 2η sin θ = 0. (S131) − − For a nonzero η, the solutions of these two equations are

θ = nπ, n = 0, 1, 2, . (S132) ± ± · · · When θ = nπ, we have eiθ = e−iθ = ( 1)n, then the eigenvalues of the matrix (S120) are given by − 1 1 λ = ( 1)n+1η, λ = ( 1)nη 8+ η2 , λ = ( 1)nη + 8+ η2 . (S133) 1 − 2 2 − − 3 2 − h p i h p i The corresponding eigenstates are given by

1 λ1 = ( f + g ), | i √2 −| i | i 1 λ = Λ ( 1)n+1η 8+ η2 e + f + g , | 2i 2 2 − − | i | i | i   1 h p i λ = Λ ( 1)n+1η + 8+ η2 e + f + g , (S134) | 3i 3 2 − | i | i | i   h p i 1 2 n 2 −1/2 where Λ2,3 = [ 4 (√8+ η ( 1) η) + 2] are normalization constants. In this case, the eigenstate λ1 is a dark state. ± − | i

S10. A POSSIBLE EXPERIMENTAL REALIZATION AND DERIVATION OF A PHASE-DEPENDENT PHONON-HOPPING INTERACTION BETWEEN TWO MECHANICAL RESONATORS

A. A possible experimental realization

In this section, we propose a possible experimental implementation of our scheme based on the circuit electromechanical system, as shown in Fig. S16(a). The circuit electromechanical system [S3, S4] consists of a microwave cavity described by the equivalent inductance L and capacitance C and two micromechanical resonators bj=1,2. The electromechanical coupling arises when the displacement xj=1,2 of each mechanical resonator independently modulates the total capacitance through Cj=1,2(xj ), and therefore the resonance frequency of the cavity ωc. This electromechanical coupling can be described by gj = (ωc/2C)∂Cj /∂xj . Meanwhile, an effective phase-dependent phonon-hopping interaction between the two mechanical resonators is introduced by coupling them to a superconducting charge qubit, as shown in Fig. S16(b). The detailed derivation of the phase-dependent phonon-exchange interaction is presented in the next subsection. 32

V1( t ) V2 ( t ) V3 ( t )

C1( x 1 ) C2( x 2 ) C3 f f f L 1 2 3 C C1( x 1 ) C2( x 2 ) E C J J f

(a) (b)

FIG. S16. (a) The circuit electromechanical system consists of a microwave cavity represented by an inductance L and three capacitances: C and Cj=1,2(xj). Here, the two capacitances Cj=1,2(xj) depend on the two micromechanical resonators bj=1,2. The displacement xj=1,2 of each mechanical resonator modulates the total capacitance and hence the cavity frequency ωc. A iθ † −iθ † phase-dependent phonon-hopping interaction η(e b1b2 + e b2b1) between the two micromechanical resonators is generated via a superconducting quantum circuit given in panel (b). (b) Schematic diagram of the superconducting quantum circuit: A Josephson junction with the Josephson energy EJ and the capacitance CJ is connected to three gate voltages Vj=1,2,3(t) through the corresponding gate capacitances Cj=1,2(xj) and C3. Two mechanical resonators are coupled to the superconducting charge qubit through the gate capacitances Cj=1,2(xj). The gate voltages are properly designed such that a phase-dependent phonon-hopping interaction between the two mechanical resonators can be induced. The phase drops across these capacitor Cj=1,2,3 and the Josephson junction are marked as φj and φ, respectively.

B. Derivation of a phase-dependent phonon-hopping interaction between two mechanical resonators

In this section, we present a detailed derivation of an effective phase-dependent phonon-hopping interaction between two mechanical resonators. Here the two mechanical resonators are coupled to a superconducting charge qubit, which is described by the circuit given in Fig. S16(b). In this circuit, a Josephson junction with the Josephson energy EJ and the capacitance CJ is connected to three gate voltages Vj=1,2,3(t) through the corresponding gate capacitances Cj=1,2(xj ) and C3. Here the two gate capacitors with capacitances Cj=1,2(xj ) are formed by one fixed plate and one mechanical resonator. The third capacitor has a constant capacitance. We denote the phase drops across these capacitor Cj=1,2,3 and the Josephson junction as φj and φ, respectively. In this circuit, the energy stored in these capacitors is the total kinetic energy [S5], which can be written as

1 1 1 1 T = C (x )Φ˙ 2 + C (x )Φ˙ 2 + C Φ˙ 2 + C Φ˙ 2, (S135) 2 1 1 1 2 2 2 2 2 3 3 2 J where Φj=1,2,3 and Φ are the generalized magnetic fluxes associated with the phase drops φj and φ across the capacitances Cj and the Josephson junction. The relation between the generalized magnetic flux and the phase drop is defined by φj=1,2,3 = 2πΦj /Φ0, where Φ0 is the magnetic flux quanta. The Josephson energy is identified as the potential energy, which takes the form as [S5]

2π U = E cos Φ , (S136) − J Φ  0 

where EJ is the Josephson energy of this junction. Based on these voltages relations in these loops, we have the relations

Vj (t)+ Φ˙ j + Φ˙ = 0, j = 1, 2, 3, (S137) 33

then the Lagrangian of this system can be expressed as L = T U 1 − 1 1 1 = C (x ) V 2 (t)+ C (x ) V 2 (t)+ C V 2 (t)+ (C (x )+ C (x )+ C + C ) Φ˙ 2 2 1 1 1 2 2 2 2 2 3 3 2 1 1 2 2 3 J 2π + [C (x ) V (t)+ C (x ) V (t)+ C V (t)] Φ+˙ E cos Φ . (S138) 1 1 1 2 2 2 3 3 J Φ  0  We introduce the momentum canonically conjugate to Φ as ∂L P = = [C1 (x1) V1 (t)+ C2 (x2) V2 (t)+ C3V3 (t)] + [C1 (x1)+ C2 (x2)+ C3 + CJ ] Φ˙ . (S139) ∂Φ˙ Then the Hamiltonian of this circuit can be derived as [S5] 1 4e2 2π H = [ˆn n (x ,x ,t)]2 E cos Φ 2 C (x ,x ) − g 1 2 − J Φ Σ 1 2  0  1 C (x ) V 2 (t)+ C (x ) V 2 (t)+ C V 2 (t) , (S140) −2 1 1 1 2 2 2 3 3   where we introduce the Cooper-pair number n, the gate capacitance CΣ (x1,x1), and the gate Cooper-pair number ng, which are defined by

P = 2en, CΣ (x1,x2)= C1 (x1)+ C2 (x2)+ C3 + CJ , (S141) and 1 n (x ,x ,t)= [C (x ) V (t)+ C (x ) V (t)+ C V (t)] . (S142) g 1 2 2e 1 1 1 2 2 2 3 3 The quantization of this circuit can be performed by introducing the commutative relation between the number operator nˆ and the phase operator φˆ as [φ,ˆ nˆ]= i. Then we can express the Hamiltonian in the eigen-representation of the number operator as 1 4e2 E H = [n n (x ,x ,t)]2 n n J ( n n + 1 + n + 1 n ) 2 C (x ,x ) g 1 2 2 Σ 1 2 n∈Z − | i h |− n∈Z | i h | | i h | X X 1 C (x ) V 2 (t)+ C (x ) V 2 (t)+ C V 2 (t) . (S143) −2 1 1 1 2 2 2 3 3   2 In this work, we consider the case where this circuit works in the charge qubit regime EC EJ , with EC = 4e /CΣ being the Coulomb energy. In particular, we choose the gate charge in the vicinity of 1/2, so≫ that the states 0 and 1 have almost degenerate energies. In this case, other states have higher energies and can be ignored in the our discussion| i | s.i Then the Hamiltonian becomes 2 1 4e 2 2 EJ H ng (x1,x2,t) 0 0 + [1 ng (x1,x2,t)] 1 1 ( 0 1 + 1 0 ) ≈ 2 CΣ (x1,x2) | i h | − | i h | − 2 | i h | | i h | 1 h i C (x ) V 2 (t)+ C (x ) V 2 (t)+ C V 2 (t) . (S144) −2 1 1 1 2 2 2 3 3 By introducing the Pauli operators 0 0 1 1 = σ and 0 0 + 1 1 = I, we can express the Hamiltonian as | i h | − | i h | z | i h | | i h | 1 4e2 1 E H = n (x ,x ,t) σ J σ + M, (S145) 2 C (x ,x ) g 1 2 − 2 z − 2 x Σ 1 2   where the term M stands for the ac voltage driving term on these two mechanical resonators

2 1 4e 2 1 2 2 2 M = 1 2ng (x1,x2,t) + 2ng (x1,x2,t) C1 (x1) V1 (t)+ C2 (x2) V2 (t)+ C3V3 (t(S146)) . 4 CΣ (x1,x2) − − 2     We consider the case in which the voltage drivings are far-off-resonance to these two mechanical resonators (namely the driving frequencies of the two voltages are much smaller than the resonance frequencies of the two mechanical resonators) and then the term M will be discarded in our following discussions. When the vibration amplitudes of the mechanical 34

D e

wd w b b 0 w 1 2 m ij1 ij2 g1 e g2 e g

FIG. S17. Schematic diagram of the energy levels and these involved resonance frequencies of this coupled qubit-resonator system. Two mechanical resonators with resonance frequency ωm are phase-dependently coupled to the superconducting charge qubit with the energy separation ω0. The ac gate voltages with frequency ωd are applied to the Josephson junction through the gate capacitors. resonators are much smaller than the distances between the fixed plate and the rest mechanical resonator of the capacitors, we can approximate the capacitances as x x C (x ) C 1 1 , C (x ) C 1 2 , (S147) 1 1 ≈ 10 − l 2 2 ≈ 20 − l  1   2  where Cj0 (for j = 1, 2) are the capacitances of the gate capacitors when the mechanical resonators are rest, and lj=1,2 are the rest distances between the fixed plate and the mechanical resonators in these gate capacitors. In addition, we choose the following gate voltages for our purpose,

e C10V1(t) C20V2(t) V1 (t)= V10 cos (ω1t + ϕ1) ,V2 (t)= V20 cos (ω2t + ϕ2) ,V3 (t)= − − . (S148) C3 In this case, we can obtain the relation 1 C V x C V x n (x ,x ,t) = 10 10 1 cos (ω t + ϕ )+ 20 20 2 cos (ω t + ϕ ) . (S149) g 1 2 − 2 − 2e l 1 1 2e l 2 2  1 2  By making the rotation for the qubit σ τ and σ τ , we can express the Hamiltonian upto the first order of the − x → z z → x mechanical displacements x1 and x2 as E E C V x C V x H J τ C 10 10 1 cos (ω t + ϕ )+ 20 20 2 cos (ω t + ϕ ) τ , (S150) I ≈ 2 z − 2 2e l 1 1 2e l 2 2 x  1 2  where E = 4e2/C under the approximation C (x ,x ) (C + C + C ) C . We should point out that C Σ0 Σ 1 2 ≈ 10 20 J ≡ Σ0 the mechanical displacement terms in CΣ (x1,x2) only introduce the second-order terms of xj=1,2/lj , which have been neglected in our considerations.

By including the free Hamiltonian of the two mechanicalresonators and using the relations xj=1,2 = ~/(2mωm)(bj + b†) and p = i ~mω /2(b b†), the total Hamiltonian of this circuit system becomes j j=1,2 − m j − j p p ω H ω b† b + ω b†b + 0 τ I ≈ m 1 1 m 2 2 2 z g b + b† ei(ωdt+ϕ1) + e−i(ωdt+ϕ1) + g b + b† ei(ωdt+ϕ2) + e−i(ωdt+ϕ2) (τ + τ )(S151), − 1 1 1 2 2 2 + − h        i where we consider the case ω1 = ω2 = ωd and introduce these parameters

EC C10V10 x10 EC C20V20 x20 g1 = , g2 = , ω0 = EJ , (S152) 4 2e l1 4 2e l2

with xj0 = ~/(2mωm) being the zero-point fluctuation of these mechanical resonators. To analyze the physical processes in this system, we now work in the rotating frame with respect to p ω H = ω b†b + ω b†b + 0 τ , (S153) 0 m 1 1 m 2 2 2 z 35

then the Hamiltonian becomes

V (t)= g (τ b† ei(ω0+ωm+ωd)teiϕ1 + b τ e−i(ω0+ωm+ωd)te−iϕ1 ) I − 1 + 1 1 − g (τ b† ei(ω0+ωm+ωd)teiϕ2 + b τ e−i(ω0+ωm+ωd)te−iϕ2 ) − 2 + 2 2 − g (τ b† ei(ω0+ωm−ωd)te−iϕ1 + b τ e−i(ω0+ωm−ωd)teiϕ1 ) − 1 + 1 1 − g (τ b† ei(ω0+ωm−ωd)te−iϕ2 + b τ e−i(ω0+ωm−ωd)teiϕ2 ) − 2 + 2 2 − g (τ b ei(ω0−ωm+ωd)teiϕ1 + b† τ e−i(ω0−ωm+ωd)te−iϕ1 ) − 1 + 1 1 − g (τ b ei(ω0−ωm+ωd)teiϕ2 + b† τ e−i(ω0−ωm+ωd)te−iϕ2 ) − 2 + 2 2 − g (τ b ei(ω0−ωm−ωd)te−iϕ1 + b†τ e−i(ω0−ωm−ωd)teiϕ1 ) − 1 + 1 1 − g (τ b ei(ω0−ωm−ωd)te−iϕ2 + b†τ e−i(ω0−ωm−ωd)teiϕ2 ). (S154) − 2 + 2 2 −

Here we can see that in this system there are eight physical processes, which are determined by the four detunings ω0 + ωm ωd and ω0 ωm ωd. From the viewpoint of the qubit and the resonators, the terms including ω0 + ωm ωd and ±ω ω ω− are± the counterrotating terms and the corotating terms, respectively. In this work, the motivation± 0 − m ± d for introducing the ac voltages V1(t) and V2(t) is to pick up the phase-sensitive interactions between the mechanical resonators and the charge qubit. For this purpose, we choose the ac voltages with the frequency ωd to pick up the terms with ω ω ω . Namely, we choose the parameters to satisfy the following parameter conditions 0 − m − d ω + ω ω ω ω + ω ω ω ω . (S155) 0 m ± d ≫ 0 − m d ≫ 0 − m − d

The terms with ω0 + ωm ωd and ω0 ωm + ωd are the far-off-resonance terms and the terms with ω0 ωm ωd are the target terms which work± in the large-detuning− regime. The energy levels and these involved resonance− frequencie− s of this coupled qubit-resonator system are shown in Fig. S17. In this case, the qubit-resonator interactions work in the large-detuning regime: ∆ gj=1,2√nj , where nj is the maximal excitation number involved in the jth mechanical resonator, and then we can obtain≫ a phase-dependent photon-hopping interaction between the two mechanical resonators. Here the phase is the difference between the two phases ϕ1 and ϕ2 associated with the qubit-resonator couplings. Based on the above analyses, we can obtain the approximate Hamiltonian as

V (t) τ g b e−iϕ1 + g b e−iϕ2 ei∆t + g b† eiϕ1 + g b† eiϕ2 τ e−i∆t , (S156) I ≈− + 1 1 2 2 1 1 2 2 − h    i where we introduce the detuning ∆= ω0 ωm ωd. The time factor can be eliminated by going back to the Schr¨odinger representation, in which the Hamiltonian of− the system− can be written as

ω0 ωd H = ω b†b + ω b†b + − τ τ g b e−iϕ1 + g b e−iϕ2 g b† eiϕ1 + g b†eiϕ2 τ . (S157) eff m 1 1 m 2 2 2 z − + 1 1 2 2 − 1 1 2 2 −    In this work, we consider the physical process associated with the detuning ∆ working in the large detuning case. Then we can adiabatically eliminate the qubit coherence in the physical processes and an effective phonon-phonon interaction between the two mechanical modes can be induced by the second-order perturbation. In this case, we can derive an effective Hamiltonian to describe the interactions using the method of the Frohlich-Nakajima transformation [S6, S7]. To this end, we express the effective Hamiltonian Heff as two parts ω ω H = ω b†b + ω b†b + 0 − d τ , 0 m 1 1 m 2 2 2 z H = τ g b e−iϕ1 + g b e−iϕ2 τ g b†eiϕ1 + g b†eiϕ2 . (S158) I − + 1 1 2 2 − − 1 1 2 2   We also introduce the operator  1 1 S = τ g b e−iϕ1 + g b e−iϕ2 g b† eiϕ1 + g b† eiϕ2 τ , (S159) ∆ + 1 1 2 2 − ∆ 1 1 2 2 −    which is determined by the equation

HI + [H0,S] = 0. (S160) 36

This equation means that the first-order physical process is eliminated. An effective Hamiltonian describing the second- order physical interaction can then be obtained as 1 H′ = H + [H ,S] eff 0 2 I ω ω g2 g2 (g2 + g2) = ω b†b + ω b†b + 0 − d τ + 1 τ b†b + 2 τ b†b + 1 2 τ τ m 1 1 m 2 2 2 z ∆ z 1 1 ∆ z 2 2 ∆ + − g1g2 + τ b†b ei(ϕ1−ϕ2) + b† b e−i(ϕ1−ϕ2) . (S161) ∆ z 1 2 2 1   The above Hamiltonian shows that there is no transition in the qubit states, and that a conditional phase-dependent interaction between the two mechanical resonators is introduced. We assume that the qubit is initial in its ground state g (τ g = g ), then a phase-dependent phonon-hopping interaction is obtained. | i z| i −| i

[email protected] [S1] D. Ramos, I. W. Frank, P. B. Deotare, I. Bulu, and M. Lonˇcar, Non-linear mixing in coupled photonic crystal nanobeam cavities due to cross-coupling opto-mechanical mechanisms, App. Phys. Lett. 105, 181121 (2014). [S2] Y.-x. Liu, J. Q. You, L. F. Wei, C. P. Sun, and F. Nori, Optical Selection Rules and Phase-Dependent Adiabatic State Control in a Superconducting Quantum Circuit, Phys. Rev. Lett. 95, 087001 (2005). [S3] F. Massel, T. T. Heikkil¨a, J.-M. Pirkkalainen, S. U. Cho, H. Saloniemi, P. J. Hakonen, and M. A. Sillanp¨a¨a, Microwave amplification with nanomechanical resonators, Nature (London) 480, 351 (2011). [S4] F. Massel, S. U. Cho, J.-M. Pirkkalainen, P. J. Hakonen, T. T. Heikkil¨a, and M. A. Sillanp¨a¨a, Multimode circuit optomechanics near the quantum limit, Nat. Commun. 3, 987 (2012). [S5] M. Nakahara and T. Ohmi, Quantum Computing: From Linear Algebra to Physical Realizations (CRC Press, Boca Raton, 2008). [S6] H. Fr¨ohlich, Theory of the Superconducting State. I. The Ground State at the Absolute Zero of Temperature, Phys. Rev. 79, 845 (1950). [S7] S. Nakajima, Perturbation theory in statistical mechanics, Adv. Phys. 4, 363 (1953).