Next Article in Journal
Perfect Solar Absorber Based on Four-Step Stacked Metamaterial
Previous Article in Journal
Ultra-Wideband Tunable Microwave Photonic Filter Based on Thin Film Lithium Niobate
Previous Article in Special Issue
Generation of Second-Order Sideband through Nonlinear Magnetostrictive Interaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Measurement-Based Control of Quantum Entanglement and Steering in a Distant Magnomechanical System

Department of Physics, Huazhong Normal University, Wuhan 430079, China
Photonics 2023, 10(10), 1081; https://doi.org/10.3390/photonics10101081
Submission received: 28 August 2023 / Revised: 13 September 2023 / Accepted: 18 September 2023 / Published: 26 September 2023
(This article belongs to the Special Issue Hybrid Quantum Magnonics)

Abstract

:
In this paper, we propose a scheme for measurement-based control of hybrid Einstein–Podolsky–Rosen (EPR) entanglement and steering between distant macroscopic mechanical oscillator and yttrium iron garnet (YIG) sphere in a system of an electromechanical cavity unidirectionally coupled to an electromagnonical cavity. We reveal that when the output of the electromagnonical cavity is continuously monitored by homodyne detection, not only the phonon–magnon entanglement and steering but also the purities of the phononic, magnonic and phonon–magnon states are considerably enhanced. We also find that the measurement can effectively retrieve the magnon-to-phonon steering, which is not yet obtained in the absence of the measurement. We show that unconditional phonon–magnon entanglement and steering can be achieved by introducing indirect feedback to drive the magnon and mechanical subsystems. The long-distance macroscopic hybrid entanglement and steering can be useful for, e.g., fundamental tests for quantum mechanics and quantum networks.

1. Introduction

In the past decade, great progress toward the realization of nonclassical phenomena in macroscopic mechanical oscillators has been achieved [1,2,3,4], motivated by its potential applications in the fields of probing the fundamental limits of quantum theory [5,6], ultrahigh precision measurements [7,8,9,10], and quantum information [11,12]. So far, various quantum effects [13,14,15,16,17,18,19,20,21], including mechanical squeezing [13], entanglement and Bell nonlocality [14,15,16], and phonon Fock states [20], have been demonstrated. Moreover, continuous-measurement-based control of mechanical oscillators has also been investigated experimentally [22,23,24,25,26,27], including cooling a mechanical resonator closely to ground states, observing quantum trajectory, and entropy production of a continuously monitored optomechanical oscillator. In addition, proposals for preparing strong mechanical squeezing, optomechanical entanglement and steering through continuous homodyne detection have been put forward [28,29].
In parallel with well-studied cavity optomechanical systems, macroscopic quantum effects in hybrid systems based on magnons in magnetic materials such as YIG sphere being around hundreds of micrometers in diameter increasingly have recently attracted a lot of attention, since magnons possess great frequency tunability, very low loss, and good coupling capability to other systems, i.e., microwave or optical photons, phonons, and qubits [30,31]. This therefore provides efficient ways to control quantum effects of magnons and meanwhile makes magnon-based systems very suitable for various quantum tasks. Qubit-based magnon quanta sensing, single-magnon control, and generation of a macroscopic Bell state between a single magnon and a superconducting qubit have already been reported [32,33,34]. Dynamical backaction magnomechanics and mechanical bistability have also been observed in cavity magnomechanics [35,36]. Recent proposals have been put forward for achieving magnon entanglement, magnon squeezing, magnon-mediated microwave photon entanglement, magnon blockade, magnon cat states [37,38,39,40,41,42,43,44,45,46]. In addition, magnon-based hybrid systems can also exhibit rich nonlinear phenomena, optomagnonic frequency combs [47], magnonic frequency combs in cavity magnomechanics [48], and magnon-induced high-order sideband generation [49].
In this paper, we consider a continuous measurement scheme to control photon-mediated EPR entanglement between a mechanical resonator and a distant YIG sphere. Such long-distance entanglement enables quantum communication among remote quantum nodes in quantum networks [50,51]. It should be noted that distant EPR entanglement between macroscopic mechanical and spin systems has been realized via unidirectional light coupling and measurement [52]. In addition, recent experiments have realized entanglement and quantum discord between superconducting qubits by virtue of continuous measurements, e.g., photon counting and homodyne detection [53,54]. We consider a system of an electromechanical cavity unidirectionally coupled to an electromagnonical cavity where a YIG is placed. It is found that when the output of the electromagnonical cavity is continuously monitored by homodyne detection, the phonon–magnon entanglement and steering can be obviously enhanced. We reveal that enhancement due to measurement enlarges the stability regime and also enhances the purities of the phononic, magnonic and phonon–magnon states. The measurement can also retrieve the steering from magnons to phonons which is unachievable in the absence of the measurement. We also show that by introducing indirect feedback to the mechanical and magnonic subsystems, unconditional phonon–magnon entanglement and steering can be achieved.
This paper is organized as follows. In Section 2, the system and working equations are presented. In Section 3, the effects of the continuous measurement on the phonon–magnon entanglement and steering are investigated in detail. Section 4, the indirect feedback is introduced to achieve unconditional entanglement and steering. In the last Section 5, the conclusion is given.

2. System and Equations

As shown in Figure 1, we consider that a driven electromechanical cavity is unidirectionally coupled to an electromagnonical cavity. For the electromechanical cavity, the cavity resonance is modulated by the mechanical oscillator, giving rise to the electromechanical coupling. While inside the electromagnonical cavity, a ferrimagnetic YIG sphere with a diameter of about hundreds of micrometers is placed and also biased in an uniform magnetic field. The magnons, characterizing the quanta of collective spin excitations in the YIG sphere, are coupled to the cavity mode via magnetic dipole interaction. In the rotating frame with respect to the driving frequency ω d , the system’s Hamiltonian reads ( = 1 )
H ^ = j = 1 , 2 δ j A ^ j A ^ j + δ m M ^ M ^ + ω b B ^ B ^ + g ¯ a b A ^ 1 A ^ 1 ( B ^ + B ^ ) + g a m ( A ^ 2 M ^ + A ^ 2 M ^ ) i ( E d A ^ 1 E d A ^ 1 ) ,
where the bosonic annihilation operators A ^ j , M ^ and B ^ denote, respectively, the jth cavity, magnon (magnetostatic), and mechanical modes of resonances ω c j , ω m and ω b , respectively. The detunings δ j = ω c j ω d and δ m = ω m ω d . g ¯ a b represent electromechanical coupling with a single microwave photon, and g a m denotes electromagnonical coupling, dependent on the number of spins. For the strong drive with the amplitude E d , Equation (1) can be linearized around the steady-state amplitudes by replacing the operators by A ^ j = A ^ j ss + a ^ j , B ^ = B ^ ss + b ^ , and M ^ = M ^ ss + m ^ , and the resulted linear Hamiltonian is given by
H ^ lin = Δ 1 a ^ 1 a ^ 1 + δ 2 a ^ 2 a ^ 2 + δ m m ^ m ^ + ω b b ^ b ^ + g a b ( a ^ 1 + a ^ 1 ) ( b ^ + b ^ ) + g a m ( a ^ 2 m ^ + a ^ 2 m ^ ) ,
with Δ 1 = δ 1 + 2 g ¯ a b Re [ B ^ ss ] and g a b = g ¯ a b A ^ 1 ss .
We further consider that the output field a ^ 2 out ( t ) of the second cavity a ^ 2 is subject to time-continuous homodyne detection. For the cascade two-cavity system, the input–output relation is
a ^ 2 out ( t ) = κ 2 a ^ 2 + ξ e κ 1 a ^ 1 + ξ e a 1 in ( t ) ,
where a ^ 1 in ( t ) is the input of the first cavity a ^ 1 satisfying a ^ 1 in ( t ) a ^ 1 in ( t ) = δ ( t t ) , κ j are the cavity dissipation rates, and ξ e accounts for the cascade–cavity coupling efficiency. When the generalized quadrature,
x ^ 2 θ out = ( a ^ 2 out e i θ + a ^ 2 out e i θ ) / 2 ,
is homodynely detected with the local phase θ , the detection current
I θ d t = η ( ξ e κ 1 a ^ 1 + κ 2 a ^ 2 ) e i θ + H . c . d t + d W .
Here, η denotes the detection efficiency and d W is the stochastic Wiener increment satisfying the Ito rule ( d W ) 2 = d t . Conditioned on the detection outcomes, the stochastic master equation for the density operator ρ ^ s c of the whole system is given in [55,56]:
d ρ ^ s c = i H ^ lin , ρ ^ s c d t + κ 1 L [ a ^ 1 ] ρ ^ s c d t + κ 2 L [ a ^ 2 ] ρ ^ s c d t + j = m , b κ j ( n ¯ j th + 1 ) L [ j ^ ] ρ ^ s c d t + n ¯ j th L [ j ^ ] ρ ^ s c d t ξ e κ 1 κ 2 a ^ 2 , a ^ 1 ρ ^ s c + ρ ^ s c a ^ 1 , a ^ 2 d t , + η H ξ e κ 1 a ^ 1 + κ 2 a ^ 2 e i θ ρ ^ s c d W ,
where the symbols L [ o ^ ] ρ ^ = o ^ ρ ^ o ^ 1 2 ( o ^ o ^ ρ ^ + ρ ^ o ^ o ^ ) and H [ o ^ ] ρ ^ = o ^ ρ ^ + ρ ^ o ^ o ^ + o ^ . The second and third terms in the first line in Equation (6) describe the cavity dissipation processes, the second line describes the damping of the mechanical and magnon modes at the rates κ b and κ m , in thermal environments with the mean thermal excitation numbers n ¯ j th ( e ω j / k B T j 1 ) 1 , for temperature T j and the Boltzmann constant k B ; the third line describes the cascade–cavity coupling with the efficiency ξ e ; and the last line characterizes the backaction due to the continuous measurement and it disappears when ensemble average is performed.
For initial Gaussian states, the system governed by Equation (6) evolves still in Gaussian, determined by the covariance matrix σ c , j j = μ j μ j + μ j μ j / 2 μ j μ j , where μ = ( x ^ 1 , p ^ 1 , x ^ 2 , p ^ 2 , x ^ b , p ^ b , x ^ m , p ^ m ) for the quadrature operators x ^ = ( o ^ + o ^ ) / 2 and p ^ = i ( o ^ o ^ ) / 2 . From Equation (6), we have
d μ ¯ T = A μ ¯ T d t + ( σ c C Γ ) d W ,
σ ˙ c = A σ c + σ c A T + D ( σ c C Γ ) ( σ c C Γ ) T ,
where μ ¯ j μ j . The matrix
A = A 1 0 A a b 0 A 12 A 2 0 A a m A a b 0 A b 0 0 A a m 0 A m ,
where A x = { 1 , 2 , b , m } = κ x 2 Δ x 2 Δ x κ x / 2 , with Δ 2 = δ 2 , Δ m = δ m and Δ b = ω b , A a b = 0 0 0 2 G a b , A a m = 0 g am g am 0 , and A 12 = ξ e κ 1 κ 2 I . The matrix
D = D 1 D 12 D 12 D 2 D b 0 0 D m ,
where D j = κ j I / 2 , D b , m = κ b , m ( n ¯ b , m th + 1 / 2 ) I , and D 12 = ξ e κ 1 κ 2 I / 2 . The vectors
C T = 2 η ( ξ e κ 1 cos θ , ξ e κ 1 sin θ , κ 2 cos θ , κ 2 sin θ , 0 , 0 , 0 , 0 ) ,
Γ T = η 2 ( ξ e κ 1 cos θ , ξ e κ 1 sin θ , κ 2 cos θ , κ 2 sin θ , 0 , 0 , 0 , 0 ) .
We see from Equation (2) that the first moments are related to the detection outcomes and are thus stochastic. Nevertheless, these stochastic moments are independent of the entanglement and steering for the Gaussian states and can be cancelled out by introducing feedback, as is shown later. On the contrary, the covariance matrix σ c is independent of the outcomes and thus deterministic and it completely determines quantum statistics of the system. The effect of continuous measurement is embodied by the last nonlinear term of Equation (8) (originating from the last term of Equation (6)), which is crucial for achieving strong EPR entanglement and steering for the present scheme.
By solving Equation (8), the covariance matrix σ b m of the mechanical and magnonic system can be obtained and expressed in the form σ b m = σ b σ b m σ b m T σ m . The phonon–magnon entanglement can be quantified by the logarithmic negativity [57],
E b m = max 0 , ln ( 2 e ) ,
where e = 2 1 / 2 Σ Σ 2 4 det σ b m and Σ = det σ c + det σ m 2 det c b m . The steering from phonons to magnons can be quantified by [58]
S m | b = max 0 , 1 2 ln det σ b 4 det σ b m ,
and, similarly, the reverse steering from magnons to phonons is determined by
S b | m = max 0 , 1 2 ln det σ m 4 det σ b m .

3. Conditional Entanglement and Steering via Measurement

In Figure 2, the phonon–magnon entanglement E b m in the steady-state regime is plotted for the two cases that the measurement is absent [(a1)–(c1), η = 0 ] and is present [(a2)–(c2), η = 1 ]. As shown in Figure 2(a1), we see that without the measurement, the maximal entanglement occurs for the detuning Δ m = ω b and also for the bad cavity κ 1 , 2 { ω b , g a b , g a m } [59]. This is because the phonon–magnon entanglement roots from the effective parametric amplification interaction between the mechanical and magnon modes result from the cascade photon coupling, which is resonant at this detuning; their strength is proportional to the product of cavity dissipation rates κ 1 , 2 . When the continuous monitoring is turned on, the entanglement appears over a wider range of detuning Δ m and its maximal value at Δ m = ω b is obviously enhanced for the given cavity dissipation rate κ = κ j [ 0.5 ω b , 15 ω b ] . From Figure 2(b1), we see that without the measurement, the entanglement is present at steady-state regime Δ 1 0 and becomes maximal in the vicinity of instability threshold Δ 1 = 0 of the optomechanical system. However, as depicted in Figure 2(b2), the continuous measurement enlarges the stable regime of the optomechanical system to blue-detuned regime Δ 1 < 0 , even with mediate optomechanical coupling g a b . As a matter of fact, this blue-detuned regime is conducive to the generation of photon–phonon entanglement, although it is prohibited by the stability for a generic optomechanical system. Since the phonon–magnon entanglement actually results from the distribution of the photon–phonon entanglement via the unidirectional cavity coupling, the stronger steady-state phonon–magnon entanglement is thus achievable at Δ 1 < 0 , with the help of the measurement. In addition, we can also see that apart from the blue-detuning regime, the entanglement is also obviously enhanced in red-detuned region Δ 1 > 0 . This is because the continuous measurement effectively reduces the fluctuations of the mechanical and magnon modes in this regime. Similarly, as shown in Figure 2(c1–c3), the entanglement is also improved by the measurement for different values of the couplings, g a b and g a m . We can see that in the presence of the measurement, stronger entanglement can be achieved even for smaller coupling g a m .
In Figure 3 and Figure 4, the steady-state steerings S m | b and S b | m are plotted, respectively. Comparing Figure 2 and Figure 3, we can see that the properties of steering S m | b are similar to those of entanglement E b m . On the contrary, steering S b | m displays quite different behaviors. As shown in plots (a1)–(b1) in Figure 3 and Figure 4, when the measurement is absent, the steering from magnons to phonons is unachievable, i.e., S b | m = 0 , and thus, one-way steering ( S m | b 0 ) is obtained in a wide range of parameters. This is mainly due to dissipation rate κ b κ m , which can be seen from Figure 5, where the steerings are plotted for decreased magnon damping rate κ m . It can be found that the magnon-to-phonon steering appears and the degrees of the entanglement and steerings increases with the decreasing of the magnon damping. When the measurement is present, as depicted in Figure 4(a2,b2), the magnon-to-phonon steering also appears, and therefore the two-way steering between phonons and magnons can be obtained via the continuous measurement. This is also due to the fact that continuous measurement suppresses fluctuations of the magnon mode. Similarly to phonon-to-magnon steering S m | b , the maximum of the reverse steering still occurs at Δ m = ω m . From Figure 4(c1–c3), we see that in the presence of the continuous measurement, steering S b | m can obviously be enhanced and also exist over a wider range of coupling g a m .
In Figure 6, we plot purity P b , P m , and P b m of the phononic states, magnonic states, and the phonon–magnon states for the cases that the measurement is present ( η = 1 ) and absent ( η = 0 ). It can be found that the purities are considerably enhanced by the measurement. Therefore, we can also understand that the measurement reduces the noise of the system and therefore enhances the entanglement and steering in the steady-state regime.
Figure 7 depicts the effects of imperfect unidirectional cavity coupling ( ξ e < 1 ) and thermal fluctuations on the entanglement and steering. It is shown that the entanglement and steering are robust against the unidirectional coupling loss. We see that E b m and S m | b can exist even for ξ e = 0.2 , whereas S b | m almost disappears for ξ e 0.8 . For the current experiments in the microwave domain, unidirectional cavity coupling efficiency ξ e 0.75 is achieved [53,54]. In addition, we can also see that the entanglement can still be achieved even for the temperature of up to T b 400 mK. The steerings are more influenced by the thermal noise, and steering S m | b is more robust than the reverse steering, due to damping rates κ b κ m . The two-way steering survives for T 100 mK. For the YIG sphere, the cooling to 10 mK ∼ 1 K by using a dilution refrigerator, just with small line broadening, is achieved in experiment [60]. We note that around temperature T b = T m 30 mK by using cryostat in experiments [60,61,62,63], the mean thermal magnon number n ¯ m th 0 ; it can thus be neglected, while the mean thermal phonon number n ¯ b th 62 , not near the ground states. Here, the parameters can be chosen as ω m / 2 π = 10 GHz, ω b / 2 π = 10 MHz, electromechanical coupling g a b / 2 π = 5 MHz for single-photon-mechanical coupling g ˜ a b 150 Hz and pump power P d 1 µW for Δ 1 = Δ 2 = Δ m = ω m and κ = 10 ω m . Experimentally, ultrastrong electromechanical coupling inducing the frequency splitting of 90 percent of bare 10-MHz mechanical frequency has already been achieved [61]. When considering a 400 µm diameter YIG sphere, electromagnonical coupling g a m N g m 0 = ω b can be obtained for single-spin coupling g m 0 / 2 π 38 mHz [60] and the number of spins in sphere N 7 × 10 16 with net spin density of the sphere ρ s = 2.1 × 10 27 m 3 . For the present scheme, the optomechanical device can be a three-dimensional superconducting cavity which is coupled to the motion of a micromechanical membrane. The electromagnonical device could also be a three-dimensional microwave cavity where an undoped single-crystal YIG sphere is coupled the the microwave field. The output ports of the two cavities can be connected via two microwave circulators and a coaxial cable to achieve the directional coupling from the electromechanical cavity to the electromagnonical cavity, as accepted in [53] where the second output is under continuous measurement. In the microwave domain, the cascade–cavity coupling efficiency of up to 0.8 can be achieved.

4. Unconditional Entanglement and Steering via Indirect Feedback

The above results are conditional, since expectation values μ ¯ ( t ) are driven by stochastic noise d W and therefore walk randomly in the phase space, depending on the detection outcomes. For many experimental runs of the system, the incoherent noise resulting from the random walk masks the conditional mechanical magnon EPR entanglement and steering studied in the above when ensemble average is performed. Therefore, to verify and apply the entanglement and steering, we need to convert the conditional results into the unconditional ones. This can be realized by employing state-based (indirect) feedback to remove the negative effect of stochastic expectation values μ ¯ ( t ) [55,56]. Once the measurement is performed at some time t, values x ¯ j ( t ) and p ¯ j ( t ) can be inferred immediately, based on which the Markovian feedback described by the Hamiltonian
H ^ fb = j = b , m d p j p ¯ j ( t ) x ^ j d x j x ¯ j ( t ) p ^ j
can be constructed, with the feedback gain parameters d x , p j . The feedback leads Equation (7) to be modified by substituting A with A ˜ A + d i a g ( 0 , 0 , 0 , 0 , d b x , d b y , d m x , d m y ) .
For the feedback described by the Hamiltonian of Equation (16), the mechanical driving can be realized by, e.g., electric actuation, as demonstrated in [64,65,66]. To achieve the magnon damping force, a drive tone at frequency ω d and supplied by a microwave source can directly drive the YIG sphere, as performed in [63]. Here, the driving magnetic field, bias magnetic field H B , and the magnetic field of the microwave cavity are orthogonal to each other at the site of the YIG sphere to avoid the mutual impact among them.
Then, ensemble average σ ¯ e 1 2 μ ¯ i ( t ) μ ¯ j ( t ) + μ ¯ j ( t ) μ ¯ i ( t ) e over many realizations of the system can be derived as
σ ¯ ˙ e = A ˜ σ ¯ e + σ ¯ e A ˜ T + ( σ c C Γ ) ( σ c C Γ ) T ,
as well as ensemble average correlation matrix
σ e = σ c + σ ¯ e
determining the system’s properties under the feedback. When σ ¯ e 0 through the feedback, the covariance matrix
σ e σ c ,
independent of the measurement outcomes and deterministic. The overlap between the states with the covariance matrices σ c and σ e can be quantified by fidelity [67]
F σ c , σ e = Θ + Λ ( Θ + Λ ) 2 Δ 1 ,
where Θ = 16 D e t [ J σ c J σ e + I / 4 ] , Λ = 16 D e t [ σ c + i J / 4 ] D e t [ σ e + i J / 4 ] , Δ = D e t [ σ c + σ e ] , and J = 0 1 1 0 .
In Figure 8a,b, fidelity F σ c , σ e , entanglement E b m and steering S m | b and S b | m are plotted for the two cases: d x j = d p j = d ( j = b , m ) and d x b = 0 and d p b = d x m = d p m = d . We see that in both cases, the fidelity, the entanglement and steering increase as feedback strength d arises. This is because the increase in the feedback strength leads to stronger damping for mean values x ^ j and p ^ j , which in turn further suppress the fluctuations (i.e., σ ¯ e ) of the mean values and even almost removes them completely in the limit of strong feedback. We see that in this limit, fidelity F σ c , σ e 1 and the entanglement and steeering recover to the conditional values, E b m 0.84 , S m | b 0.39 , and S b | m 0.09 in Figure 3 when d x j = d p j = d and d ω b . It is also explicitly shown that a greater amount of unconditional entanglement and steerings can be recovered when simultaneously driving mechanical position x ^ b and momentum p ^ b in the feedback compared to when just driving the mechanical position alone. In addition, as shown in Figure 8c,d, the effect of the unequal feedback parameters is taken into account. We see that the unconditional entanglement and steering become maximal approximately near the symmetric feedback and they decrease when d m exceeds d b . For the asymmetric feedback, one-way steering also appears, showing that phonon–magnon entanglement and steering can be controlled via feedback.

5. Conclusions

In conclusion, in this paper, we consider a continuous measurement scheme to enhance photon-mediated EPR entanglement and steering between a mechanical resonator and a distant YIG sphere, a system of an electromechanical cavity unidirectionally coupled to an electromagnonical cavity where a YIG is placed. We reveal that when the output of the electromagnonical cavity is continuously monitored by homodyne detection, the phonon–magnon entanglement and steering can be obviously enhanced. It is also revealed that the enhancement due to the measurement enlarges the stability regime and also enhances the purities of the phononic, magnonic and phonon–magnon states. It is found that continuous measurement can retrieve the steering from magnons to phonons, which is unachievable in the absence of the measurement. We finally propose an indirect feedback scheme to achieve unconditional phonon–magnon entanglement and steering. The long-distance macroscopic hybrid entanglement and steering can be useful for, e.g., fundamental tests for quantum mechanics and quantum networks.

Funding

This work is supported by the National Natural Science Foundation of China (No.12174140).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Meystre, P. A short walk through quantum optomechanics. Ann. Phys. 2013, 525, 215. [Google Scholar] [CrossRef]
  2. Poot, M.; van der Zant, H.S.J. Mechanical systems in the quantum regime. Phys. Rep. 2012, 511, 273. [Google Scholar] [CrossRef]
  3. Aspelmeyer, M.; Kippenberg, T.J.; Marquardt, F. Cavity optomechanics. Rev. Mod. Phys. 2014, 86, 1391. [Google Scholar] [CrossRef]
  4. Barzanjeh, S.; Xuereb, A.; Gröblacher, S.; Paternostro, M.; Regal, C.A.; Weig, E. Optomechanics for quantum technologies. Nat. Phys. 2022, 18, 15. [Google Scholar] [CrossRef]
  5. Zurek, W.H. Decoherence and the Transition from Quantum to Classical. Phys. Today 1991, 44, 36. [Google Scholar] [CrossRef]
  6. Zurek, W.H. Decoherence, einselection, and the quantum origins of the classical. Rev. Mod. Phys. 2003, 75, 715. [Google Scholar] [CrossRef]
  7. Caves, C.M.; Thorne, K.S.; Drever, R.W.P.; Sandberg, V.D.; Zimmermann, M. On the measurement of a weak classical force coupled to a quantum-mechanical oscillator. I. Issues of principle. Rev. Mod. Phys. 1980, 52, 341. [Google Scholar] [CrossRef]
  8. Metcalfe, M. Applications of cavity optomechanics. Appl. Phys. Rev. 2014, 1, 031105. [Google Scholar] [CrossRef]
  9. Li, B.-B.; Ou, L.; Lei, Y.; Liu, Y.-C. Cavity optomechanical sensing. Nanophotonics 2021, 10, 2799. [Google Scholar] [CrossRef]
  10. Motazedifard, A.; Dalafi, A.; Naderi, M.H. Ultraprecision quantum sensing and measurement based on nonlinear hybrid optomechanical systems containing ultracold atoms or atomic Bose-Einstein condensate. AVS Quantum Sci. 2021, 3, 024701. [Google Scholar] [CrossRef]
  11. Rips, S.; Hartmann, M.J. Quantum Information Processing with Nanomechanical Qubits. Phys. Rev. Lett. 2013, 110, 120503. [Google Scholar] [CrossRef] [PubMed]
  12. Muschik, C.A.; Krauter, H.; Hammerer, K.; Polzik, E.S. Quantum information at the interface of light with atomic ensembles and micromechanical oscillators. Quantum Inf. Process. 2011, 10, 839. [Google Scholar] [CrossRef]
  13. Wollman, E.E.; Lei, C.U.; Weinstein, A.J.; Suh, J.; Kronwald, A.; Marquardt, F.; Clerk, A.A.; Schwab, K.C. QUANTUM MECHANICS. Quantum squeezing of motion in a mechanical resonator. Science 2015, 349, 952. [Google Scholar] [CrossRef] [PubMed]
  14. Pirkkalainen, J.M.; Damskägg, E.; Brandt, M.; Massel, F.; Sillanpää, M.A. Squeezing of Quantum Noise of Motion in a Micromechanical Resonator. Phys. Rev. Lett. 2015, 115, 243601. [Google Scholar] [CrossRef]
  15. Ockeloen-Korppi, C.F.; Damskägg, E.; Pirkkalainen, J.M.; Asjad, M.; Clerk, A.A.; Massel, F.; Woolley, M.J.; Sillanpää, M.A. Stabilized entanglement of massive mechanical oscillators. Nature 2018, 556, 478. [Google Scholar] [CrossRef]
  16. Kotler, S.; Peterson, G.A.; Shojaee, E.; Lecocq, F.; Cicak, K.; Kwiatkowski, A.; Geller, S.; Glancy, S.; Knill, E.; Simmonds, R.W.; et al. Direct observation of deterministic macroscopic entanglement. Science 2021, 372, 622. [Google Scholar] [CrossRef]
  17. Palomaki, T.A.; Teufel, J.D.; Simmonds, R.W.; Lehnert, K.W. Entangling mechanical motion with microwave fields. Science 2013, 342, 710. [Google Scholar] [CrossRef]
  18. Ho, M.; Oudot, E.; Bancal, J.; Sangouard, N. Witnessing Optomechanical Entanglement with Photon Counting. Phys. Rev. Lett. 2018, 121, 023602. [Google Scholar] [CrossRef]
  19. Riedinger, R.; Hong, S.; Norte, R.A.; Slater, J.A.; Shang, J.; Krause, A.G.; Anant, V.; Aspelmeyer, M.; Gröblacher, S. Non-classical correlations between single photons and phonons from a mechanical oscillator. Nature 2016, 530, 313. [Google Scholar] [CrossRef]
  20. Hong, S.; Riedinger, R.; Marinkovic, I.; Wallucks, A.; Hofer, S.G.; Norte, R.A.; Aspelmeyer, M.; Gröblacher, S. Hanbury Brown and Twiss interferometry of single phonons from an optomechanical resonator. Science 2017, 358, 203–206. [Google Scholar] [CrossRef]
  21. Marinkovic, I.; Wallucks, A.; Riedinger, R.; Hong, S.; Aspelmeyer, M.; Gröblacher, S. An optomechanical Bell test. arXiv 2018, arXiv:1806.10615. [Google Scholar] [CrossRef] [PubMed]
  22. Rossi, M.; Mason, D.; Chen, J.; Tsaturyan, Y.; Schliesser, A. Measurement-based quantum control of mechanical motion. Nature 2018, 563, 53. [Google Scholar] [CrossRef] [PubMed]
  23. Rossi, M.; Mason, D.; Chen, J.; Schliesser, A. Observing and Verifying the Quantum Trajectory of a Mechanical Resonator. Phys. Rev. Lett. 2019, 123, 163601. [Google Scholar] [CrossRef] [PubMed]
  24. Mason, D.; Chen, J.; Rossi, M.; Tsaturyan, Y.; Schliesser, A. Continuous force and displacement measurement below the standard quantum limit. Nat. Phys. 2019, 15, 745. [Google Scholar] [CrossRef]
  25. Rossi, M.; Mancino, L.; Landi, G.T.; Paternostro, M.; Schliesser, A.; Belenchia, A. Experimental Assessment of Entropy Production in a Continuously Measured Mechanical Resonator. Phys. Rev. Lett. 2020, 125, 080601. [Google Scholar] [CrossRef]
  26. Tebbenjohanns, F.; Mattana, M.L.; Rossi, M.; Frimmer, M.; Novotny, L. Quantum control of a nanoparticle optically levitated in cryogenic free space. Nature 2021, 595, 378. [Google Scholar] [CrossRef]
  27. Magrini, L.; Rosenzweig, P.; Bach, C.; Deutschmann-Olek, A.; Hofer, S.G.; Hong, S.; Kiesel, N.; Kugi, A.; Aspelmeyer, M. Real-time optimal quantum control of mechanical motion at room temperature. Nature 2021, 595, 373. [Google Scholar] [CrossRef]
  28. Tan, H.; Zhan, H. Strong mechanical squeezing and optomechanical steering via continuous monitoring in optomechanical systems. Phys. Rev. A 2019, 100, 023843. [Google Scholar] [CrossRef]
  29. Brunelli, M.; Malz, D.; Nunnenkamp, A. Conditional Dynamics of Optomechanical Two-Tone Backaction-Evading Measurements. Phys. Rev. Lett. 2019, 123, 093602. [Google Scholar] [CrossRef]
  30. Quirion, D.L.; Tabuchi, Y.; Gloppe, A.; Usami, K.; Nakamura, Y. Hybrid quantum systems based on magnonics. Appl. Phys. Express 2019, 12, 070101. [Google Scholar] [CrossRef]
  31. Yuan, H.Y.; Cao, Y.; Kamra, A.; Duine, R.A.; Yan, P. Quantum magnonics: When magnon spintronics meets quantum information science. Phys. Rep. 2022, 965, 1. [Google Scholar] [CrossRef]
  32. Lachance-Quirion, D.; Piotr Wolski, S.; Tabuchi, Y.; Kono, S.; Usami, K.; Nakamura, Y. Entanglement-based single-shot detection of a single magnon with a superconducting qubit. Science 2020, 367, 425. [Google Scholar] [CrossRef]
  33. Xu, D.; Gu, X.-K.; Li, H.-K.; Weng, Y.-C.; Wang, Y.-P.; Li, J.; Wang, H.; Zhu, S.-Y.; You, J.Q. Quantum Control of a Single Magnon in a Macroscopic Spin System. Phys. Rev. Lett. 2023, 130, 193603. [Google Scholar] [CrossRef]
  34. Xu, D.; Gu, X.-K.; Weng, Y.-C.; Li, H.-K.; Wang, Y.-P.; Zhu, S.-Y.; You, J.Q. Deterministic generation and tomography of a macroscopic Bell state between a millimeter-sized spin system and a superconducting qubit. arXiv 2023, arXiv:2306.09677. [Google Scholar]
  35. Potts, C.A.; Varga, E.; Bittencourt, V.A.S.V.; Viola Kusminskiy, S.; Davis, J.P. Dynamical Backaction Magnomechanics. Phys. Rev. X 2021, 11, 031053. [Google Scholar] [CrossRef]
  36. Shen, R.-C.; Li, J.; Fan, Z.-Y.; Wang, Y.-P.; You, J.Q. Mechanical Bistability in Kerr-modified Cavity Magnomechanics. Phys. Rev. Lett. 2022, 129, 123601. [Google Scholar] [CrossRef]
  37. Zhang, Z.; Scully, M.O.; Agarwal, G.S. Quantum entanglement between two magnon modes via Kerr nonlinearity driven far from equilibrium. Phys. Rev. Res. 2019, 1, 023021. [Google Scholar] [CrossRef]
  38. Zhan, H.; Sun, L.; Tan, H. Chirality-induced one-way quantum steering between two waveguide-mediated ferrimagnetic microspheres. Phys. Rev. B 2022, 106, 104432. [Google Scholar] [CrossRef]
  39. Ren, Y.; Xie, J.; Li, X.; Ma, S.; Li, F. Long-range generation of a magnon-magnon entangled state. Phys. Rev. B 2022, 105, 094422. [Google Scholar] [CrossRef]
  40. Liu, Z.-X.; Xiong, H.; Wu, Y. Magnon blockade in a hybrid ferromagnet-superconductor quantum system. Phys. Rev. B 2019, 100, 134421. [Google Scholar] [CrossRef]
  41. Xie, H.; He, L.-W.; Shang, X.; Lin, G.-W.; Lin, X.-M. Nonreciprocal photon blockade in cavity optomagnonics. Phys. Rev. A 2022, 106, 053707. [Google Scholar] [CrossRef]
  42. Yuan, H.Y.; Zheng, S.; Ficek, Z.; He, Q.Y.; Yung, M.-H. Steady Bell State Generation via Magnon-Photon Coupling. Phys. Rev. Lett. 2020, 124, 053602. [Google Scholar] [CrossRef] [PubMed]
  43. Li, J.; Zhu, S.-Y.; Agarwal, G.S. Magnon-Photon-Phonon Entanglement in Cavity Magnomechanics. Phys. Rev. Lett. 2018, 121, 203601. [Google Scholar] [CrossRef] [PubMed]
  44. Li, J.; Zhu, S.-Y.; Agarwal, G.S. Squeezed states of magnons and phonons in cavity magnomechanics. Phys. Rev. A 2019, 99, 021801. [Google Scholar] [CrossRef]
  45. Yu, M.; Shen, H.; Li, J. Magnetostrictively Induced Stationary Entanglement between Two Microwave Fields. Phys. Rev. Lett. 2020, 124, 213604. [Google Scholar] [CrossRef]
  46. Sun, F.-X.; Zheng, S.-S.; Xiao, Y.; Gong, Q.; He, Q.; Xia, K. Remote Generation of Magnon Schrödinger Cat State via Magnon-Photon Entanglement. Phys. Rev. Lett. 2021, 127, 087203. [Google Scholar] [CrossRef]
  47. Liu, Z.-X.; Li, Y.-Q. Optomagnonic frequency combs. Photon. Res. 2022, 12, 2786. [Google Scholar] [CrossRef]
  48. Xiong, H. Magnonic frequency combs based on the resonantly enhanced magnetostrictive effect. Fund. Res. 2023, 3, 8. [Google Scholar] [CrossRef]
  49. Liu, Z.-X.; Wang, B.; Xiong, H.; Wu, Y. Magnon-induced high-order sideband generation. Opt. Lett. 2018, 43, 3698. [Google Scholar] [CrossRef]
  50. Kimble, H.J. The quantum internet. Nature 2008, 453, 1023. [Google Scholar] [CrossRef]
  51. Simon, C. Towards a global quantum network. Nat. Photon. 2017, 11, 678. [Google Scholar] [CrossRef]
  52. Thomas, R.A.; Parniak, M.; ØStfeldt, C.; Moller, C.B.; Bærentsen, C.; Tsaturyan, Y.; Schliesser, A.; Appel, J.; Zeuthen, E.; Polzik, E.S. Entanglement between distant macroscopic mechanical and spin systems. Nat. Phys. 2021, 17, 228. [Google Scholar] [CrossRef]
  53. Roch, N.; Schwartz, M.E.; Motzoi, F.; Macklin, C.; Vijay, R.; Eddins, A.W.; Korotkov, A.N.; Whaley, K.B.; Sarovar, M.; Siddiqi, I. Observation of Measurement-Induced Entanglement and Quantum Trajectories of Remote Superconducting Qubits. Phys. Rev. Lett. 2014, 112, 170501. [Google Scholar] [CrossRef] [PubMed]
  54. Chantasri, A.; Kimchi-Schwartz, M.E.; Roch, N.; Siddiqi, I.; Jordan, A.N. Quantum Trajectories and Their Statistics for Remotely Entangled Quantum Bits. Phys. Rev. X 2016, 6, 041052. [Google Scholar] [CrossRef]
  55. Wiseman, H.M.; Milburn, G.J. Quantum Measurement and Control; Cambridge University Press: Cambridge, UK, 2010. [Google Scholar]
  56. Zhang, J.; Liu, Y.-X.; Wu, R.-B.; Jacobs, K.; Nori, F. Quantum feedback: Theory, experiments, and applications. Phys. Rep. 2017, 679, 1. [Google Scholar] [CrossRef]
  57. Plenio, M.B. Logarithmic Negativity: A Full Entanglement Monotone That is not Convex. Phys. Rev. Lett. 2005, 95, 090503. [Google Scholar] [CrossRef] [PubMed]
  58. Kogias, I.; Lee, A.R.; Ragy, S.; Adesso, G. Quantification of Gaussian Quantum Steering. Phys. Rev. Lett. 2015, 114, 060403. [Google Scholar] [CrossRef]
  59. Tan, H.; Li, J. Einstein-Podolsky-Rosen entanglement and asymmetric steering between distant macroscopic mechanical and magnonic systems. Phys. Rev. Res. 2021, 3, 013192. [Google Scholar] [CrossRef]
  60. Tabuchi, Y.; Ishino, S.; Ishikawa, T.; Yamazaki, R.; Usami, K.; Nakamura, Y. Hybridizing Ferromagnetic Magnons and Microwave Photons in the Quantum Limit. Phys. Rev. Lett. 2014, 113, 083603. [Google Scholar] [CrossRef]
  61. Peterson, G.A.; Kotler, S.; Lecocq, F.; Cicak, K.; Jin, X.Y.; Simmonds, R.W.; Aumentado, J.; Teufel, J.D. Ultrastrong Parametric Coupling between a Superconducting Cavity and a Mechanical Resonator. Phys. Rev. Lett. 2019, 123, 247701. [Google Scholar] [CrossRef]
  62. Zhang, D.; Luo, X.-Q.; Wang, Y.-P.; Li, T.-F.; You, J.Q. Observation of the exceptional point in cavity magnon-polaritons. Nat. Commun. 2017, 8, 1368. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, Y.-P.; Zhang, G.-Q.; Zhang, D.; Li, T.-F.; Hu, C.-M.; You, J.Q. Bistability of Cavity Magnon Polaritons. Phys. Rev. Lett. 2018, 120, 057202. [Google Scholar] [CrossRef] [PubMed]
  64. Poggio, M.; Degen, C.L.; Mamin, H.J.; Rugar, D. Feedback Cooling of a Cantilever’s Fundamental Mode below 5 mK. Phys. Rev. Lett. 2007, 99, 017201. [Google Scholar] [CrossRef] [PubMed]
  65. Schäfermeier, C.; Kerdoncuff, H.; Hoff, U.B.; Fu, H.; Huck, A.; Bilek, J.; Harris, G.I.; Bowen, W.P.; Gehring, T.; Andersen, U.L. Quantum enhanced feedback cooling of a mechanical oscillator using nonclassical light. Nat. Commun. 2016, 7, 13628. [Google Scholar] [CrossRef] [PubMed]
  66. Fan, L.; Fong, K.Y.; Poot, M.; Tang, H.X. Cascaded optical transparency in multimode-cavity optomechanical systems. Nat. Commun. 2015, 6, 5850. [Google Scholar] [CrossRef]
  67. Banchi, L.; Braunstein, S.L.; Pirandola, S. Quantum Fidelity for Arbitrary Gaussian States. Phys. Rev. Lett. 2015, 115, 260501. [Google Scholar] [CrossRef]
Figure 1. The schematic plot of a driven electromechanical cavity is unidirectionally coupled to an electromagnonical cavity whose output is continuously monitored by homodyne detection. Based on the detection outcome I ( t ) , indirect (state-based) feedback with gains d b , m is employed to achieve unconditional entanglement and steering between the macroscopic mechanical oscillator and the YIG sphere.
Figure 1. The schematic plot of a driven electromechanical cavity is unidirectionally coupled to an electromagnonical cavity whose output is continuously monitored by homodyne detection. Based on the detection outcome I ( t ) , indirect (state-based) feedback with gains d b , m is employed to achieve unconditional entanglement and steering between the macroscopic mechanical oscillator and the YIG sphere.
Photonics 10 01081 g001
Figure 2. The dependence of the phonon–magnon entanglement E b m on different parameters for the measurement being absent ( η = 0 ) in (a1,b1,c1) and present ( η = 1 ) in (a2,b2,c2). In (a1,a2), Δ 1 = Δ 2 = ω b , g a b = 0.5 g a m = 0.5 ω b ; In (b1,b2), κ = 10 ω b , Δ m = ω b , g a b = 0.5 g a m = 0.5 ω b ; In (c1,c2), κ = 10 ω b , Δ m = ω b , Δ 1 = Δ 2 = ω b . The other parameters ω b / 2 π = 10 MHz, ω m / 2 π = 10 GHz, κ m = 0.15 ω b , κ b = 10 5 ω b , T 1 = T m = T b = 30 mK, and ξ e = 1 . The plots (a3,b2,c3) are the same as (a1,b1,c1), respectively, with Δ m = ω m [in (a3)], Δ 2 = ω m [in (b3)], and g a b = ω m [in (c3)].
Figure 2. The dependence of the phonon–magnon entanglement E b m on different parameters for the measurement being absent ( η = 0 ) in (a1,b1,c1) and present ( η = 1 ) in (a2,b2,c2). In (a1,a2), Δ 1 = Δ 2 = ω b , g a b = 0.5 g a m = 0.5 ω b ; In (b1,b2), κ = 10 ω b , Δ m = ω b , g a b = 0.5 g a m = 0.5 ω b ; In (c1,c2), κ = 10 ω b , Δ m = ω b , Δ 1 = Δ 2 = ω b . The other parameters ω b / 2 π = 10 MHz, ω m / 2 π = 10 GHz, κ m = 0.15 ω b , κ b = 10 5 ω b , T 1 = T m = T b = 30 mK, and ξ e = 1 . The plots (a3,b2,c3) are the same as (a1,b1,c1), respectively, with Δ m = ω m [in (a3)], Δ 2 = ω m [in (b3)], and g a b = ω m [in (c3)].
Photonics 10 01081 g002
Figure 3. The dependence of phonon-to-magnon steering S m | b on different parameters for the measurement being absent ( η = 0 ) (in (a1,b1,c1)) and present ( η = 1 ) (in (a2,b2,c2)). The other parameters are the same as in Figure 2.
Figure 3. The dependence of phonon-to-magnon steering S m | b on different parameters for the measurement being absent ( η = 0 ) (in (a1,b1,c1)) and present ( η = 1 ) (in (a2,b2,c2)). The other parameters are the same as in Figure 2.
Photonics 10 01081 g003
Figure 4. The dependence of magnon-to-phonon steering S b | m on different parameters for the measurement being absent ( η = 0 ) (in (a1,b1,c1)) and present ( η = 1 ) (in (a2,b2,c2)). The other parameters are the same as in Figure 2.
Figure 4. The dependence of magnon-to-phonon steering S b | m on different parameters for the measurement being absent ( η = 0 ) (in (a1,b1,c1)) and present ( η = 1 ) (in (a2,b2,c2)). The other parameters are the same as in Figure 2.
Photonics 10 01081 g004
Figure 5. Purity P b , P m , and P b m of the phononic states, magnonic states, and phonon–magnon states in (a,b,c) for the same parameters as those in Figure 2(a3,b3,c3), for the measurement being present ( η = 1 , thick lines) and absent ( η = 0 , thin lines).
Figure 5. Purity P b , P m , and P b m of the phononic states, magnonic states, and phonon–magnon states in (a,b,c) for the same parameters as those in Figure 2(a3,b3,c3), for the measurement being present ( η = 1 , thick lines) and absent ( η = 0 , thin lines).
Photonics 10 01081 g005
Figure 6. Density plots of steerings S m | b (a1c1) and S b | m (a2c2) in the absence of the measurement for the same parameters as (a1c1) in Figure 3 and (a1c1) in Figure 4, respectively, except for κ m = 0.01 ω b .
Figure 6. Density plots of steerings S m | b (a1c1) and S b | m (a2c2) in the absence of the measurement for the same parameters as (a1c1) in Figure 3 and (a1c1) in Figure 4, respectively, except for κ m = 0.01 ω b .
Photonics 10 01081 g006
Figure 7. Density plot of entanglement E b m and steering S m | b and S b | m versus temperature T and unidirectional coupling efficiency ξ e for g a b = 0.5 ω b and Δ 1 = ω b in (a1,b1,c1), g a b = ω b and Δ 1 = ω b in (a2,b2,c2), g a m = ω b , κ = 10 ω b , Δ 2 = ω b , Δ m = ω b , and the other parameters are the same as in Figure 2.
Figure 7. Density plot of entanglement E b m and steering S m | b and S b | m versus temperature T and unidirectional coupling efficiency ξ e for g a b = 0.5 ω b and Δ 1 = ω b in (a1,b1,c1), g a b = ω b and Δ 1 = ω b in (a2,b2,c2), g a m = ω b , κ = 10 ω b , Δ 2 = ω b , Δ m = ω b , and the other parameters are the same as in Figure 2.
Photonics 10 01081 g007
Figure 8. (a,b) The dependence of unconditional entanglement E b m , fidelity F σ c , σ e and steerings S m | b and S b | m on feedback parameter d, respectively, for d x j = d p j = d ( j = b , m ) (thick lines) and d x b = 0 and d p b = d x m = d p m = d (thin lines). (c,d) The effect of unequal feedback parameters d b and d m on entanglement and steerings for d b = 0.5 ω b (thin lines) and d b = ω b (thick lines). We have Δ 1 = Δ 2 = ω b , g a b = 0.5 g a m = 0.5 ω b , κ = 10 ω b , Δ m = ω b , η = 1 , ξ e = 1 , and the other parameters are the same as in Figure 2.
Figure 8. (a,b) The dependence of unconditional entanglement E b m , fidelity F σ c , σ e and steerings S m | b and S b | m on feedback parameter d, respectively, for d x j = d p j = d ( j = b , m ) (thick lines) and d x b = 0 and d p b = d x m = d p m = d (thin lines). (c,d) The effect of unequal feedback parameters d b and d m on entanglement and steerings for d b = 0.5 ω b (thin lines) and d b = ω b (thick lines). We have Δ 1 = Δ 2 = ω b , g a b = 0.5 g a m = 0.5 ω b , κ = 10 ω b , Δ m = ω b , η = 1 , ξ e = 1 , and the other parameters are the same as in Figure 2.
Photonics 10 01081 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tan, H. Measurement-Based Control of Quantum Entanglement and Steering in a Distant Magnomechanical System. Photonics 2023, 10, 1081. https://doi.org/10.3390/photonics10101081

AMA Style

Tan H. Measurement-Based Control of Quantum Entanglement and Steering in a Distant Magnomechanical System. Photonics. 2023; 10(10):1081. https://doi.org/10.3390/photonics10101081

Chicago/Turabian Style

Tan, Huatang. 2023. "Measurement-Based Control of Quantum Entanglement and Steering in a Distant Magnomechanical System" Photonics 10, no. 10: 1081. https://doi.org/10.3390/photonics10101081

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop