Coherent states on a circle: the Higgs-like approach
Ali Mahdifar1,2 , Ehsan Amooghorban3,4
arXiv:2109.02251v1 [quant-ph] 6 Sep 2021
1
2
Department of Physics, University of Isfahan, Hezar Jerib, Isfahan, 81746-73441, Iran.
Department of Physics, Quantum optics group, University of Isfahan, Hezar Jerib St.
Isfahan 81764-73441, Iran.
3
4
Department of Physics, Faculty of Science, Shahrekord University, Shahrekord,
88186-34141, Iran.
Nanotechnology Research Center, Shahrekord University, Shahrekord 88186-34141,
Iran.
Abstract
In this paper, the Higgs-like approach is used to analyze the quantum dynamics of a harmonic oscillator constrained on a circle. We obtain the Hamiltonian of this system as a function of the Cartesian coordinate of the tangent
line through the gnomonic projection and then quantize it in the standard
way. We then recast the Hamiltonian in a shape-invariant form and derive
the spectrum energy of the confined harmonic oscillator on the circle. With
help of the f-deformed oscillator algebra, we construct the coherent states
on the circle and investigate their quantum statistical properties. We find
that such states show nonclassical features like squeezing and sub-Poissonian
statistics even in small curvatures of the circle.
Keywords: Quantum Harmonic Oscillator on a Circle, Shape-Invariant
Hamiltonian, Coherent States, Quantum Statistical Properties.
PACS: 03.65.Fd, 42.50.Dv
Email addresses:
[email protected] (Ali Mahdifar1,2 ),
[email protected] (Ehsan Amooghorban3,4 )
Preprint submitted to Elsevier
September 7, 2021
1. Introduction
One of the fundamental problems in quantum mechanics is to find a compatible quantization procedure of a classical system. The canonical quantization is the usual well-known approach which is based on using the classical
form of observables as a real function in phase space and replacing the coordinates x and p with operators q̂ and p̂ in a symmetrical order. However,
this procedure works only for systems with phase space R2N , provided that
the quantization is carried out in Cartesian coordinates. In a non-Cartesian
coordinate, we can start from the Lagrangian formalism and obtain canonical
conjugate momenta of the corresponding dynamical variables of the system
to construct the Hamiltonian, and impose canonical commutation relations
between them. However, it raises a host of difficulties like a mismatch between the canonical momenta and the Noether momenta [1]. Instead, we
can consistently quantize the system in Cartesian coordinates and make the
change to curvilinear coordinates at the quantum level. In general, the canonical quantization becomes cumbersome when the classical coordinates and/or
their conjugate momenta do not vary over the full interval from −∞ to +∞.
There are various reasons why physicists try to develop the correct quantum mechanics in curved spaces. Despite extending our knowledge on quantum mechanics in more general mathematical frameworks, it provides the
capacity for application in many different fields of physics, such as the quantum Hall effect [2], the quantum dots [3, 4], and the coherent state quantization [5]. To the best of our knowledge, there are three known approaches
to study nonrelativistic quantum mechanics in the spaces of constant curvature: The first one is the Noether quantization which is concerned with the
identification of the Killing vector fields and the Noether momenta. In this
manner, one can construct the Hamiltonian of the system as a function of
Noether momenta, and then apply the quantization process to the components of these momenta [6, 7, 8]. The second method is the so-called thin
layer quantization [9, 10, 11, 12]. In this approach, the 2D curved surface is
embedded into the larger 3D Euclidean space, and the introduction of effective potential results in a dimensional reduction in the Schrodinger equation.
In the last approach, which is known as the Higgs method, dynamical symmetries of a system with a spherical geometry are extracted[13, 14]. In this
way, the Hamiltonian of the system is expressed in terms of the Casimir operators of the Lie algebra of such dynamical symmetries, and subsequently,
the energy eigenvalues are obtained through the eigenvalues of the Casimir
2
operators. In this contribution, we follow the Higgs-like approach.
On the other hand, the standard coherent states (CSs) [15] as well as the
generalized CSs associated with various algebras [16, 17, 18] play an important role in different fields of physics. Among them, the so-called nonlinear or
f-deformed CSs [19, 20] have attracted much interest due to their nonclassical
properties, such as quantum interference and amplitude squeezing [21, 22].
Such states indeed are the eigenstates of the annihilation operator of the
f-deformed oscillator [19, 20] can be realized in the center of mass motion
of a laser-driven trapped ion [22, 23], and in a micromaser under intensitydependent atom-field interaction [24]. It is well known, the resolution of
identity is the most important mathematical feature of the CSs, i.e., they
form an overcomplete set. Furthermore, the CSs include properties of the
space that the relevant system defined on it. Therefore, if we consider, for
example, the harmonic oscillator on a curved space and construct the associated CSs, they contain the curvature effects. Thus, employing the curvaturedependent CSs manifold makes it possible to investigate the curvature effects
on the physical phenomena within the framework of the CSs approach [25].
This can enable us to study the curvature effects of some physical structures
like nanostructures on their electronic and optical properties [26].
In Refs. [27] and [28], based on the Higgs model, the generalized (nonlinear) coherent states of a two-dimensional harmonic oscillator and a KeplerCoulomb potential on a spherical surface are constructed and some of their
quantum optical properties are studied. Later on, the thermal nonlinear coherent states on a sphere is presented in [29] and the relation of the nonclassical properties of the constructed CSs with the curvature and temperature
is investigated.
We have recently presented the classical and quantum-mechanical treatment of the Higgs model in the presence of dissipation using a continuum of
oscillators as a reservoir [30]. There, we showed that the transition probability between energy levels is dependent on the curvature of the physical space
and the dissipative effect, and significant probabilities of transition are only
possible if the transition and reservoir’s oscillators frequencies are almost in
resonance. We have then investigated the dynamics of both a free particle
and an isotropic harmonic oscillator constrained to move on a spheroidal
surface using two consecutive gnomonic projections [31]. We have obtained
the Hamiltonian of the systems as a function of the Cartesian coordinates of
the tangent plane and finally quantized them. We have shown that the effect
of nonsphericity of the surface appears like an effective potential. We found
3
that the deviation from the sphericity leads to the split of the energy levels
of the constrained oscillator on a sphere and lifts the degeneracy.
With the above background, the goal of this contribution is twofold. In
the first step, we intend to obtain the classical and quantum Hamiltonian
of a harmonic oscillator confined on a circle which is viewed as the simplest
curvature in geometry. To do this, we use the gnomonic projection, which
is a projection onto the tangent line from the center of the circle into the
embedding line. We first obtain the metric of the circle background in the
gnomonic coordinates and construct the Hamiltonian of the harmonic oscillator. We follow the canonical quantization method and derive the quantum
counterpart of the classical Hamiltonian. We then recast the Hamiltonian in
a shape-invariant form and compute the exact spectrum energy of the confined harmonic oscillator on a circle. We show that in the limit of the large
radius of the circle (straight line), it properly reduces to the well-known
energy spectrum of the conventional quantum harmonic oscillator. In the
second step, we obtain a deformation function for such oscillator utilizing
the f-deformed oscillator algebra. In the end, we construct the CSs on a
circle employing the nonlinear CSs approach and investigate their quantum
statistical properties.
This paper is organized as follows: In Sec. 2, by making use of the
gnomonic projection onto the tangent line, we obtain the Hamiltonian of
a harmonic oscillator confined on a circle. We subsequently quantize the
aforementioned Hamiltonian using a factorization method and calculate its
exact spectrum energy. In Sec. 3, we obtain a deformation function for the
oscillator on a circle. In Sec. 4, we construct the corresponding CSs and
examine their resolution of identity. The Sec. 5 is devoted to the study of
the quantum statistical properties of the constructed CSs. Especially, the influence of curvature of the circle on their nonclassical properties is analyzed.
Finally, the summary and concluding remarks are given in section 6.
2. Quantum harmonic oscillator on circle
In this section, we obtain the Hamiltonian of a harmonic oscillator confined on a circle. To this end, we obtain the metric of the circle background
in the gnomonic coordinates and then construct the Hamiltonian of the harmonic oscillator.
Let us designate the Cartesian coordinates of the circle background by
4
(q1 , q2 ), and assume that they satisfy the circle equation,
q12 + q22 = R2 ,
(1)
where, R is the radius of the circle. If we denote the Cartesian coordinate
Figure 1: Coordinate systems and projection from a circle with radius R onto a line. The
gnomonic projection of the point from the circle’s center onto the point x on the tangent
line.
of the tangent line to the circle by x (see Fig. 1), the relationship between
these two coordinates is given by
x
q1 = p
,
[1 + λx2 ]
1
q2 = p
,
(2)
λ[1 + λx2 ]
where, λ = R12 is the curvature of the circle. Accordingly, a point on the
circle can be represented as,
x 1
~r ≡ (q1 , q2 ) = ( , √ ),
Λ λΛ
where,
Λ=
√
1 + λx2 .
(3)
(4)
Now the differential of ~r is given by,
√
λx
1
d~r = ~rx dx = ( 3 , − 3 )dx.
Λ
Λ
5
(5)
Here, ~rx is the derivatives of ~r with respect to x. After a straightforward
calculation, we obtain the metric of the circle as,
1
λx2
1
.
ds2 = d~r · d~r = ( 6 + 6 )dx2 = 4 dx2 .
Λ
Λ
Λ
(6)
From equation (6), we get
ṡ2 =
ẋ2
.
Λ4
(7)
We can therefore write the Lagrangian of a harmonic oscillator on circular
background as,
1 ẋ2
. 1 2
L =
ṡ − VHO (x) =
− VHO (x),
2
2 Λ4
(8)
where VHO (x) is the potential of the confined oscillator on the circle in terms
of the tangent coordinate x. In what follows, we obtain the aforementioned
potential. Let s be the arc-length of the geodesics from the north pole of the
circle, ~r0 = (0, R), to the point ~r = (q1 , q2 ). Using Eq.(1), we have
s2 = q12 + q22 = g q12 .
(9)
Here, the one-dimensional metric tensor on the circle is given by
g =1+
q12
.
R2 − q12
(10)
Accordingly, the potential energy for a harmonic oscillator on the circular
curve can be written as (in this paper we put ~ = m = ω = 1):
1
1
q2
VHO (s) ≡ s2 = g q12 = (1 + 2 1 2 )q12 .
2
2
R − q1
(11)
Let us reconsider the gnomonic projection, which is the projection onto
the tangent line from the center of the circle. By making use of Eq. (2),
the points on the circle are expressed in terms of the coordinates of this
projection. Thus, the potential of the oscillator on the circle in terms of x is
given by:
1
x2
1
VHO (x) = (1 + λx2 )(
) = x2 .
(12)
2
2
1 + λx
2
6
Now, by using the Lagrangian (8) and the potential (12), we can calculate
the momentum as,
p=
∂L
ẋ
= 4.
∂ ẋ
Λ
(13)
With proper calculation, we get the classical Hamiltonian of the harmonic
oscillator constrained to the circular background as,
2 1
1
1
1
.
(1 + λx2 )p + x2 ≡ π 2 + x2 ,
Hcl = ẋp − L =
2
2
2
2
(14)
where π = (1 + λx2 )p. By replacing the classical momentum p by related
d
), we can obtain the quantum counterpart of the classical
operator (−i dx
Hamiltonian (14) as,
d
1
d
1
2
2
(1 + λx )(−i )(1 + λx )(−i ) + x2
H =
2
dx
dx
2
2
d
1
d
1
−(1 + λx2 )2 2 − 2λx(1 + λx2 )
+ x2 .
(15)
=
2
dx
dx
2
Now, for any real number γ we define the following linear operator:
1
1
2 d
A(γ) = √ [ıπ + γx] = √ (1 + λx ) + γx .
dx
2
2
(16)
Indeed, the adjoint of A(γ) is given by:
1
1
2 d
A (γ) = √ [−ıπ + γx] = √ −(1 + λx ) + γx .
dx
2
2
†
(17)
We can easily show that
H̃1 (γ) := A† (γ)A(γ)
(18)
2
d
d
1
−(1 + λx2 )2 2 − 2λx(1 + λx2 ) − γ + (γ 2 − λγ)x2 .
=
2
dx
dx
√
2
By choosing γ = λ+ 2λ +4 ≡ γ̃, for which (γ 2 − λγ) = 1, we can get the
following factorization form:
γ
H̃1 = A† A = H − ,
2
7
(19)
where H is the Hamiltonian of the quantum oscillator (15). On the other
hand, for the partner Hamiltonian H̃2 := AA† we find
H̃2 (γ) = A(γ)A† (γ)
(20)
2
1
d
d
=
−(1 + λx2 )2 2 − 2λx(1 + λx2 ) + γ + (γ 2 + λγ)x2 .
2
dx
dx
Now, after straightforward calculations, the following relationship is hold:
H̃2 (γ) = H̃1 (γ1 ) + R(γ1 ),
(21)
where γ1 = γ + λ and R(γ) = γ − λ2 . As it is seen, the Hamiltonian H̃1 (γ)
admitting a factorization form A† (γ)A(γ) in such a way that the partner
Hamiltonian H̃2 (γ) = A(γ)A† (γ) is of the form as H̃1 (γ) but for a different
value of the γ. In this case, it is usually said that there is shape invariance
and makes it possible to calculate exact spectrum of the Hamiltonian [32, 33].
The ground state |ψ0 (γ)i with zero energy is given by A(γ)|ψ0 (γ)i = 0.
Subsequently, by using Eq. (21) it is seen that |ψ0 (γ1 )i is an eigenstate of
H̃2 (γ) with the energy Ẽ1 = R(γ1 ):
H̃2 (γ)|ψ0 (γ1 )i = [H̃1 (γ1) + R(γ1 )] |ψ0 (γ1 )i = R(γ1 )|ψ0 (γ1 )i.
(22)
Similarity, using Eq. (18) we have
H̃1 (γ)[A† (γ)|ψ0 (γ1 )i] = A† (γ)H̃2 (γ)|ψ0 (γ1 )i
= A† (γ)[H̃1 (γ1 ) + R(γ1 )] |ψ0 (γ1 )i
= R(γ1 )[A† (γ)|ψ0 (γ1 )i].
(23)
In other words, A† (γ)|ψ0 (γ1 )i is the first excited state of H̃1 (γ) with the
energy Ẽ1 = R(γ1 ). We can iterate this process and obtain the energy
eigenvalues of H̃1 (γ) as following:
Ẽn =
n
X
k=1
λ
R(γ + kλ) = n(γ + n).
2
√
2
(24)
Finally, by substituting γ = γ̃ = λ+ 2λ +4 and using Eq. (19), we can obtain
the eigenvalues En of the harmonic oscillator on the circle as
γ̃
1
λ
En = Ẽn + = γ̃(n + ) + n2 .
(25)
2
2
2
It is obvious that in the flat limit λ → 0, γ̃ → 1 and Eq. (25) reduces to
the energy eigenvalues of the quantum harmonic oscillator on a flat line, i.e,
En(λ=0) = n + 21 .
8
3. Circular oscillator Hamiltonian as an f -deformed quantum oscillator
The f -deformed quantum oscillators [20], as the nonlinear oscillators are
characterized by the following deformed dynamical variables  and †
 = âf (n̂) = f (n̂ + 1)â,
† = f (n̂)↠= ↠f (n̂ + 1),
(26)
where â and ↠are usual bosonic annihilation and creation operators. The
deformation function f (n̂) is an operator-valued function of the number operator n̂ (= ↠â) and the nonlinear properties of this system are governed
by it. As usual, without loss of generality, we choose f (n) = f † (n). From
equation (26), it follows that the f -deformed operators Â, † and n̂ satisfy
the following closed algebra
[Â, † ] = (n̂ + 1)f 2 (n̂ + 1) − n̂f 2 (n̂),
[n̂, Â] = −Â,
[n̂, † ] = † .
(27)
The above-mentioned algebra represents a deformed Weyl-Heisenberg algebra
whose nature depends on the nonlinear deformation function f (n̂). An f deformed oscillator is a nonlinear system characterized by a Hamiltonian of
the harmonic oscillator form
1
Ĥ = (†  + † ).
2
(28)
Using Eq. (26) and the number state representation n̂|ni = n|ni, the eigenvalues of the Hamiltonian (28) can be written as
1
En = [(n + 1)f 2(n + 1) + nf 2 (n)].
2
(29)
It is worth noting that in the limiting case f (n) → 1, the deformed algebra
(27) and the deformed energy eigenvalues (29), respectively, will reduce to
the standard Weyl-Heisenberg algebra and the harmonic oscillator spectrum.
Comparing the energy spectrum of the harmonic oscillator on a circular
background, Eq. (25), with the energy spectrum of an f -deformed oscillator,
Eq. (29), we obtain the corresponding deformation function for the oscillator
on the circle as
1/2
n̂ − 1
f (n̂) = γ̃ + (
)λ
.
(30)
2
9
Furthermore, the annihilation and creation operators of the harmonic oscillator on the circle can be written in terms of the conventional operators â
and ↠as follows
1/2
1/2
n̂ − 1
n̂ − 1
†
 = â γ̃ + (
)λ
)λ
,
 = γ̃ + (
↠.
(31)
2
2
These two operators satisfy the deformed Weyl-Heisenberg commutation relation
[Â, † ] = n̂λ + γ̃,
(32)
and act upon the quantum number states |ni, corresponding to the energy
eigenvalues (25), as
√
Â|ni = f (n) n|n − 1i,
√
† |ni = f (n + 1) n + 1|n + 1i.
(33)
4. Nonlinear CSs on a circle
We can construct CSs for the f -deformed oscillator similar to those of
the harmonic oscillator. The nonlinear transformation of the creation and
annihilation operators naturally provides the notion of nonlinear CSs or f coherent states. These states as eigenstates of the f -deformed annihilation
operator defined [20],
Â|z, f i = z|z, f i
(34)
From Eq. (26), we can obtain an explicit form of the nonlinear coherent states
on a circle (NCSsOC) in the number state representation as,
|z, f i = N (|z|2 )−1/2
∞
X
n=0
√
zn
n!f (n)!
|ni,
(35)
where by definition f (0)! = 1 and
f (n)! = f (n)f (n − 1) · · · f (1),
(36)
z is a complex number, and the normalization constant N is given by
N (|z|2 ) =
∞
X
n=0
10
|z|2n
.
n![f (n)!]2
(37)
Here, the deformation function corresponding to the oscillator on a circle
f (n)!, using Eq. (30) and after some calculations, is given by
λ Γ(β + n)
f (n)! = ( )n
,
2
Γ(β)
(38)
where β = 2γ̃
+ 1 and Γ is the gamma function. Similarly, we can define
λ
the f-deformed CSs corresponding to the harmonic oscillator on the circle as
below:
∞
X
zn
2 −1/2
p
|z, f icircle ≡ |ziλ = N (|z| )
|ni,
(39)
ρ(n)
n=0
h
i2
λ 2n Γ(β+n)
. It is worth noting that to have states
where ρ(n) = n! ( 2 )
Γ(β)
belonging to the Fock space, is required that 0 < N (|z|2 ) < ∞, which implies
that |z| ≤ limn→∞ n[f (n)]2 = ∞ , where in the last equality we have used
Eq. (30). Moreover, it is obvious that in the flat limit, λ → 0, ρ(n) → n!
and the above deformed CSs reduce to the standard CSs:
∞
n
|z|2 X z
− 2
√ |ni.
(40)
|zi = e
n!
n=0
4.1. Resolution of identity
In this subsection, let us show the constructed NCSsOC form an overcomplete set. In other words, the following resolution of identity has to be
satisfied
Z
N
X
d2 z |ziλ W (|z|2 ) λ hz| =
|nihn| = I.
(41)
n=0
The resolution of identity can be achieved by finding a measure function
W (|z|2 ). For this purpose, we substitute |ziλ from the Eq. (39) into Eq. (41),
and yields
Z
Z
∞
X
W (|z|2 )
|nihn|
2
2
d z |ziλ W (|z| ) λ hz| = π
d(|z|2 )|z|2n
,
(42)
2)
ρ(n)
N
(|z|
n=0
where we have used: z = |z|eiθ and d2 z = 12 d(|z|2 )dθ. By using the change
(|z|2 )
, we have:
of variable x = |z|2 and considering w(x) = π W
N (|z|2 )
Z ∞
1 Γ(n + 1) Γ2 (β + n)
w(x)xn dx = ρ(n) = 2 2 n
.
(43)
Γ2 (β)
[( λ ) ]
0
11
The above integral is called the moment problem and well-known mathematical methods such as Mellin transformations can be used to solve it [34, 35].
From definition the of Meijers G-function, it follows that,
Z
a1 , · · · , an , an+1 , · · · , ap
k−1 m,n
dx x Gp,q αx
b1 , · · · , bm , bm+1 , · · · , bq
Qn
Qm
1
j=1 Γ(1 − aj − k)
j=1 Γ(bj + k)
Qp
= k Qq
.
(44)
α
j=m+1 Γ(1 − bj − k) j=n+1 Γ(aj + k)
Comparing Eq. (43) with Eq. (44), we find that the measure function can be
written as,
2
4x
4
0
3,0
.
(45)
G
w(x) =
0, β − 1, β − 1
λ2 Γ2 (β) 0,3 λ
In this manner, it is seen that the NCSsOC satisfy the resolution of identity
and consequently form an overcomplete set.
5. Quantum statistical properties of the NCSsOC
In this section, we turn to investigate some quantum optics features of the
NCSsOC, such as probability of finding n quanta, mean number of photons,
the Mandel parameter and quadrature squeezing.
5.1. Photons-number distribution
Using Eq.(27), the mean number of photons in the NCSsOC is obtained
as follows:
N
X
†
np(n),
(46)
hn̂i = λ hz|â â|ziλ =
n=0
where the probability of finding n photon in the NCSsOC is given by
p(n) =
|z|2n
.
N (|z|2 )ρ(n)
(47)
As it is clear from the above equation, it is difficult to predict the results
analytically. Thus, in Fig. 2 we show the effect of the parameters z and λ on
the probability of finding n photons and also the mean number of photons
in the NCSsOC.
12
Figure 2: The probability of finding n photon in the NCSsOC as a function of λ and z
(a,b). Here, z = 3 and λ = 0.5 in panels (a) and (b), respectively. The mean number of
photons in the NCSsOC versus λ and z (c,d).
In Fig. 2(a) the probability p(n) of the NCSsOC for n > 1 is characterized
by a fast initial increase followed by a very slow decrease and tends to zero,
while the p(0) always enhances with increasing λ and eventually approaches
unity, as expected from the conservation of probability. This behavior is also
true for all other values of z. Moreover, the probability of finding n photon for
the large value of n is less than the NCSsOC with small photons. In Fig. 2(b),
the probability p(n) for n > 1 show the same trend in Fig. 2(b). However,
the probability p(0) decreases from its initial unit-value with increasing z
and eventually approaches zero. To get further insight, let us consider the
limiting case λ → 0. From Eq. (47), we find that the probability p(n) tends
to a Poissonian distribution, and the NCSsOC approaches the standard CSs.
The mean number of photons in the NCSsOC as a function of λ and z
are shown in Fig. 2(c) and 2(d), respectively. These results show that the
highest value of the photons mean number occurs at the large (small) value
of z (λ). In other words, the mean number of photons in the NCSsOC is not
13
Figure 3: The Mandel parameter of the NCSsOC as a function of λ and z (a,b). (c) The
density plot of Mandel parameter for the NCSsOC versus z and λ.
significant when both parameters z and λ are simultaneously small or large.
5.2. Mandel parameter
In this subsection, we investigate deviation from the Poisson distribution for the NCSsOC using the Mandel parameter. This parameter is given
by [36]:
(∆n)2 − hni
Q=
,
(48)
hni
where the positive, zero, and negative values represent super-Poissonian,
Poissonian, and sub-Poissonian distribution, respectively. Due to the complexity of the final form of the Mandel parameter for the NCSsOC, we do
not attempt to obtain the analytic form. Instead, we numerically investigate
the Mandel parameter for such states using Eqs. (46) and (47). In Fig. 3, we
14
have plotted the Mandel parameter of the NCSsOC as a function of λ and
z. The results show that, for some fixed values of z = 0.1, 1, 2 and 3, the
Mandel parameter of the NCSsOC first becomes more sub-Poissonian with
increasing λ and then tends to zero at large values of λ. Furthermore, this
parameter is more negative with increasing z (see Fig. 3 (a)). Whereas, for
some fixed values of λ = 0.01, 1, 2 and 3, the Mandel parameter becomes
more negative with increasing z and then tends to −0.6 (−1) at large (small)
values of λ, as seen in Fig. 3 (b). We can justify this result by the fact that
the NCSsOC behaves as the standard CSs at small λ and when the parameter z is large enough. Indeed, the NCSsOC is the statndard CSs at exactly
zero curvature, as shown in Eq. (40). In Fig. 3 (c), we observe that at the
small value of λ the most negative amount of the Mandel parameter occurs
in the NCSsOC with the small z values.
5.3. Quadrature squeezing
In this subsection, we consider the quadrature operators X̂1 and X̂2 defined in terms of creation and annihilation operators â and ↠as follows :
X̂1 = 21 âeiφ + ↠e−iφ ,
(49)
X̂2 = 2i1 âeiφ − ↠e−iφ .
By using the commutation relation of â and ↠, the following uncertainty
relation is obtained
iE 2
1 Dh
1
(∆X1 )2 (∆X2 )2 ≥
X̂1 , X̂2
= .
(50)
16
16
As is known, the quadrature squeezing occurs if we have (∆Xi )2 < 1/4(i =
1or2) or equally Si ≡ 4(∆Xi )2 − 1 < 0. Employing Eqs. (46) and (49), after
some algebra, the squeezing parameter S1 is obtained as
!2
p
√
∞
∞
2 cos 2φ X z 2n n(n − 1)
z 2n n
2 cos φ X
p
p
S1 = 2
−
z N (z 2 ) n=2 ρ(n − 2)ρ(n)
zN (z 2 ) n=1 ρ(n − 1)ρ(n)
+2hn̂i.
(51)
Here, z is assumed to be a real parameter, and the mean number of photons,
hn̂i, is already defined in Eq. (46). Figs. 4(a) and 4(b) display the squeezing
parameter S1 for the NCSsOC as a function of ϕ for z = 1 and λ = 0.5,
respectively. The red solid, dashed blue, dashdotted orange and dotted green
15
Figure 4: The squeezing parameter S1 versus φ for (a) z = 1, and (b) λ = 0.5. Here,
solid red, dashed blue, dashdotted orange and dotted green lines correspond to (a) ((b))
λ = 0.01 (z = 0.1), λ = 0.5 (z = 0.5), λ = 1 (z = 1) and λ = 1.5 (z = 1.5), respectively.
The density plots of (c) S1 , and (d) S2 as functions of z and λ for φ = π/6.
lines depict numerical results related to λ = 0.01, λ = 0.5, λ = 1 and λ = 1.5
for Fig. 4(a), and z = 0.1, z = 0.5, z = 1 and z = 1.5 for Fig. 4(b),
respectively. These figures show a periodic dependant on φ, and in areas
that the NCSsOC is squeezed the amount of squeezing always enhances with
increasing z (see Fig. 4(b)). While, for the fixed value z = 1, the squeezing
first enhances with increasing λ and then leads to decreasing squeezing, as
seen in Fig. 4(a)). This behavior can be easily understood from Fig. 4(c)
that at the small value of λ the highest amount of the squeezing appears in
the NCSsOC with small z, which is also in agreement with the results of the
Mandel parameter. At λ = 0, the squeezing properly vanished, as expected
from the standard CSs. It is worth noting that the squeezing parameter S2
can be obtained from Eq. (51) by replacing φ = φ + π/2. In this manner, the
16
squeezing parameter S2 is plotted for different values of λ and z in Fig. 4(d).
This plot clearly exhibits the same behavior of the squeezing parameter S1
with φ = 5π/6 (not shown here).
6. Summary and Concluding Remarks
In this paper, we have obtained the Hamiltonian of a harmonic oscillator
confined on a circle using the gnomonic projection. It is shown that the algebra of such harmonic oscillator can be regarded as f -deformed harmonic
oscillator algebra. We have constructed the nonlinear CSs for such harmonic
oscillator so that the NCSsOC properly tends to the standard CSs at zero
curvature. We then studied the quantum statistical properties of the NCSsOC, and found that the nonclassical features of these states enhance with
increasing the parameter z even in small curvatures of the circle.
7. Acknowledgment
A.M. wishes to thank The Office of Graduate Studies and Research Vice
President of The University of Isfahan for their support. E.A. also wishes to
thank the Shahrekord University for their support.
References
[1] R. Shankar, Principles of Quantum Mechanics, Springer, 1994.
[2] P. F. Bracken, Int. J. Theor. Phys. 46, 116 (2007).
[3] V. V. Gritsev and Y. A. Kurochkin, Phys. Rev. B 64, 035308 (2001).
[4] D. V. Bulaev, V. A. Geyler, and V. A. Margulis, Phys. Rev. B 69,
195313 (2004).
[5] J. P. Gazeau and W. Piechocki, J. Phys. A: Math. Gen. 37, 6977 (2004).
[6] J. F. Carinena, M. F. Ranada, and M. Santander, Ann. Phys. 322, 2249
(2007).
[7] J. F. Carinena, M. F. Ranada, and M. Santander, J. Math. Phys. 52,
072104 (2011).
17
[8] P. Bracken, J. Math. Phys. 55, 102102 (2014).
[9] H. Jensen and H. Koppe, Ann. Phys. 63, 586 (1971).
[10] R. C. T. da Costa, Phy. Rev. A 23, 1982 (1981).
[11] R. C. T. da Costa, Phys. Rev. A 25, 2893 (1982).
[12] Sh. Dehdashti, R. Roknizadeh, A. Mahdifar, and H. Chen H, Int. J.
Theor. Phys. 55, 124 (2016).
[13] P. W. Higgs, J. Phys. A: Math. Gen. 12, 309 (1979).
[14] H. I. Leemon, J. Phys. A: Math. Gen. 12, 489 (1979).
[15] R. J. Glauber, Phys. Rev. 136, 2529 (1963); R. J. Glauber, Phys. Rev.
131, 2766 (1963); R. J. Glauber, Phys. Rev. Lett. 10, 84 (1963).
[16] A.P. Perelomov, Generalized Coherent States and their Applications,
Springer, 1989.
[17] S. T. Ali, J.P. Antoine, J. P. Gazeau, Coherent States, Wavelets and
their Generalizations, Springer, 2000.
[18] J. R. Klauder, B. S. Skagerstam, Coherent States, Applications in
Physics and Mathematical Physics, World Scientific, 1985.
[19] A. I. Solomon, Phys. Lett. A 196, 29 (1994); J. Katriel and A. I.
Solomon, Phys. Rev. A 49, 5149 (1994); P. Shanta, S. Chaturvedi,
V. Srinivasan, and R. Jagannathan, J. Phys. A 27, 6433 (1994).
[20] V. I. Man’ko, G. Marmo, E. C. G. Sudarshan, and F. Zaccaria, Phys.
Scr. 55, 528 (1997).
[21] W. Vogel, R.L. de Matos Filho, Phys. Rev. A 52 4214 (1995).
[22] R. L. de Matos Filho, W. Vogel, Phys. Rev. A 54 4560 (1996).
[23] A. Mahdifar, W. Vogel, Th. Richter, R. Roknizadeh, and M. H. Naderi,
Phys. Rev. A 78, 063814 (2008).
[24] M. H. Naderi, M. Soltanolkotabi, and R. Roknizadeh, Eur. Phys. J. D
32, 397 (2005).
18
[25] A. Mahdifar, R. Roknizadeh and M. H. Naderi, Int. J. of Geom. Meth.
in Mod. Phys. 9, 1250009 (2012).
[26] V. N. Popov, New J. Phys. 6, 17 (2004).
[27] A. Mahdifar, R. Roknizadeh, M.H. Naderi, J. Phys. A 39, 70037014
(2006).
[28] A. Mahdifar, T. Hoseinzadeh, M. Bagheri Harouni, Annals of Physics
355, 21 (2015).
[29] H. Bagheri and A. Mahdifar 54, 052104 (2013).
[30] E. Amooghorban and A. Mahdifar, Annals of Physics 360, 237 (2015).
[31] A. Mahdifar and E. Amooghorban, J. Math. Phys. 60, 082106 (2019).
[32] B. K. Bagchi, Supersymmetry In Quantum and Classical Mechanics,
Chapman and Hall/CRC, 2000.
[33] J. F. Cariñena, M. F. Rañada, M. Santander, Rep. Math. Phys., 54,
285 (2004).
[34] J. R. Klauder, K. A. Peson, J. M. Sixdeniers, Phys. Rev. A 64, 013817
(2001).
[35] A. M. Mathai, R. K. Saxena, Generalized Hypergeometric Functions
with Applica-tions in Statistics and Physical Sciences, Springer-Verlag,
1973.
[36] L. Mandel and E. Wolf, Optical Coherence and Quantum Optics, Cambridge University Press, Cambridge, 1995.
19