1. Introduction
One of the most significant open problems in quantum optics is to prepare entangled states, which have potential applications in quantum sensing and networking [
1]. Entangled states have been experimentally created in several microscopic systems, such as atoms [
2], photons [
3], and ions [
4]. However, the creation of entangled states at the macroscopic level is still challenging. In the past two decades, optomechanical systems, in which a cavity field is coupled with a macroscopic mechanical oscillator via linear momentum transfer, have been proven to be a promising platform on which to prepare macroscopic entanglement [
5] because the linear vibration of the mechanical oscillator can, experimentally, be cooled down close to its quantum ground state [
6,
7]. Meanwhile, Laguerre–Gaussian-(LG)-cavity optomechanical systems have been intensively studied [
8,
9,
10,
11,
12,
13,
14,
15,
16,
17,
18,
19,
20,
21]. In these, an LG-cavity mode is coupled to a rotating mirror due to the orbital angular momentum transfer from the LG-cavity mode to the rotating mirror. It has been shown that it is also possible to cool down the rotational motion of a macroscopic mirror close to its quantum ground state [
8]. Thus, it is possible to observe a variety of nonlinear and quantum phenomena in such systems, including optomechanically induced transparency [
9,
10,
11,
12], optomechanically induced amplification [
13], Fano resonance [
14], fast and slow light [
11,
14], optical sum-sideband generation [
15], second-order sideband effects [
16], entanglement between an LG-cavity mode and a rotating mirror [
17,
18], stationary entanglement between two rotating mirrors [
19,
20], and tripartite entanglement between an LG-cavity mode, a magnon mode, and a phonon mode [
21].
An LG laser beam carries an orbital angular momentum of
per photon, where
l is the topological charge [
22]. Its phase distribution is helical and it contains a phase singularity at which the intensity is zero [
23]. An LG laser beam can be produced by using spatial light modulators [
24], computer-generated holograms [
25], and spiral-phase elements [
23]. The topological charge of the LG laser beam generated by the spiral phase element can be tuned by changing the step height and the refractive index of the spiral-phase element [
23].
It is well known that the cross-Kerr effect is the change in the refractive index of a material in the presence of an applied electromagnetic field, and that the induced refractive index change is proportional to the intensity of the applied electromagnetic field [
26]. It has been shown that the cross-Kerr effect can be used to realize a quantum non-demolition measurement [
27], the generation of multiphoton Greenberger–Horne–Zeilinger states [
28], the construction of a controlled NOT gate [
29], the creation of strong micro–macro entanglement [
30], and so on. In optomechanical systems, it has been theoretically proposed [
31] and experimentally demonstrated [
32] that a cross-Kerr type of coupling between the mechanical oscillator and the microwave cavity field can be realized by using a two-level system (qubit or artificial atom). In Ref. [
31], it was shown that cross-Kerr coupling can lead to a cavity frequency shift, which is related to the phonon number of the mechanical oscillator. In the past few years, cross-Kerr coupling in optomechanical systems has been widely studied. It has been found that cross-Kerr coupling can affect the mechanical frequency shift and the optical damping rate in red and blue mechanical sidebands [
33]. Cross-Kerr coupling can also make the bistable behavior of an intracavity photon number occurring at lower laser power [
34,
35]. Moreover, cross-Kerr coupling can lead to an improvement in optical nonreciprocity, nonreciprocal photon amplification, and optical squeezing [
36]. Additionally, cross-Kerr coupling can be used to enhance the slow light effect [
37]. Furthermore, cross-Kerr coupling affects photon blockade, the magnitude of single-photon mechanical displacement, and the generation of mechanical cat states [
38]. Moreover, cross-Kerr coupling can improve the steady-state entanglement between optical and mechanical modes [
39,
40].
In this paper, we study the steady-state entanglement between an optical cavity field and a rotating mirror in an LG-cavity optomechanical system in the presence of cross-Kerr nonlinearity. We analyze how the cross-Kerr coupling strength, the cavity detuning, the input laser power, the topological charge of the LG-cavity mode, and the temperature of the environment affect the steady-state optomechanical entanglement. In the presence of cross-Kerr nonlinearity, we find that the steady-state optomechanical entanglement is greatly improved and is more robust against the thermal noise of the surrounding environment.
The article is organized as follows. In
Section 2, we introduce the LG-cavity optomechanical system with cross-Kerr nonlinearity, we present the Hamiltonian of the whole system, we provide the quantum Langevin equations, and we calculate the steady-state amplitudes of the cavity and mechanical modes. In
Section 3, we use logarithmic negativity as a measure of entanglement between the cavity and mechanical modes in the steady state. In
Section 4, we discuss the influence of cross-Kerr coupling strength, cavity detuning, input laser power, the topological charge of an LG-cavity mode, and the temperature of the environment on steady-state optomechanical entanglement. In
Section 5, we summarize our results.
2. Model
We consider an LG-cavity optomechanical system with one fixed mirror and one rotating mirror, as shown in
Figure 1. The left fixed mirror transmits a small portion of the incident light, and the right rotating mirror totally reflects the incident light. The two cavity mirrors are spiral-phase elements which are designed to impart a fixed topological charge onto the incident light [
8,
23]. We assume that an incident Gaussian beam with charge 0 interacts with the fixed mirror, that the charge of the reflected beam is
less than that of the incident beam, and that the charge of the transmitted beam is the same as that of the input beam. Then, the beam with charge 0 interacts with the rotating mirror, and the charge of the reflected beam is
larger than that of the incident beam. In addition, the beam with charge
returns to the fixed mirror and interacts with it, the reflected component has charge 0, and the transmitted component has charge
. Thus, a single LG-cavity mode is created in the optical cavity. Due to the exchange of the orbital angular momentum between the intracavity photons and the rotating mirror, the LG-cavity mode exerts a radiation torque on the rotating mirror so that the rotating mirror vibrates around the
z axes. We can approximate the rotating mirror with mass
m, radius
R, frequency
, and damping rate
as a harmonic oscillator. The optomechanical coupling strength between the cavity field and the rotating mirror is
, where
c is the light speed in vacuum and
L is the length of the optical cavity [
8]. We assume that there is a quantum two-level system on the rotating mirror which induces the cross-Kerr coupling between the cavity field and the rotating mirror [
32]. The cross-Kerr coupling strength is denoted by
. It has been shown that the cross-Kerr coupling strength
can be enhanced by using the optical parametric amplification [
41], periodically modulating the mechanical spring constant [
42], and using the Josephson capacitance of a Cooper-pair box [
43]. The motion of the rotating mirror is described by the angular displacement
and the angular momentum
, which satisfies the commutation relation
and can be, respectively, expressed in terms of the annihilation operator
b and the creation operator
of the rotating mirror, i.e.,
,
. The operators
b and
of the rotating mirror satisfy the commutation relation
.
The Hamiltonian of the system in the rotating frame at the frequency
of the input Gaussian beam is given by
where
a (
) is the annihilation (creation) operator of the cavity mode with the resonance frequency
;
is the frequency of the input Gaussian laser beam;
g is the single-photon optomechanical coupling strength given by
, with
being the moment of inertia of the rotating mirror; and
is the amplitude of the input Gaussian beam given by
, depending on the power
℘ of the input Gaussian beam and the decay rate
of the cavity field. In Equation (
1), the first two terms are the free energies of the cavity field and the rotating mirror, respectively, the next two terms represent the optomechanical interaction and the cross-Kerr coupling between the cavity field and the rotating mirror, respectively, and the last term describes that the cavity mode is driven by the input Gaussian beam.
It is noted that the photon loss in the cavity through the left mirror is characterized with the decay rate
, and the energy loss of the rotating mirror is characterized by the damping rate
. Applying the Heisenberg equation of motion and adding the damping terms and the noise terms, we obtain the time evolution of the optical and mechanical modes
where
is the input optical noise operator associated with the vacuum fluctuations of the continuum of modes outside the optical cavity and has zero mean value, and
is the Brownian thermal noise operator of the rotating mirror due to its coupling to the thermal environment and has zero mean value. The time correlation functions for the noises
and
are given by
where
is the mean thermal excitation number at the frequency
of the rotating mirror,
is the Boltzmann constant, and
T is the environmental temperature. The steady-state mean values of the system operators are given by
where
is the cavity detuning,
is the steady-state amplitude of the cavity field, and
is the steady-state amplitude of the mechanical mode. It is seen that both
and
depend on the optomechanical coupling strength
g and the cross-Kerr coupling strength
. Without the optomechanical coupling (
), the steady-state amplitude of the mechanical mode is
.
4. The Stationary Entanglement between the Cavity Field and the Rotating Mirror in the Presence of Cross-Kerr Nonlinearity
In this section, we discuss the effects of the cross-Kerr coupling strength , the cavity detuning , the input laser power ℘, the topological charge l of the LG-cavity mode, and the temperature T of the environment on the stationary entanglement between the cavity field and the rotating mirror.
The parameters we choose are similar to those in Ref. [
17]: the wavelength of the input laser is
nm, the length of the optical cavity is
mm, the photon decay rate is
MHz, the mass of the rotating mirror is
ng, the radius of the rotating mirror is
m, the resonance frequency of the rotating mirror is
MHz, and the damping rate of the rotating mirror is
Hz. Thus, the system is working in the resolved-sideband regime
. It has been demonstrated experimentally that an LG beam with a topological charge of
can be generated by using spiral-phase elements [
48].
In
Figure 2, we plot the photon number
in the cavity field and the phonon number
in the rotating mirror at the steady state of the system as a function of the input laser power
℘ (mW) when
,
, and
. Based on the stable conditions of the system, it is found that the system is stable only if the power
℘ of the input laser is not larger than 0.87 mW. It is seen that an input laser with a higher power can generate a larger photon number
in the cavity, and it also can lead to a larger phonon number
in the rotating mirror due to the optomechanical coupling and the cross-Kerr coupling. As the phonons grow, the input pump photons interact with the phonons, leading to the creation of more cavity photons via absorption of the phonons. It is noted that
and
are much larger than 1, so the linearization assumption we make above is valid.
In
Figure 3, we plot the normalized effective single-photon optomechanical coupling strength
and the normalized effective mechanical frequency
of the rotating mirror as a function of the input laser power
℘ (mW) when
,
, and
. It is seen that the the effective single-photon optomechanical coupling strength
G increases with increasing the power
℘ of the input laser, while the effective mechanical frequency
of the rotating mirror decreases with increasing the power of the input laser. These results are consistent with those in [
39], which studies that the cross-Kerr coupling improves the optomechanical entanglement induced by the radiation pressure.
Figure 4 shows the logarithmic negativity
as a function of the normalized cavity detuning
for different topological charges
l of the LG-cavity mode when
mW,
K, and
. Without the cross-Kerr coupling (
), for
, and 110, the stability conditions of the system require that the cavity detuning satisfies
, 0.02, 0.03, 0.05, and 0.07, respectively. With the cross-Kerr coupling, for
, and 110, the stability conditions of the system require that the cavity detuning satisfies
, 0.18, 0.25, 0.30, and 0.35, respectively, when
and
, 0.26, 0.33, 0.39, and 0.45, respectively, when
. Thus, for a fixed value of the cross-Kerr coupling strength
, as the topological charge
l increases, the stable regime of the system becomes narrower. Moreover, for a fixed value of the topological charge
l, increasing the cross-Kerr nonlinearity also makes the stable regime of the system smaller. Next, we first look at the case without the cross-Kerr coupling (
). For
, as the cavity detuning
increases, the logarithmic negativity
first increases and then decreases. For
, and 110, as the cavity detuning
increases, the logarithmic negativity
first increases rapidly to the maximum value and then decreases slowly to the minimum value, again increases, and then decreases. Then, we consider the case of
. For
, with increasing the cavity detuning
, the logarithmic negativity
first increases and then decreases. For
, and 110, the logarithmic negativity
decreases from the maximum value with the increase in the cavity detuning
, and the logarithmic negativity
takes the maximum value at a cavity detuning
close to the unstable regime of the system. When
, the behavior of the logarithmic negativity
for
, and 110 is the same as that for
, and 110 when
. Importantly, for a fixed value of the topological charge
l, it is seen that the maximum entanglement between the cavity field and the rotating mirror in the presence of the cross-Kerr nonlinearity is much larger than that in the absence of the cross-Kerr nonlinearity. Without the cross-Kerr coupling (
), for
, and 110, the logarithmic negativity
takes the maximum value of about 0.001, 0.003, 0.007, 0.012, and 0.018 at
, and
, respectively. In the presence of the cross-Kerr coupling, for
, and 110, the logarithmic negativity
takes the maximum value of about 0.031, 0.201, 0.211, 0.282, and 0.272 at
, 0.18, 0.25, 0.30, and 0.35 when
, and takes the maximum value of about 0.255, 0.449, 0.459, 0.405, and 0.307 at
, and
when
. Thus, for a given nonzero value of the cross-Kerr coupling strength
, as the topological charge
l increases, the maximum optomechanical entanglement occurs at a larger cavity detuning
, but is not always increased. For a given value of the topological charge
l, as the cross-Kerr coupling strength increases, the maximum optomechanical entanglement occurs at a larger cavity detuning
, and it is always enhanced.
Figure 5a plots the logarithmic negativity
as a function of the normalized cross-Kerr coupling strength
for different topological charges
l of the LG-cavity mode when
mW,
, and
K. For
, and 110, the system is stable if the normalized cross-Kerr coupling strength
is not larger than
,
,
,
, and
, respectively. Thus, the larger the topological charge
l, the narrower the stable regime of the system. For a fixed value of the topological charge
l, the logarithmic negativity
increases with increasing the cross-Kerr coupling strength
, and the logarithmic negativity
takes the maximum value at a cross-Kerr coupling strength
close to the unstable regime. It is noted that the real part of the effective single-photon optomechanical coupling strength
G is found to be much larger than its imaginary part by at least six orders of magnitude. For example, when
,
, so
G can be approximated as a real number. From
Figure 5b, it is seen that the effective single-photon optomechanical coupling strength
G is increased with the increase in the cross-Kerr coupling strength
, thereby leading to the enhancement of the entanglement as shown in
Figure 5a. Moreover, for a larger topological charge
l,
increases with increasing
at a faster rate. For
, and 110, the maximum value of the logarithmic negativity
is about 0.252, 0.373, 0.459, 0.402, and 0.449 at
, and
, respectively. Therefore, as the topological charge
l increases, the maximum optomechanical entanglement occurs at a weaker cross-Kerr coupling strength
, but is not always increased.
Figure 6 plots the logarithmic negativity
against the input laser power
℘ (mW) for different cross-Kerr coupling strengths
when
,
, and
K. For
and
, the stable conditions of the system require that the input laser power
℘ is not larger than 11.2 mW, 2.38 mW, 1.47 mW, 1.08 mW, and 0.87 mW, respectively. Thus, the stronger the cross-Kerr coupling, the smaller the stable regime of the system. Without the cross-Kerr coupling (
), the logarithmic negativity
first increases and then decreases with increasing the input laser power
℘. The maximum value of the logarithmic negativity
is only about 0.063 at
mW. For a nonzero cross-Kerr coupling strength
, the logarithmic negativity
increases as the input laser power
℘ increases. The reason is that increasing the input laser power
℘ leads to a stronger optomechanical coupling and a stronger cross-Kerr coupling between the cavity and mechanical modes due to an increase in the cavity photon number and an increase in the mechanical phonon number as shown in
Figure 2. It is also seen that the logarithmic negativity
takes the maximum value at an input laser power
℘ near the unstable regime. For a larger cross-Kerr coupling strength
, the logarithmic negativity
increases faster with increasing the input laser power
℘. For
and
, the maximum value of the logarithmic negativity
is about 0.403, 0.421, 0.390, and 0.459 at
mW, 1.47 mW, 1.08 mW, and 0.87 mW, respectively. Therefore, the maximum entanglement for the case with the cross-Kerr coupling is larger than that for the case without the cross-Kerr coupling. As the cross-Kerr coupling strength
increases, the maximum entanglement happens at a smaller input laser power
℘, but is not always enhanced.
Figure 7 shows the logarithmic negativity
versus the topological charge
l of the LG-cavity mode for different cross-Kerr coupling strengths
when
,
mW, and
K. For
and
; the stable conditions of the system require that the topological charge
l is not larger than 252, 144, 105, 83, and 70, respectively. Thus, when the cross-Kerr coupling strength
is larger, the stable regime of the system is narrower. We first consider the case without the cross-Kerr coupling (
). When the cross-Kerr coupling strength
increases, the logarithmic negativity
first increases and then decreases. The logarithmic negativity
begins to be nonzero (the entanglement between the cavity field and the rotating mirror starts to appear) when the topological charge
l is equal to 26. The maximum value of the logarithmic negativity
is only about 0.063 at
. Next, we consider the case with the cross-Kerr coupling (
). For
and
, the logarithmic negativity
begins to be nonzero when the topological charge
l is equal to 23, 22, 20, and 19, respectively. Hence, for a stronger cross-Kerr coupling, the entanglement between the cavity field and the rotating mirror appears at a smaller topological charge
l, and it reaches the maximum value just before the unstable regime. For a stronger cross-Kerr coupling, the logarithmic negativity
increases more quickly with increasing the topological charge
l. For
and
, the maximum value of the logarithmic negativity
is about 0.326, 0.400, 0.369, and 0.459 at
and 70, respectively, which is much larger than that without the cross-Kerr coupling (
). Hence, the presence of the cross-Kerr coupling makes the maximum entanglement larger than that without the cross-Kerr coupling. For a larger cross-Kerr coupling strength
, the maximum entanglement happens at a smaller topological charge
l. In addition, it is noted that the maximum entanglement happens when
and
since the maximum value of the logarithmic negativity
is the largest when
and
.
Figure 8 is a plot of the logarithmic negativity
versus the temperature
T of the environment for different cross-Kerr coupling strengths
when
mW,
, and
. For a fixed value of the cross-Kerr coupling strength
, it is seen that the logarithmic negativity
decreases with the increase in the temperature
T of the environment. For
and
, the logarithmic negativity
takes the maximum value of about 0.008, 0.044, 0.122, 0.230, and 0.460 at
K, respectively, and the logarithmic negativity
becomes zero when
K, 3.2 K, 9.6 K, 23.4 K, and 70.8 K, respectively. Hence, for a stronger cross-Kerr coupling, the entanglement between the cavity field and the rotating mirror can survive at a higher temperature
T of the environment.
Finally, we discuss the possibility of experimentally detecting the generated entanglement between the LG-cavity mode and the rotating mirror. In order to measure the logarithmic negativity
, we need to measure the ten independent elements of the correlation matrix
V. For the cavity mode, its second moments can be directly measured by homodyne detecting the output field from the cavity. For the mechanical mode, we need to add a second cavity, adjacent to the first one, formed by the rotating mirror and a third transmissive fixed spiral phase element. There is no coupling between the two cavities. The angular displacement and the angular momentum of the rotating mirror can be measured by homodyning the output field of the second cavity via choosing the suitable parameters of the second cavity [
49].
Without the cross-Kerr coupling, it is noted that the entanglement (
) between the LG-cavity mode and the rotating mirror in Ref. [
17] is much larger than the entanglement (
) between the LG-cavity mode and the rotating mirror in our scheme. This is because they use a high laser power, 50 mW, and a small cavity length, 0.1 mm. In our scheme, a larger entanglement (
) between the LG-cavity mode and the rotating mirror can be achieved in the presence of the cross-Kerr coupling at a low laser power, 0.87 mW, and a large cavity length, 1 mm.