Parity Deformed Jaynes-Cummings Model: "Robust Maximally Entangled States"
Parity Deformed Jaynes-Cummings Model: "Robust Maximally Entangled States"
com/scientificreports
The Jaynes-Cummings model (JCM) which is used extensively in quantum optics describes the interaction of
a single quantized radiation field with a two-level atom. The solvability and applications of this model has long
been discussed1,2. This simple model describes various quantum mechanical phenomena, for example, Rabi oscil-
lations2,3, collapse and revivals of the atomic population inversion4 and entanglement between atom and field5.
Specifically, it plays an important role in recent quantum information processing6–11. Furthermore, JCM is one
of the several possible schemes for producing the nonclassical states12–15. The dynamics predicted by JCM have
been proven in experiments with Rydberg atom in high quality cavities16. Since JCM is an ideal model in quantum
optics, its various extensions such as intensity dependent coupling, two photons or multi-photon transitions, two-
or three- cavity modes for three-level atoms and the Tavis-Cummings model have been proposed17,18. In 1984,
Sukumer and Buck studied the above mentioned models by using algebraic operator methods19. On the other
hand, the supergroup theoretical approach to JCM leads to the exact solvability of this model and the representa-
tion theory of super-algebras20. More recently, a lot of researchers have found that the ordinary creation and anni-
hilation operators in JCM may be replaced by the q-deformed partners, namely, the q-deformed JCM21. Later on,
JCM has been adopted with a Kerr nonlinearity within the framework of f-oscillator formalism22. Furthermore,
the investigations of a class of shape-invariant bound state problem, which represents a two-level system, leads to
the generalized JCM23.
In addition to the above -cited generalizations, in recent years, a lot of interest has been given to to the exten-
sion of the boson oscillator algebra. One of the most interesting cases which is not related to the q or f-calculus is
R-deformed Heisenberg algebra (RDHA). Such generalized schemes naturally lead to the concept of para-fields
and para-statistics18–36. The same RDHA was also used for solving the quantum mechanical Calogero model or
pseudo harmonic oscillator (PHO) 37–41. Recently, this algebra has been employed for bosonization of
super-symmetric quantum mechanics41–43 and for describing anyons in (2 + 1)42,44 and (l + 1) dimensions45,46. All
the applications as well as the para-bosonic constructions32,33 have used infinite-dimensional unitary representa-
tions of RDHA. According to Wigner’s new quantization method, the RDHA is raised as a unital algebra by the
generators {1, a , a†, Rˆ }, which satisfy the (anti-) commutation relations
1
Department of Physics, Payame Noor University, P.O. Box 19395-3697, Tehran, I.R. of Iran. 2Department of Physics,
Azarbaijan Shahid Madani University, P.O. Box 51745-406, Tabriz, Iran. Correspondence and requests for materials
should be addressed to A.D. (email: [email protected]) or B.M. (email: [email protected]) or S.S.
(email: [email protected])
R̂ n = ( − 1)n n , (3)
which means that R̂ commutes with number operator N̂ that includes the eigenvector n , such that N̂ n = n n .
The number operator N̂ is in general different from the product a†a , and is postulated to satisfy the following
relations:
a 2n = 2 n 2 n − 1 , a 2n + 1 = 2n + 2λ + 1 2n , (6)
a† 2n = 2n + 2λ + 1 2n + 1 , a† 2n + 1 = 2n + 2 2n + 2 . (7)
†
The explicit differential forms of the generators a , a in terms of the well known annihilation and creation opera-
tors a, a† are obtained37 as follows:
1 d λ λ ˆ
a= + x − Rˆ = a − R,
2 dx x 2x (8)
1 d λ λ ˆ
a† = − + x + Rˆ = a† + R,
2 dx x 2x (9)
which provide us with the coordinate representation for the position x̂ and its λ−deformed canonical pair p̂ as
a + a†
xˆ = ,
2 (10)
a − a† d λ
pˆ = = −i + i Rˆ .
i 2 dx x (11)
It is evident that the above representation is really different from the f- deformed realization of the Heisenberg
algebra47. Therefore, the introduced RDHA in (1) can be considered as a new deformation of the simple harmonic
oscillator with significant features in quantum optics48,49.
Due to the physical significance of the deformed JCM in quantum optics on the one hand, and the central role
of the parity operator in the theory of deformation on the other hand, we then generalize the well-known JCM to
a parity deformed-Hermitian case in terms of parity deformed boson operators which are not related to the q or
f-deformation. It should be noted that the parity deformed JCM introduced here could be compared, with a major
difference, to what has been already discussed in other studies as the q-, f-deformed JCM and so on. It is one of
the most interesting cases of the parity operator R̂ appeared in the context of quantization schemes generalizing
bosonic commutation relations. These Hermitian operators arise from a special deformation of canonical bosonic
commutation relations, allowing us a mathematically rigorous treatment of our deformed interaction
Hamiltonian, extraction of the energy spectrum, and the corresponding eigen-vectors. Preparing the initial field
in the λ−deformed cat states, we investigate the collapse and revival phenomena in the Rabi oscillations of the
atomic inversion.
Considering the experimental realization, interaction with the environment is an unavoidable feature of real
quantum systems. Dissipation phenomenon of energy into the environment is the main feature of such real sys-
tems which leads to the loss of entanglement generated in those systems. Some scholars have studied this subject
and found interesting phenomena such as entanglement revival50, entanglement sudden death51,52 and sudden
change for quantum discord53. The description of dissipative systems Hamiltonian is a polemic topic in the lit-
erature54,55. The dissipative systems can be studied in a phenomenological way in which a closed system formed
by the cavity including its environment is considered. This larger system is therefore closed, and it is possible to
apply quantum mechanics to it in a usual way56. The mentioned environment can be regarded as a discrete set of
degrees of freedom, or a set of harmonic oscillators57. In addition, the surrounding environment can be modeled
as a set of continuum harmonic oscillators58. The Hamiltonian describing this model of dissipation is called the
Gardiner-Collett Hamiltonian59. Also, based on the dissipation of the atomic upper level, a model describing
dissipative atom-field system is introduced60. This model described by a non-Hermitian Hamiltonian is called
damping JCM. The authors have studied decaying behaviour of the entanglement between atom and field. Also,
one of the interesting topics in dissipative systems is to find a way to fight against the decay of the entanglement.
Many schemes have been proposed in order to preserve entanglement in such systems. For instance, it has been
shown that, the addition of a laser field leads to high stationary entanglement57. Another approach to overcome
this problem, relies on active feedback61–63. In addition to these, the quantum Zeno effect is a promising way to
avoid the decaying behaviour of the entanglement in dissipative systems64–68. In the dissipation regime, we use
the damping deformed JCM to describe a two-level atom interaction with the dissipative cavity and investigate
individually and simultaneously the effects of dissipation and detuning parameters on the entanglement between
the two-level atom and the deformed field. It has been shown that by detuning modulation, and setting the defor-
mation parameter, coupling constant and average photon number of the field, the degree of entanglement of the
atom-field states, and the atomic population inversion may be controlled.
ω ω λ (λ − 1) ω 2 gλ ˆ
Hλ = {a, a†} + 2
+ 0 σ 3 + g (a†σ− + aσ +) − i Rσ y ,
2 2 x 2 x (13)
ω λ (λ − 1) 2 gλ ˆ
= H JCM + −i Rσ y .
2 x2 x (14)
It recalls a system of a two-level atom coupled simultaneously to a single-mode of the quantized electromag-
netic field and a centrifugally external classical field. It reduces to an ordinary JCM, while λ →0. Here, σ± and σ3
are usual rising (lowering) and inversion operators for the atomic states, |±〉, satisfying [σ + , σ−] = σ3 and
[σ3, σ ±] = ±2σ±, a† and a are the creation and annihilation operators for the cavity mode [a, a†] = 1, g is the
coupling constant between the atom and the cavity field mode, ω and ω0 are the cavity as well as the external mode
frequency and the atomic transition frequency, respectively, λ represents the strength of the external field,
E ext d ≡ x Rˆ is the coupling constant between the atom and the external classical field where d recalls the
2 gλ
atomic dipole matrix element for the transition and Eext refers to the amplitude of the external field. It is notewor-
thy that the Hamiltonian Hλ is super-symmetric when ω = ω0 (exact resonance) and g = 0 (absence of coupling).
This exactly solvable model includes the inversely quadratic potential, 1 λ (λ −2 1) , first proposed by Post in 1956
2 x
when he studied the one-dimensional many identical particles problem in the case of the pair-force interaction
between the particles69. In addition to the two typical model potentials such as the Morse potential and Poschl-
Teller potential, this an-harmonic oscillator potential can be considered as a good candidate to describe the
molecular vibrations too70–72. Since 1956, this quantum system was studied by other researchers, extensively. For
example, Landau and Lifshitz studied its exact solutions in three dimensions73. Recently, Sage has studied the
vibrations and rotations of the pseudo-gaussian oscillator in order to describe the diatomic molecule74, in which
he briefly reinvestigated some properties of the PHO in order to study the pseudo-gaussian oscillator. Advantages
of the pseudo-harmonic potential have been considered for improvements in the conventional presentation of
molecular vibrations75. Hurley found that this kind of PHO interaction between the particles can be exactly
solved by the separation of variables while studying the three-body problem in one dimension76. A few years later,
Calogero studied the one-dimensional three- and N-body problems interacting pairwise via harmonic and
inverse square (centrifugal) potential77,78. On the other hand, this potential was generalized by Camiz and
Dodonov et al. to the non-stationary (varying frequency) PHO potential79,80. In addition, such a physical problem
was also studied in arbitrary dimension. Also, Dong et al. have studied its dynamical group in two dimensions81.
This parity deformed JCM model possesses an exact solution because of the existence of an integral of motion,
1
a†a + 2 σ 3, which commutes with the Hamiltonian Hλ and allows us to decompose all the representation space of
the atom-field system as the tensor product of the Hilbert space associated with the field, λ , times the Hilbert
space associated with the spin, f ,
∞
2n
:=
⊗ f = 2n, + = , 2n + 1, − = 0 ,
λ
0 2 n + 1
n =0 (15)
or
∞
2n + 1 0
2n + 1, + =
, 2n + 2, − = 2n + 2 .
0
n =0 (16)
±
Figure 1. Dependence of eigenvalues En, λ on detuning Δ. The continuous curve corresponds to g = 0.01. The
dashed curves (a) and (b) correspond to (g = 0, λ = 0) and (g = 0, λ = 50), respectively. The dashed curves with
positive and negative slopes correspond respectively to En,+λ and En,−λ . Lower part of the figure is for n = 1 and
upper part for n = 2.
Here, 2n, + is the bare state in which the atom is in the excited state + and the field has 2n photons, and a
similar description holds for the bare state 2n + 1, − , where − is the atom ground state. Using the Fock space
given in (15), we can find the following matrix representation of the λ−deformed JC Hamiltonian Hλ:
ω 2n + λ + 1 + ω0 g 2n + 2λ + 1
2
2
Hλ = .
3 ω
g 2n + 2λ + 1 ω 2n + λ + − 0
2 2 (17)
It is easy to see that the corresponding dressed eigen-states of Hλ are
2g 2n + 2λ + 1
c2 = ,
(∆ − Ωn , λ )2 + 4g 2(2n + 2λ + 1) (21)
2 2
in which Δ(= ω − ω0) and Ωn, λ ( = ∆ + 4g (2n + 2λ + 1) ) are defined as detuning parameter and a general-
ized Rabi frequency, respectively. The energy eigenvalues corresponding to the eigen-states in Eqs (18) and (19)
are
Ωλ
En,±λ = (2n + λ + 1)ω ± .
2 (22)
The energy difference between the levels En,+λ and En,−λ
is Ωn,λ. The minimum of the separation occurs when Δ
equals to zero and the corresponding difference is 2g 2n + 2λ + 1 . In Fig. 1(a) and (b), respectively, we have
plotted the energy eigenvalues En,+λ and En,−λ as functions of Δfor given values of λ = 0, 50. The dotted lines rep-
resent the eigenvalues when g = 0, i.e. En,±λ = (2n + λ + 1)ω ± ∆ . In this case, the eigenvalues cross each other
2
as they increase from negative to positive values. The continuous lines represent the energy eigenvalues for
g = 0.01. The diverging eigenvalue separation beyond the minimum separation indicates level repulsion in the
eigenvalues of the dressed states. As Fig. 1(a) and (b) show, the repulsion between energy levels increases as the
deformation parameter λgets bigger. The latter, also, leads to shift the energy levels to the positive side.
1 1
H 0 = ω a†a + + λRˆ + ω0σ 3,
2 2 (24)
H ′ = g (a†σ− + aσ +) . (25)
In the interaction picture generated by H0, the Hamiltonian of the system can be written as
Ω ∆ Ω
c +,2n (t ) = c +,2n(0)cos n , λ t + i sin n , λ t
2 Ωn , λ 2
2n + 2 λ + 1 Ω ∆
− 2ig c−,2n +1(0) sin n , λ t e−i 2 t ,
Ωn , λ 2 (29)
Ω ∆ Ω
c−,2n +1(t ) = c−,2n +1(0)cos n , λ t − i sin n , λ t
2 Ωn , λ 2
2n + 2 λ + 1 Ωn , λ i ∆ t
− 2ig c +,2n(0) sin t e 2 ,
Ωn , λ 2 (30)
where the constants c−,2n+1(0) and c+,2n(0) are determined from the initial conditions of the system, which is sup-
posed initially in the upper level, i.e. c+,2n(0) = c2n(0) and c−,2n+1(0) = 0. Here, the initial condition for the field is
described by c2n(0). For this case in particular, we have
Ω ∆ Ω ∆
c +,2n (t ) = c 2n(0)cos n , λ t + i sin n , λ t e−i 2 t ,
2 Ωn , λ
2 (31)
2n + 2λ + 1 Ω ∆
c−,2n +1(t ) = − 2ig c 2n(0) sin n , λ t e i 2 t ,
Ωn , λ 2 (32)
This set of equations gives us the solution for the problem. In order to calculate some physical quantities of
interest, we need only to specify the initial photon number distribution of the field |c2n(0)|2.
The field we are considering in this work is being treated as an λ-deformed oscillator which could be described
in different ways. We focus on the situations in which the field as an eigen-state of the λ-deformed annihilation
operator is introduced48, i.e. a2 W = w 2 W λ, +, and its number state expansion is
λ, +
( )
2λ− 1
w
2
∞
w 2n
w λ, +
:=
I λ− 1 ( w 2
)
∑ 2n 1
2n .
2
n =0 2 n ! Γ (n + λ + 2 ) (33)
Therefore, in this case, we have
( )
4n + 2λ− 1
w
2
c 2n(0) 2 = 1
.
n! Γ (n + λ + 2 ) I λ− 1 (|w|2 ) (34)
2
It needs to be noted that, as a second option, one may choose the initial state of the field as the Wigner neg-
ative binomial states already studied49. They are mathematically equivalent to those nonlinear ones and may be
expected to bring new quantum features.
1.0 1.0
b
0.5 0.5
0.0 0.0
0.5 0.5
Z
0 20 40 60 80 0 20 40 60 80
gt gt
1.0
c
0.5
0.0
0.5
Z
1.0
0 20 40 60 80
gt
Figure 2. Temporal evolution of the atomic inversion 〈σz〉 for the field initially prepared in WCS, w ,
λ, +
with |w|2 = 30 and g = 0.01. The parameters are (a) λ = 0, Δ = 0, (b) λ = 0, Δ = 0.01 and (c) λ = 50, Δ = 0.01.
∞ 2 g
2
∆
σz = ∑ c 2n(0) 2 + (8n + 8λ + 4) cos(Ωλ t ) .
n =0 Ωλ Ω
λ (36)
The numerical results of the atomic inversion when the field is in a standard cat state (i.e. when λ = 0), have
been shown against the scaled time gt in Fig. 2(a) and (b), respectively. The temporal evolution of the atomic
inversion 〈σz〉reveals significant discrepancies of the well-known phenomenon of collapses and revivals82. Note
that the collapse, i.e. when the envelope of the oscillations collapses to zero, is due to the destructive interference
among the probability amplitudes at different Rabi frequencies, Ωn,λ, for different photon number eigen-states. At
the revival times, on the other hand, constructive interference occurs. This phenomenon also takes place when
the initial field state is a WCS. In Fig. 2(c), we have pictured the function 〈σz〉for the value λ = 50. In this case,
〈σz〉exhibits quasi periodic behaviour very similar to the atomic inversion of a two-photon JCM83–85, with an
( ω
)
effective Hamiltonian defined as H eff = ω a†a + 1 + 0 σ 3 + g (a† 2σ− + a2σ +). However, in this case, note
2 2
how the structure of the oscillations is much more complex than the standard Rabi oscillations. As the detuning
factor Δincreases, these structures are disappeared (see Fig. 3), i.e. the inhibition of the radiation decay is more
transparent. It is clear that the inhibited decay even occurs in the case λ = 0. This behaviour is due to the influence
of the parity deformation via the generalized Rabi frequency Ωn,λ. Figure 3(a–c) indicate that, with increasing λ
the inhibition decay of the excited state will be balanced.
The finite lifetime of the atomic levels can be described by adding phenomenological decay terms, − i γ + + ,
2
to the Hamiltonian (9), where γ ∈ is a decay constant60. For an atom initially stated in the upper level, |+〉, the
population inversion dynamics are analyzed for a deformed JCM was surrounded by a dissipative environment
where the dissipation of the upper-level is considered (see Fig. 4). Here, we plot the dynamics of the population
inversion in the presence of the decay term against the scaled time gt. In Fig. 4, we compare the effects of increas-
ing γ on the Rabi oscillations, where the latter is destroyed. However, as is illustrated in Fig. 5, the patterns of the
revivals are restored while λis enhanced.
Fidelity
We now calculate the fidelity
2
F= Ψ(0) Ψ (t ) , (37)
1.0 1.0
b
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
Z
0.4 0.4
0.3 0.3
0 50 100 150 200 0 50 100 150 200
gt gt
1.0
c
0.9
0.8
0.7
0.6
0.5
Z
0.4
0.3
0 50 100 150 200
gt
Figure 3. Temporal evolution of the atomic inversion with |w|2 = 30, g = 0.01 and Δ = 0.5. The parameters
equal to (a) λ = 0, (b) λ = 30 and (c) λ = 100.
1.0 1.0
a
b
0.5 0.5
0.0 0.0
Z
0.5 0.5
1.0 1.0
0 20 40 60 80 0 20 40 60 80
gt gt
1.0 1.0
0.8
d
c
0.5 0.6
0.4
0.0
0.2
0.0
Z
0.5
0.2
0.4
1.0
0 20 40 60 80 0 20 40 60 80
gt gt
Figure 4. Temporal evolution of the atomic inversion with |w|2 = 30, g = 0.01, Δ = 0.5 and λ = 30. The
parameters equal to (a) γ = 0, (b) γ = 0.001, (c) γ = 0.01 and γ = 0.1.
which measures the “closeness” of the two quantum states Ψ (t ) and Ψ(0) = w ⊗ + , which indicate that
λ, +
F is unity when these two quantum states are identical. Figure 6 indicates that with increasing λthe fidelity
decreases but remains close to its initial value ( see Fig. 5(a)–(d)). To obtain fidelity around 1, one needs to
enhance the deformation parameter λto 100 and gt = 95. In this case, Ψ (t ) becomes minimum uncertainty state
which minimize the uncertainty relation, Eq. (28) in ref. 48.
1.0 1.0
a
0.8
b
0.8
0.6
0.6
0.4
0.4
0.2
0.2
Z
0.0
0.0
0.2
0 1 2 3 4 5 0 1 2 3 4 5
gt gt
1.0 1.0
0.8 0.8
d
c
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
Z
Z
0.2 0.2
0.4 0.4
0 1 2 3 4 5 0 1 2 3 4 5
gt gt
Figure 5. Temporal evolution of the atomic inversion with |w|2 = 30, g = 0.01, Δ = 0.5 and γ = 0.05. The
parameters equal to (a) λ = 0, (b) λ = 200, (c) λ = 500 and λ = 1000.
1.0 1.0
0.8 0.8
a
b
0.6 0.6
0.4 0.4
0.2 0.2
F
0.0 0.0
0 20 40 60 80 100 0 20 40 60 80 100
gt gt
1.0 1.0
0.8 0.8
d
c
0.6 0.6
0.4 0.4
F
0.2 0.2
F
0.0 0.0
0 20 40 60 80 100 0 20 40 60 80 100
gt gt
Figure 6. Fidelity as a function of the scaled time of gt in the small coupling regime with g = 0.01 for
different λ( = −0.25, 0, 10, 100), other parameters are |w|2 = 9 and Δ = 0.1.
operational measure of the disorder of a system and of the purity of a quantum state. For given density operator
ρ the Von Neumann entropy is given by
S (ρ ) : = Tr[ρ ln ρ], (38)
where “Tr” often abbreviated to the trace and S(ρ) ranges from 0 for a separable state to 1 for a maximally entan-
gled one. As the deformed JCM is a bipartite system, a Schmidt decomposition is assured. Based on the solution
of the time-dependent Schrodinger equation (23), and according to the Schmidt decomposition, for any instant
in time t, we can always find the reduced density operator of the atom in the bare basis specified by |±〉and obtain
1.0 1.0
b
0.8 0.8
0.6 0.6
0.4 0.4
SA
SA
0.2 0.2
0.0 0.0
0 20 40 60 80 100 0 20 40 60 80 100
gt gt
1.0 1.0
d
c
0.8 0.8
0.6 0.6
0.4 0.4
SA
SA
0.2 0.2
0.0 0.0
0 20 40 60 80 100 0 20 40 60 80 100
gt gt
Figure 7. Plots of entropy SA versus gt with g = 0.01, |w|2 = 9 and Δ = 0.1 for various deformation parameters
respectively (a) λ = −0.25, (b) λ = 0, (c) λ = 10 and (d) λ = 50.
1.0
0.8
0.6
0.4
0.2 F
SA
0.0
95.0 95.5 96.0 96.5 97.0
gt
Figure 8. The Von Neumann entropy (dash) and the fidelity (solid) are plotted together. In each case a
deformation parameter with λ = 100, g = 0.01, |w|2 = 9, Δ = 0.1 is used.
∞
∑ c +,2n (t ) 2 0
n =0
ρ A = .
∞
2
0 ∑ c−,2n +1(t )
n =0 (39)
Clearly, eigenvalues of the density operator for the atom, g±, can be expressed in terms of the coefficients c+,2n(t)
2 ∞ 2
and c−,2n+1(t) i.e. g + = ∑ ∞
n = 0 c +,2n (t ) and g − = ∑ n = 0 c−,2n + 1(t ) . Then, it is easy to obtain an expression for
the Von Neumann entropy. For atomic subsystem, this is:
S A = − g + ln g + − g − ln g −. (40)
In Fig. 7, for an atom initially in an excited state and the field initially in WCS, we plot the Von Neumann
entropy SA, all against the scaled time gt. We can also see that the entropy makes quasi-period oscillation. This
means that the deformed field can help to realize and stabilize the degree of entanglement between the atom and
the field at a high level. Sometimes, as the deformation parameter λis enhanced, the atom- field system becomes
maximally entangled (Fig. 7(c) and (d)). In Fig. 8, we compare the Von Neumann entropy and fidelity of the quan-
tum state, Ψ( t), associated with the λ−d
eformed JCM investigated here. Bear in mind that entanglement between
the atom and field is maximized i.e. the quantum state Ψ(t) becomes maximally entangled, where fidelity of the
quantum state passes the smallest value.
1.0 1.0
b
0.8 0.8
0.6 0.6
0.4 0.4
SA
SA
0.2 0.2
0.0 0.0
0 20 40 60 80 100 0 20 40 60 80 100
gt gt
1.0 1.0
d
c
0.8 0.8
0.6 0.6
0.4 0.4
SA
SA
0.2 0.2
0.0 0.0
0 20 40 60 80 100 0 20 40 60 80 100
gt gt
Figure 9. The time evolution of atomic entropy SA in the dissipative regime for g = 0.01, |w|2 = 9, Δ = 0.1 and
λ = 2, (a) γ = 0, (b) γ = 0.01, (c) γ = 0.1 and (d) γ = 1.
1.0 1.0
b
a
0.8 0.8
0.6 0.6
0.4 0.4
SA
SA
0.2 0.2
0.0 0.0
0 10 20 30 40 50 0 10 20 30 40 50
gt gt
1.0 1.0
c
0.8 0.8
0.6 0.6
0.4 0.4
SA
SA
0.2 0.2
0.0 0.0
0 10 20 30 40 50 0 10 20 30 40 50
gt gt
Figure 10. The time evolution of atomic entropy SA in the dissipative regime with g = 0.01, |w|2 = 9, λ = 2 and
γ = 0.1, (a) Δ = 0, (b) Δ = 0.01, (c) Δ = 0.1 and (d) Δ = 1.
1.0 1.0
0.8 0.8
b
0.6 0.6
0.4 0.4
SA
SA
0.2 0.2
0.0 0.0
0 20 40 60 80 100 0 20 40 60 80 100
gt gt
1.0 1.0
0.8 0.8
d
c
0.6 0.6
0.4 0.4
SA
SA
0.2 0.2
0.0 0.0
0 20 40 60 80 100 0 20 40 60 80 100
gt gt
Figure 11. The time evolution of atomic entropy SA in the dissipative regime with g = 0.01, |w|2 = 9, γ = 0.1 and
Δ = 0.1, (a) λ = 0, (b) λ = 10, (c) λ = 20 and (d) λ = 80.
while it prolongs the entanglement time. Comparison of Fig. 10(b) with Fig. 10(d), reveals that larger modulation
is unfavorable to the maintenance of the entanglement for a long time.
Fig. 11 refers to the effect of the parity deformation on time evolution of the entanglement. As is shown, in the
absence of λ, the atomic entropy after suddenly increasing to its maximum value at the beginning of the interac-
tion, decay to a minimum asymptotic (stable) value in enough large times (see Fig. 11(a)). As is seen, the presence
of λnot only decreases the decay time but also causes the enhancement of the asymptotic (stable) values of atomic
entropy (see Fig. 11(c–d)) in which for large values of λ, the stable values of entanglement reaches to 1. On the
other words, by increasing λ the generated maximally entangled states are preserved, in the dissipation regime at
the beginning of the interaction.
Figure 12. Plots of the normalized variance, Qλ, as a function of the normalized time gt, for the field
initially prepared in an annihilation operator coherent state w with |w|2 = 20, g = 0.01 and Δ = 0.01.
λ, +
The parameters are (a) λ = −0.25, (b) λ = 0, (c) λ = 2 and (d) λ = 10.
2
1 + 2λR̂
〈σ xx 〉〈σ pp〉 ≥ ,
4 (43)
ˆ ˆ + yx
〈xy ˆ ˆ〉
where 〈σ xy
ˆˆ〉 = 2
− 〈xˆ 〉〈yˆ 〉 and the angular brackets denote averaging over an arbitrary normalizable state
for which the mean values are well defined, yˆ = Ψ (t ) yˆ Ψ (t ) . It can be said that a state is squeezed if the con-
1 + 2λRˆ 1 + 2λRˆ 〉
dition σ xx < or 〈σ pp〉 < is fulfilled98,99. In other words, a quantum state is called squeezed
2 2
state if it has less uncertainty, in one parameter (x̂ or p̂), than a coherent state. Then to measure the degree of
squeezing, we introduce the squeezing factors Sx(p)100, corresponding with the state Ψ(t), respectively
1 + 2λRˆ
σ xx − 2
Sx = ,
1 + 2λRˆ
2 (44)
〈1 + 2λRˆ 〉
〈σ pp〉 − 2
Sp = ,
〈1 + 2λRˆ 〉
2 (45)
which results that the squeezing condition takes the simple form Sxλ(p), i < 0. By using the mean values of the gen-
erators of the WHA,
a = a† = 0 (46)
∞
a2 = ∑ [c+⁎ ,2n (t ) c+,2n +2 (t ) (2n + 2)(2n + 2λ + 1)
n
+ c−⁎ ,2n +1 (t ) c−,2n +3(t ) (2n + 2)(2n + 2λ + 3) ] (47)
a† 2 = a2
∞
a†a = ∑ [(2n) c+,2n (t ) 2 + (2n + 2λ + 1) c−,2n +1(t ) 2 ]
n (48)
∞
aa† = ∑ [(2n + 2λ + 1) c +,2n (t ) 2 + (2n + 2) c−,2n +1(t ) 2 ]
n (49)
Figure 13. Squeezing in the p and x quadratures against gt for different values of λ with g = 0.01, |w|2 = 9 and
Δ = 0.1 as well as for fixed values of φ = 0 and π correspond with (a) and (b), respectively. The solid curve is
2
plotted for λ = 0.
∞
1 + 2λR̂ = ∑ [(1 + 2λ) c+,2n (t ) 2 + (1 − 2λ) c−,2n +1(t ) 2 ]
n (50)
one can derive the variance and covariance of the operators x̂ and p̂. From Eqs (44) and (45), we also stress that
the squeezing is very sensitive to the deformation parameter λ, which can be discussed as follows
igure 13(a) and (b) visualize variations of the squeezing factors Sp and Sx in terms of gt for different values of
• F
the deformed parameter λ = −0.25, 0 and 5 when we choose the phase φ = 0 and π , respectively. These show
2
that the squeezing effect in the field operator p may be considerable for φ = 0 and small values of gt while
λ > 0. As seen in Fig. 13(a), the squeezing factor Sp tends to zero which indicates that the states Ψ(t) become
minimum uncertainty ones.
• For the case φ = 0, our calculations show that the squeezing factors Sp are really dependent on λ. Figure 13(a)
shows that, with a rise in λ, the degree of squeezing or depth of non-classicality increases at first and then
decreases when time goes on.
• Squeezing in the p quadrature disappears when φ reaches π , where squeezing in the x quadrature is raised (see
2
Fig. 13(b)).
Conclusions
A parity deformed Jaynes-Cummings Hamiltonian in terms of spin and λ-deformed bosonic operators, was
introduced. Its eigen-states and eigenvalues were obtained explicitly. Mathematical and physical implications and
applications of our results were discussed in detail. The deformed JCM introduced here may add new insights to
nonclassical states of radiation in cavity QED. It, also, can be used to further investigate the interaction between
an atomic system and a single mode of an electromagnetic field, including damping or amplifying processes,
which are of fundamental importance, in quantum optics. By assuming that the atom is initially prepared in the
excited state and the field is in the WCS, its quantum dynamical features such atomic inversion, quantum statistics
and squeezing of obtained wave functions of the system were investigated. It was found that the atomic inver-
sion exhibits Rabi oscillations including quasi-periodic behavior. Further examination of non-classical signs of
the atom-field states, revealed that a significant squeezing can be achieved for positive deformation parameters.
Furthermore, increasing the deformation parameter (the stronger external field) changes their statistics from
the Poissonian to super-Poissonian. The λ-deformed JCM, driven JCM, can be applied to generate maximally
entangled states. In other words, the small detuning and coupling regimes with a large deformation parameter
may lead to a long-lasting robust maximally entangled quantum state. It was illustrated that for large values of λ,
the generated maximally entangled quantum state was preserved as time goes on, despite the presence of the dis-
sipation. It is worth mentioning that the approach presented here can be potentially compared with some others,
already discussed in the literature, where other researchers tried to generate and stabilize the entanglement. In
other words, we investigate how an appropriate choice of the external field allows one to control atom-field entan-
glement. Finally, a possible generalization to the three-level system can be discussed.
References
1. Jaynes, E. & Cummings, F. Comparison of quantum and semiclassical radiation theories with application to the beam maser. Proc.
IEEE 51, 89 (1963).
2. Narozhny, N. B., Sanchez-Mondragon, J. J. & Eberly, J. H. Coherence versus incoherence: Collapse and revival in a simple quantum
model. Phys. Rev. A 23, 236 (1981).
3. Agarwal, G. S. Vacuum-Field Rabi Splittings in Microwave Absorption by Rydberg Atoms in a Cavity. Phys. Rev. Lett. 53, 1732
(1984).
4. Cummings, F. W. Stimulated Emission of Radiation in a Single Mode. Phys. Rev 140, A 1051 (1965).
5. Phoenix, S. J. D. & Knight, P. L. Fluctuations and entropy in models of quantum optical resonance. Ann. Phys 186, 381 (1988).
6. Chang, D. E., Vuletic, V. & Lukin, M. D. Quantum nonlinear optics — photon by photon. Nat. Photon 8, 685 (2014).
7. Zhang, Y. et al. Photon Devil’s staircase: Photon long-range repulsive interaction in lattices of coupled resonators with Rydberg
atoms. Sci. Rep 5, 11510 (2015).
8. Pedernales, J. S. et al. Quantum Rabi Model with Trapped Ions. Sci. Rep 5, 15472 (2015).
9. Zhou, L. & Sheng, Y. B. Complete logic Bell-state analysis assisted with photonic Faraday rotation. Phys. Rev. A 92, 042314 (2015).
10. Forn-Diaz, P., Romero, G., Harmans, C. J. P. M., Solano, E. & Mooij, J. E. Broken selection rule in the quantum Rabi model. Sci. Rep
6, 26720 (2016).
11. Song et al. Heralded quantum repeater based on the scattering of photons off single emitters using parametric down-conversion
source. Sci. Rep 6, 28744 (2016).
12. Short, R. & Mandel, L. Observation of Sub-Poissonian Photon Statistics. Phys. Rev. Lett. 51, 384 (1983).
13. Lo Franco, R., Compagno, G., Messina, A. & Napoli, A. Bell’s inequality violation for entangled generalized Bernoulli states in two
spatially separate cavities, Phys. Rev. A 72, 053806 (2005).
14. Lo Franco, R., Compagno, G., Messina, A. & Napoli, A. Single-shot generation and detection of a two-photon generalized binomial
state in a cavity. Phys. Rev. A 74, 045803 (2006).
15. Lo Franco, R., Compagno, G., Messina, A. & Napoli, A. Generating and revealing a quantum superposition of electromagnetic-
field binomial states in a cavity. Phys. Rev. A 76, 011804(R) (2007).
16. Goy, P., Raimond, J. M., Gross, M. & Haroche, S. Observation of Cavity-Enhanced Single-Atom Spontaneous Emission. Phys. Rev.
Lett. 50, 1903 (1983).
17. Singh, S. Field statistics in some generalized Jaynes-Cummings models. Phys. Rev. A 25, 3206 (1982).
18. Tavis, M. & Cummings, F. W. N atoms interacting with a single mode radiation field. Phys. Rev 170, 379 (1968).
19. Sukumar, C. V. & Buck, B. Some soluble models for periodic decay and revival. J. Phys. A 17, 885 (1984).
20. Buzano, C., Rasetti, M. G. & Rastello, M. L. Dynamical Superalgebra of the “Dressed” Jaynes-Cummings Model. Phys. Rev. Lett 62,
137 (1989).
21. Chaichian, M., Ellinas, D. & Kulish, P. Quantum algebra as the dynamical symmetry of the deformed Jaynes-Cummings model.
Phys. Rev. Lett 65, 980 (1990).
22. de los Santos-Sanchez, O. & Recamier, J. The f-deformed Jaynes–Cummings model and its nonlinear coherent states. J. Phys. B: At.
Mol. Opt. Phys 45, 015502 (2012).
23. Aleixo, A. N. F., Balantenkin, A. B. & Candido Ribeiro, M. A. A generalized Jaynes-Cummings Hamiltonian and supersymmetric
shape invariance. J. Phys. A: Math. Gen 33, 3173 (2000).
24. Wigner, E. Do the Equations of Motion Determine the Quantum Mechanical Commutation Relations?. Phys. Rev 77, 711 (1950).
25. Green, H. S. A Generalized Method of Field Quantization. Phys. Rev. 90, 270 (1953).
26. Volkov, D. V. On the quantization of half-integer spin fields. Sov. Phys. JETP 9, 1107 (1959).
27. Volkov, D. V. SU(3)×SU(3) symmetry and the baryon meson coupling constants. Sov. Phys. JETP 11, 375 (1960).
28. Greenberg, O. W. Spin and Unitary Spin Independence in a Paraquark Model of Baryons and Mesons. Phys. Rev. Lett. 13, 598
(1964).
29. Greenberg, O. W. & Messiah. A. M. L. Selection Rules for Parafields and the Absence of Para Particles in Nature. Phys. Rev. B 138,
1155 (1965).
30. Govorkov, A. B. Parastatistics and Parafields. Theor. Math. Phys 54, 234 (1983).
31. Plyushchay, M. S. Deformed Heisenberg algebra with reflection. Nucl. Phys. B 491, 619 (1997).
32. Ohnuki, Y. & Kamefuchi, S. Quantum Field Theory and Parastatistics. University Press of Tokyo (1982).
33. Macfarlane, A. J. Generalized Oscillator Systems and Their Parabosonic Interpretation, in Proc. Inter. Workshop on Symmetry
Methods in Physics, ed. Sissakian, A. N., Pogosyan, G. S. & Vinitsky, S. I. (JINR, Dubna, 1994).
34. Macfarlane, A. J. Algebraic structure of parabose Fock space. I. The Green’s ansatz revisited. J. Math. Phys 35, 1054 (1994).
35. Jayaramants, J. & de Lima Rodrigues, R. The Wigner-Heisenberg algebra as an effective operator technique for simpler spectral
resolution of general oscillator-related potentials and the connection with the SUSYQM algebra. J. Phys. A: Math. Gen 23, 3123
(1990).
36. Meljanac, S., Milekovic, M. & Stojic, M. Permutation invariant algebras, a Fock space realization and the Calogero model. Eur.
Phys. J. C 24, 331 (2002).
37. Yang, L. M. A Note on the Quantum Rule of the Harmonic Oscillator. Phys. Rev 84, 788 (1951).
38. Polychronakos, A. P. Exchange operator formalism for integrable systems of particles. Phys. Rev. Lett 69, 703 (1992).
39. Brink, L., Hansson, T. H. & Vasiliev, M. A. Explicit solution to the N body Calogero problem. Phys. Lett. B 286, 109 (1992).
40. Brink, L., Hansson, T. H., Konstein, S. & Vasiliev, M. A. The Calogero model: Anyonic representation, fermionic extension and
supersymmetry. Nucl. Phys. B 401, 591 (1993).
41. Brzezinski, T., Egusquiza, I. L. & Macfarlane, A. J. Generalized harmonic oscillator systems and their Fock space description. Phys.
Lett. B 311, 202 (1993).
42. Plyushchay, M. S. Deformed Heisenberg algebra, fractional spin fields and supersymmetry without fermions. Ann. Phys. 245, 339
(1996).
43. Plyushchay, M. S. Minimal bosonization of supersymmetry. Mod. Phys. Lett. A 11, 397 (1996).
44. Plyushchay, M. S. Deformed Heisenberg algebra and fractional spin field in (2+1)-dimensions. Phys. Lett. B 320, 91 (1994).
45. Aglietti, U., Griguolo, L., Jackiw, R., Pi, S. Y. & Seminara, D. Anyons and Chiral Solitons on a Line. Phys. Rev. Lett 77, 4406 (1997).
46. Horváthy, P. A. & Plyushchay, M. S. Anyon wave equations and the noncommutative plane. Phys. Lett. B 595, 547 (2004).
47. de Matos Filho, R. L. & Vogel, W. Nonlinear coherent states. Phys. Rev. A 54, 4560 (1996).
48. Dehghani, A., Mojaveri, B., Shirin, S. & Saedi, M. Cat-states in the framework of Wigner–Heisenberg algebra. Ann. Phys 362, 659
(2015).
49. Mojaveri, B. & Dehghani, A. Even and odd Wigner negative binomial states: Nonclassical properties. Mod. Phys. Lett. A 30,
1550198 (2015).
50. Bellomo, B., Lo Franco, R. & Compagno, G. Non-Markovian Effects on the Dynamics of Entanglement. Phys. Rev. Lett 99, 160502
(2007).
51. Yu, T. & Eberly, J. H. Quantum Open System Theory: Bipartite Aspects. Phys. Rev. Lett 97, 140403 (2006).
52. Yu, T. & Eberly, J. H. Sudden Death of Entanglement. Science 323, 598 (2009).
53. Mazzola, L., Piilo, J. & Maniscalco, S. Sudden Transition between Classical and Quantum Decoherence. Phys. Rev. Lett 104, 200401
(2010).
54. Bateman, H.On Dissipative Systems and Related Variational Principles. Phys. Rev 38, 815 (1931).
55. Stuckens, C. & Kobe, D. H. Quantization of a particle with a force quadratic in the velocity. Phys. Rev. A 34, 3565 (1986).
56. Dutra, S. M. Cavity Quantum Electrodynamics: The Strange Theory of Light in a Box. John Wiley and Sons, New York (2005).
57. Soltani, M. et al. Control of entanglement between two dissipative non-interacting qubits in a common heat bath by a laser field.
Eur. Phys. J. D 67, 256 (2013).
58. Nourmandipour, A. & Tavassoly, M. K. A novel approach to entanglement dynamics of two two-level atoms interacting with
dissipative cavities. Eur. Phys. J. Plus 130, 148 (2015).
59. Collett, M. J. & Gardiner, C. W. Squeezing of intracavity and traveling-wave light fields produced in parametric amplification. Phys.
Rev. A 30, 1386 (1984).
60. Scully, M. O. & Zubairy, M. S. Quantum Optics. Cambridge. Univ. Press (2001).
61. Sayrin, C. et al. Real-time quantum feedback prepares and stabilizes photon number states. Nature 477, 73 (2011).
62. Vijay, R. et al. Stabilizing Rabi oscillations in a superconducting qubit using quantum feedback. Nature 490, 77 (2012).
63. Schindler, P. et al. Quantum simulation of dynamical maps with trapped ions. Nat. Phys 9, 361 (2013).
64. Misra, B. & Sudarshan, E. C. G. The Zeno’s paradox in quantum theory. J. Math. Phys 18, 756 (1977).
65. Koshino, K. K. & Shimizu, A. Quantum Zeno effect by general measurements. Phys. Rep 412, 191 (2005).
66. Facchi, P., Nakazato, H. & Pascazio, S. From the Quantum Zeno to the Inverse Quantum Zeno Effect. Phys. Rev. Lett. 86, 2699
(2001).
67. Facchi, P. et al. Control of decoherence: Analysis and comparison of three different strategies. Phys. Rev. A 71, 022302 (2005).
68. Nourmandipour, A. & Tavassoly, M. K. Dynamics and protecting of entanglement in two-level systems interacting with a
dissipative cavity: the Gardiner–Collett approach. J. Phys. B: At. Mol. Opt. Phys 48, 165502 (2015).
69. Post, H. R. Many-particles systems: II. Proc. Phys. Soc. (London). A 69, 936 (1956).
70. Kuramotos, Y. & Kato, Y. Dynamics of One-Dimensional Quantum Systems: Inverse-Square Interaction Models. Cambridge. Univ.
Press (2009).
71. Dong, S. H. Wave Equations in Higher Dimensions. Springer (2011).
72. Dong, S. H. Factorization Method in Quantum Mechanics. Springer (2007).
73. Landau, L. D. & Lifshitz, E. M. Quantum Mechanics, Non-relativistic Theory 3rd edition. Pergamon, Oxford (1977).
74. Sage, M. L. The vibrations and rotations of the pseudogaussian oscillator. Chem. Phys 87, 431 (1984).
75. Sage, M. L. & Goodisman, J. Improving on the conventional presentation of molecular vibrations: Advantages of the
pseudoharmonic potential and the direct construction of potential energy curves. Am. J. Phys 53, 350 (1985).
76. Hurley, J. One-dimensional three-body problem. J. Math. Phys 8, 813 (1967).
77. Calogero, F. Ground state of one-dimensional N body system. J. Math. Phys 10, 2197 (1969).
78. Calogero, F. Solution of the one-dimensional N body problems with quadratic and/or inversely quadratic pair potentials. J. Math.
Phys. 12, 419 (1971).
79. Camiz, P., Gerardi, A., Marchioro, C., Presutti, E. & Scacciatelli, E. Exact solution of a time-dependent quantal harmonic oscillator
with a singular perturbation. J. Math. Phys. 12, 2040 (1971).
80. Dodonov, V. V., Malkin, I. A. & Man’ko, V. I. Green function and excitation of a singular oscillator. Phys. Lett. A 39, 377 (1972).
81. Dong, S. H. & Ma, Z. Q. Algebraic Approach To The Pseudo-Harmonic Oscillator In 2D. Int. J. Mod. Phys. E 11, 155 (2002).
82. Bayfield, J. E. Quantum Evolution: An Introduction to Time-Dependent Quantum Mechanics (New York: Wiley) (1999).
83. Alsing, P. & Zubairy, M. S. Collapse and revivals in a two-photon absorption process. J. Opt. Soc. Am. B 4, 177 (1987).
84. Puri, R. R. & Bullough, R. K. Quantum electrodynamics of an atom making two-photon transitions in an ideal cavity. J. Opt. Soc.
Am. B 5, 2021 (1987).
85. Puri, R. R. & Agarwal, G. S. Collapse and revival phenomena in the Jaynes-Cummings model with cavity damping. Phys. Rev. A 33,
3610(R) (1986).
86. Einstein, A., Podolsky, B. & Rosen, N. Can quantum mechanical description of physical reality be considered complete?. Phys. Rev
47, 777 (1935).
87. Barenco, A., Deutsch, D., Ekert, A. & Jozsa, R. Conditional quantum dynamics and logic gates. Phys. Rev. Lett 74, 4083 (1983).
88. Bengtsson, I. & Zyczkowski, K. Geometry of quantum states: An Introduction to Quantum Entanglement. Cambridge. Univ. Press
(2006).
89. Zhou, L. & Sheng, Y. B. Detection of nonlocal atomic entanglement assisted by single photons. Phys. Rev. A 90, 024301 (2014).
90. Bennett, C. H. et al. Teleporting an unknown quantum state via dual classical and Einstein-Podolsky-Rosen channels. Phys. Rev.
Lett 70, 1895 (1993).
91. Li, X. et al. Quantum Dense Coding Exploiting a Bright Einstein-Podolsky-Rosen Beam. Phys. Rev. Lett 88, 047904 (2002).
92. Bennett, C. H. & Wiesner, S. J. Communication via one- and two-particle operators on Einstein-Podolsky-Rosen states. Phys. Rev.
Lett 69, 2881 (1992).
93. Ekert, A. K. Quantum cryptography based on Bell’s theorem. Phys. Rev. Lett 67, 6961 (1991).
94. Horodecki, R., Horodecki, P., Horodecki, M. & Horodecki, K. Quantum entanglement. Rev. Mod. Phys 81, 865 (2009).
95. Wootters, W. K. Entanglement of formation of an arbitrary state of two qubits. Phys. Rev. Lett 80, 2245 (1998).
96. Vidal, G. & Werner, R. F. Computable measure of entanglement. Phys. Rev. A 65, 032314 (2002).
97. Vedral, V., Plenio, M. B., Jacobs, K. & Knight, P. L. Statistical inference, distinguishability of quantum states, and quantum
entanglement. Phys. Rev. A 56, 4452 (1997).
98. Walls, D. F. Squeezed states of light. Nature 306, 141–146 (1983).
99. Wodkiewicz, K. & Eberly, J. H. Coherent states, squeezed fluctuations, and the SU(2) am SU(1,1) groups in quantum-optics
applications. J. Opt. Soc. Am. B 2, 458 (1985).
100. Buzek, V. SU(1, 1) squeezing of SU(1, 1) generalized coherent states. J. Mod. Opt. 37, 303 (1990).
Acknowledgements
The authors would like to thank Professor M. K. Tavassoly for useful discussions and the referees as well as R. Lo
Franco for their efficient and worthwhile suggestions to improve the presentation.
Author Contributions
A.D. conceived and designed the research. A.D. and B.M. performed the analysis and wrote the manuscript. S.S.
and S.A. prepared all the figures.
Additional Information
Competing financial interests: The authors declare no competing financial interests.
How to cite this article: Dehghani, A. et al. Parity Deformed Jaynes-Cummings Model: “Robust Maximally
Entangled States”. Sci. Rep. 6, 38069; doi: 10.1038/srep38069 (2016).
Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.
This work is licensed under a Creative Commons Attribution 4.0 International License. The images
or other third party material in this article are included in the article’s Creative Commons license,
unless indicated otherwise in the credit line; if the material is not included under the Creative Commons license,
users will need to obtain permission from the license holder to reproduce the material. To view a copy of this
license, visit http://creativecommons.org/licenses/by/4.0/