Generalized squeezed states
Kevin Zelaya,1, 2, ∗ Sanjib Dey,3, † and Véronique Hussin2, 4, ‡
arXiv:1810.02947v1 [quant-ph] 6 Oct 2018
2
1
Physics Department, Cinvestav, AP 14-740, 07000 México City, Mexico
Centre de Recherches Mathématiques, Université de Montréal, Montréal–H3C 3J7, QC, Canada
3
Department of Physical Sciences, Indian Institute of Science Education and Research Mohali,
Sector 81, SAS Nagar, Manauli 140306, India
4
Départment de Mathématiques et de Statistique, Université de Montréal,
Montréal–H3C 3J7, QC, Canada
Squeezed states are one of the most useful quantum optical models having various applications
in different areas, especially in quantum information processing. Generalized squeezed states are
even more interesting since, sometimes, they provide additional degrees of freedom in the system.
However, they are very difficult to construct and, therefore, people explore such states for individual
setting and, thus, a generic analytical expression for generalized squeezed states is yet inadequate in
the literature. In this article, we propose a method for the generalization of such states, which can
be utilized to construct the squeezed states for any kind of quantum models. Our protocol works
accurately for the case of the trigonometric Rosen-Morse potential, which we have considered as
an example. Presumably, the scheme should also work for any other quantum mechanical model.
In order to verify our results, we have studied the nonclassicality of the given system using several
standard mechanisms. Among them, the Wigner function turns out to be the most challenging from
the computational point of view. We, thus, also explore a generalization of the Wigner function and
indicate how to compute it for a general system like the trigonometric Rosen-Morse potential with
a reduced computation time.
I.
INTRODUCTION
Squeezed states are interesting quantum optical systems
exhibiting nonclassical properties [1–3]. They produce
less noise in optical communication than a vacuum state.
This is why squeezed light has various applications in different areas including optical communications [4], optical
measurement [5], detection of gravitational waves [6, 7],
universal quantum computing [8], dense coding [9], etc.
Squeezed states are also utilized in quantum metrology
not only to improve the quantum metrology technique
itself [10–12], but also to increase the sensitivity of gravitational wave detectors [13–15], especially the LIGO [16].
Moreover, squeezed light serves as a primary resource in
continuous variable quantum information processing and,
is utilized to distribute secret keys in quantum cryptography [17]. For an extensive list of applications one may
refer, for instance [18, 19].
The primitive idea of squeezing follows from the Heisenberg uncertainty principle. It is well-known that the coherent states of light minimize the uncertainty relation
∆x∆p = 1/2, with both of the quadraturepuncertainties
being identical to each other ∆x = ∆p = 1/2 (~ = 1).
Consequently, the uncertainty region of a coherent state
can be represented by a circle in the optical phase space.
However, for some states the uncertainty circle may be
squeezed in one quadrature and elongated correspondingly
in the other so that the uncertainty circle is deformed to
form an ellipse and, the corresponding states are often
∗
†
‡
kdzelaya@fis.cinvestav.mx
dey@iisermohali.ac.in; sanjibdey4@gmail.com
veronique.hussin@umontreal.ca
familiar as squeezed states. Although, the uncertainty relation does not necessarily have to be minimized in the
latter case, however, there are rare examples for squeezed
states with minimum uncertainties [20, 21], which are popular as ideal squeezed states. Nevertheless, so far we have
discussed a particular type of squeezed states, namely the
quadrature squeezed states, which are defined as the states
whose standard deviation in one quadrature is less than
that of the coherent states or a vacuum state. Squeezing
can also occur in photon number distribution, and a state
is said to be number squeezed if the photon number uncertainty of the corresponding states becomes lower than
that of the coherent states. However, physically both of
the scenarios refer to the notion of nonclassicality.
Coherent states are not nonclassical, in fact, they are
the most classical analogue of quantum systems. But,
the statement is true only for the coherent states of the
harmonic oscillator, which are sometimes referred to the
Glauber coherent states. However, there are various generalized coherent states [22–29] for which the quadrature
and/or number squeezing occur and, thus, they are nonclassical. People sometimes refer these types of coherent
states or any other nonclassical states to be squeezed states
for which the quadrature and/or number squeezing occurs.
It should be noted that in this article we do not refer these
types of nonclassical states to be squeezed states, rather,
we talk about a particular class of states as defined in Sec.
II, which are constructed in such a way that the quadrature squeezing is inherited to them by construction and,
thus, they are always nonclassical.
Generalization of different quantum optical systems
provides a deeper understanding, since sometimes it
brings additional degrees of freedom in the system so that
they can be applied more efficiently to physical models
[22]. The essence of the generalization of several non-
2
classical states; such as, cat states [30–33], photon-added
coherent states [34–37], pair-coherent states [38], binomial
states [39], etc., have been explored in various contexts.
Squeezed states are probably the most well-behaved nonclassical states which can be prepared more systematically
and elegantly in the laboratory [40]. However, as per our
knowledge, there is no particular form of the generalized
squeezed states available in the literature, because they
are extremely difficult to construct. In this article, we
propose a generalization to such states followed by an example of a general system, viz., the trigonometric RosenMorse potential, where we apply our proposal directly.
The reason behind choosing the Rosen-Morse potential for
our analysis is mainly because it is a widely used model
in optics, but mostly because the squeezed states for such
model have not been studied notably. Therefore, we have
the opportunity to explore in a two-fold way. Firstly, the
generalization of the squeezed states is verified via an example through a popular model. Secondly, at the same
time, we can shed light on the behavior of the RosenMorse squeezed states, which is inadequate in the literature. It should be noted that there are many articles
available, for instance [41–53], entitled by “generalized
squeezed states”, however, they contain either the “generalized coherent states” having the squeezing properties
or, they are any other type of nonclassical states whose
quadrature and/or photon number is/are squeezed. However, according to our knowledge there is no trace of the
generalization of the particular state that we discuss in
the following section.
In Sec. II, we discuss the detailed procedure for the
generalization of the squeezed states by introducing a set
of generalized ladder operators followed by an explicit
analytic solution of the generalized squeezed states. In
Sec. III, we apply the obtained general solution to a particular type of model, namely the Rosen-Morse potential.
Sec. IV is composed of the analysis of nonclassicality and
squeezing properties of the given example by means of the
analysis of quadrature squeezing, sub-Poissonian photon
statistics and Wigner distribution function. Finally, our
concluding remarks are stated in Sec. V.
II.
Squeezed states for harmonic oscillator |α, δiho are constructed by operating the displacement operator D(α) =
exp(αa† − α∗ a) on the squeezed vacuum S(δ)|0i [46]
1
S(δ) = e 2 (δa
† †
a −δ ∗ aa)
,
(1)
with α, δ ∈ C being displacement and squeezing parameters, respectively. Alternatively they can be formulated
by performing the Holstein-Primakoff/Bogoliubov transformation on S(δ) arising from the solution of the following eigenvalue equation [50, 54, 55]
(a + ξa† )|α, ξiho = α|α, ξiho , ξ =
with k(n) being an operator-valued function of the number operator n = a† a associated with generalized models. The generalized ladder operators (3) have been introduced long time back [27, 29, 31, 32] in order to generalize
various quantum optical models, especially coherent and
cat states. These generalizations are mostly familiar as
nonlinear generalization and they are well-established and
widely accepted in the community. The existence of such
states have also been found in many experiments using
Kerr type nonlinearity and nonlinear cavity [56–58]. For
further information in this regard one may follow some
review articles in the context [22, 59]. Note that, the generalized ladder operators (3) are given in such a form that
A† A behaves as the number operator of the generalized
system and, therefore, the function k(n) can be associated with the eigenvalues en of the model as follows
A† A|ni = k(n)|ni,
k(n) ∼ en ,
(4)
which holds in general for the function k(n). The appearance of additional constant terms in the eigenvalues can
be realized by rescaling the composite system of A and A†
correspondingly. Therefore, by computing the eigenvalues
of the system, one can construct the function k(n) and,
thus, various quantum optical states by using the nonlinear generalization procedure that is discussed above. Nevertheless, in order to solve the eigenvalue equation (2) in
the generalized scenario, let us first expand the squeezed
states in Fock basis
|α, ξi =
∞
X
1
J (α, ξ, n)
p
|ni ,
N (α, ξ) n=0
k(n)!
(5)
Qn
where k(n)! = i=1 k(i) and k(0) = 1. By inserting (5)
into (2) replaced by the generalized ladder operators (3),
we end up with a three term recurrence relation
J (α, ξ, n + 1) = α J (α, ξ, n) − ξ k(n) J (α, ξ, n − 1), (6)
GENERALIZATION
|α, δiho = D(α)S(δ)|0i,
which reduces to the coherent states for ξ = 0. Generalization are usually carried out [50, 60] by replacing the
usual ladder operators a, a† by the generalized ladder operators A, A† in (2)
p
p
A† |ni = k(n + 1)|n + 1i, A|ni = k(n)|n − 1i, (3)
δ
tanh(|δ|), |ξ| < 1,
|δ|
(2)
with J (α, ξ, 0) = 1 and J (α, ξ, 1) = α, which when solved
one obtains the explicit form of the squeezed states [60].
But, the recurrence relation (6) is extremely difficult to
solve for general k(n), in fact, there are only few examples
where it has been possible, that too for some particular
models [60, 61]. Notice that, Eq. (6) reduces to a simple
form for ξ = 0, which is the case of coherent states and,
indeed, the corresponding solution leads to the nonlinear
coherent states [29, 31, 62]
|αi =
∞
1 X αn
p
|ni ,
N (α) n=0 k(n)!
(7)
which are the generalized version of the Glauber coherent
states. Furthermore, in order to obtain the standard form
3
of the harmonic oscillator squeezed states, we must consider k(n) = n and, with this the recurrence relation (6)
is solved to obtain
n/2
∞
X
1
1
α
ξ
√
Hn ( √ )|ni , (8)
|α, ξiho =
N (α, ξ) n=0 n! 2
2ξ
where Hn (α) denote the complex Hermite polynomials.
We intend to provide a general solution applicable for any
model, which is obtained by the general solution of (6) as
follows
[n/2]
J (α, ξ, n) =
X
(−ξ)m αn−2m g(n, m),
A, A† defined by (3) along with
k(n) = En −E0 = n(n+2d+2) 1 +
b2
,
(d + 1)2 (n + d + 1)2
(13)
so that the operator H = A† A + E0 reproduces the spectrum (12) in the Fock basis, with E0 being the ground
state energy. Thus, it becomes straightforward to compute the squeezed states for the model by utilizing (9)
along with the k(n) (13). Let us now analyze their nonclassical behavior to ensure that the constructed squeezed
states are well-behaved.
(9)
m=0
IV.
NONCLASSICAL PROPERTIES
with
(n−1)
(n−2m+1) (n−2m+3) (n−2m+5)
g(n, m) =
X
j1 =1
X
j2 =j1 +2
X
j3 =j2 +2
···
X
µ,
jm =jm−1 +2
(10)
where µ = k(j1 )k(j2 ) · · · k(jm ) and g(n, 0) = 1. The notation [n] in (9) represents the Floor function, whose output
is an integer less than or equal to the corresponding real
number n. The solution (9) is achieved initially by computing the first few terms of the recurrence relation (6)
with initial conditions J (α, ξ, 0) = 1 and J (α, ξ, 1) = α.
Thereafter, a relatively closed form of the series is acquired by the trial and error method, which is never an elegant way to solve a recurrence relation. However, once the
solution is obtained it is straightforward to verify whether
the solution satisfies the recurrence relation. For the detailed proof one may refer to the Appendix. Eq. (9) along
with (10), thus, provides a powerful solution of (6), which
can be employed to obtain the squeezed states for any
model corresponding to the known eigenvalues k(n). Let
us now study how the general solution (9) works for a
given system and verify their nonclassical properties. The
main aim is to extract the factor k(n) from the expression of the eigenvalues of the given model and apply it
to the general solution (9) to obtain the squeezed states
corresponding to it.
III.
EXAMPLE: ROSEN-MORSE POTENTIAL
The trigonometric Rosen-Morse potential [63, 64]
V (x) = −2b cot x + d(d + 1) csc2 x,
∀x ∈ [0, π],
(11)
is one of the well-studied models and the corresponding
energy eigenvalues are well-known [64]
En = (n + d + 1)2 −
b2
,
(n + d + 1)2
n = 0, 1, 2, ...., (12)
with b, d being positive constants. While the model can be
solved by any standard method available in the literature,
we consider a particular procedure suitable for our purpose. We can construct the generalized ladder operators
The behavior of the coherent states being analogous to
that of the classical objects, they are often familiar as
classical like states, although originally they are the superposition of large number of quantum states. According
to the convention of Glauber and Sudarshan [65, 66], the
quantum states which are less classical than the coherent states can be called nonclassical states. Technically
they can be realized in terms of the Glauber-Sudarshan’s
P -function for arbitrary density matrices
Z
ρ = d Reα d Imα P (α)|αihα|,
(14)
R
with d Reα d Imα P (α) = 1. In case of coherent states,
the weight function P (α) is a delta function and, thus, the
function P (α) represents a probability density. On the
other hand, the states for which the P -distribution fails
to be a probability density are called nonclassical states.
More precisely, if the singularities of the P -functions are
either of types stronger than those of the delta functions
(e.g. derivatives of delta function) or they are negative,
the corresponding states have no classical analogue. For
further details in this regard one may refer, for instance
[22, 59]. The convention of nonclassicality in case of our
generalized systems is same as described above. The reason is that the generalization is carried out in such a way
that the generalized ladder operators A and A† (3) operate
on the Fock states |ni producing the generalized coherent
states (7) as well as the squeezed states (9) in the Fock basis. Therefore, the whole construction is in the optical basis dealing with the optical photon number a† a. The only
thing that remains is to understand the role of the extra
nonlinear factor k(n) that appears due to the generalization. Note that, the signature of k(n) underlies within
the generalized ladder operators A† and A, which constitute the commutation relation [Q, P ]. And this commutation relation [Q, P ] is the driving factor in the RHS of
the generalized uncertainty relation (15) as given in the
next section. Therefore, only a slight modification in the
statement of the quadrature squeezing may be sufficient
to understand the quadrature squeezing properties of our
system. In the studies of photon number squeezing and
Wigner function, we may not require any further modification to the standard definition, since the generalized
4
Figure 1. Quadrature squeezing for the Rosen-Morse squeezed states.
squeezed states are constructed in the standard optical
basis. However, a proper generalization of the Mandel
parameter could provide further deeper understanding on
the photon number squeezing of our systems, but it is not
a trivial work. At least, to our knowledge, a proper generalization of such thing is not yet adequate in the literature.
We have provided a more rigorous and technical analysis
in this regard in the respective places in the subsequent
sections. Let us now analyze some standard techniques to
realize the nonclassical properties of our system.
A.
Quadrature squeezing
The quadrature operators
as
√ for the generalized system
√
defined by Q = (A + A† )/ 2 and P = (A − A† )/ 2i obey
the generalized uncertainty relation
∆Q∆P ≥
1
hα, ξ|[Q, P ]|α, ξi .
2
(15)
The relation (15) holds for any arbitrary state |ψi in general, however, since our analysis is based on the squeezed
states |α, ξi, we have written it in a particular form in
terms of the squeezed states |α, ξi. For the case of the
vacuum state |0i, the RHS of (15) and square of each
of the variances become equal each other, i.e. (∆Q)2 =
(∆P )2 = 21 h0|[Q, P ]|0i = k(1)/2, and the same identity
holds for the generalized coherent states (7) also. Note
that, since [Q, P ] is not in general proportional to the
identity, the RHS of (15) strongly depends on the state
that is used to compute the expectation values. So, it can
not be guaranteed that the variances associated with the
generalized squeezed states are necessarily below to those
of the vacuum state |0i, as it occurs in the case of harmonic oscillator. Therefore, in our case, the quadrature
squeezing does not correspond to the case when the variance of any of the quadratures becomes lower than the
square root of the RHS of (15) for the vacuum state |0i.
But, it corresponds to that of the particular state that is
being studied, which is |α, ξi in our case. The behavior
is depicted in Fig. 1, where we plot the variance of the
quadratures for the squeezed states ∆Q, ∆P as well as for
the vacuum state (∆Q)0 , (∆P )0 along with the RHS of
(15) for different values of ξ. Panel (a) corresponds to the
case of coherent states, and we notice that ∆Q, ∆P and
RHS of (15) coincide. It means that the states belong to
the category of intelligent states [67] implying no quadrature squeezing with the variance of both of the quadratures being identical to each other. More interesting behavior is observed in the remaining two panels, in panel
(b), the Q quadrature is squeezed for ξ = 0.2, whereas
the P quadrature squeezes for ξ = −0.2 (see the subpanel
in panel (b)). The same holds true for panel (c), however, in this case the squeezing has increased compared to
(b), as expected. Interestingly, in both of the cases the
uncertainty relation (15) is minimized. This is actually
more exciting, since this is found very rarely in the literature, and is familiar as ideal squeezed states [20, 21]. The
other interesting observation from the panel (b) and (c)
is that the variances for the generalized squeezed states
∆Q, ∆P are not always below to those of the vacuum
state (∆Q)0 , (∆P )0 , however, they are always below the
generalized uncertainty minimum 21 |h[Q, P ]i||α,ξi . It confirms that the information of nonclassicality in case of the
generalized squeezed states must be extracted from the
minimum of the generalized uncertainty relation.
B.
Photon number squeezing
Photon number squeezing is another excellent method
by which the nonclassicality of a state can be tested. A
photon number squeezed state or a nonclassical state must
satisfy the relation ∆n)2 < hni, where n = a† a is the number operator for the generalized model. This implies that
for a squeezed/nonclassical state the Mandel parameter
Q [68]
Q=
(∆n)2
− 1,
hni
(16)
must be negative, so that the corresponding distribution is
sub-Poissonian. Clearly, Q = 0 corresponds to the Poissonian distribution and, thus, the number squeezing is
absent. Similarly, the super-Poissonian case Q > 0 is also
not important to us, since it does not identify any squeezing property. Note that, the same analysis is applicable
to our generalized squeezed states, since they have been
constructed in the Fock basis. However, for curiosity one
may ask a question what happens if we generalize the
definition of the Mandel parameter, say by defining the
5
is the quantum analogue of the phase space probability
density. Thus, any negative region in the Wigner function indicates that the state possesses nonclassicality [70].
However, in order to study the Wigner function in the
given scenario, we need to consider a construction for the
most general density matrix, in terms of the Fock basis
∞
X
ρ=
m1 ,m2 =0
Cm1 ,m2 |m1 ihm2 |,
(17)
with Cm1 ,m2 = hm1 |ρ|m2 i so that the corresponding
Wigner function [34, 70] becomes
W (z) =e
∞
X
2|z|2
m1 ,m2
Figure 2. Mandel parameter Q (16) for the Rosen-Morse
squeezed states for different values of ξ.
†
generalized number operator N = A A. Firstly, it may
be inappropriate in our case to deal with the generalized
Mandel parameter, since we are effectively dealing in the
Fock basis and not in the generalized basis. But, most
importantly, even if we construct a generalized Mandel
parameter as Qg = [(∆N )2 /hN i] − 1, the Poissonian distribution does not correspond to the case of Qg = 0. This
is because the Mandel parameter in the generalized coherent state basis takes the form Qg = h[A, A† ]i − 1, and it
is not guaranteed that [A, A† ] = 1 is fulfilled in general.
Therefore, the information of photon number squeezing
can not be extracted from the Qg measured in the generalized coherent state basis. A similar kind of observation
was made earlier by two of the authors of this article in different studies [33, 69]. However, a proper understanding
of the generalization of the definition of Mandel parameter is yet underway. Nevertheless, for our purpose we
may not require that analysis, or if it requires, we have
to keep it as an open problem, which is beyond our scope
in the present article. The behavior of the Mandel parameter for different values of the squeezing parameter
for the trigonometric Rosen-Morse system is depicted in
Fig. 2. Notice that the Mandel parameter becomes negative in some region in all of the cases, including the case
corresponding to the coherent states, ξ = 0. This is not
surprising since the generalized coherent states are sometimes known to be slightly nonclassical as discussed in the
introduction. Nevertheless, upon increasing the squeezing
parameter ξ, the nonclassicality increases accordingly as
expected. Note that, the subpanel in Fig. 2 demonstrates
the behavior of the same functions as in the main panel,
but for the higher value of α, where it is clearly visible
that the line corresponding to ξ = 0.6 moves below than
that of ξ = 0.4, thus, showing a consistent behavior.
C.
Wigner quasi-probability distribution
The nonclassicality can also be described by the study
of the Wigner quasi-probability distribution [70], which
×
Z
C
√ m1 ,m2
m1 !m2 !
=0
(18)
∗
∗
d2 β −|β|2
e
(−β ∗ )m1 (β)m2 e2(zβ −z β) ,
π
where z, β are the eigenvalues of the Glauber coherent states.
Introducing a change of variable γ =
2z and, subsequently, using the identity G(γ) =
R d2 β −|β|2 γβ ∗ −γ ∗ β
2
e
= e−|γ| , the Wigner function (18)
π e
can be rewritten in the following form
W (γ) = e|γ|
2
/2
∞
X
m1 ,m2 =0
Cm1 ,m2 Fm1 ,m2 (γ),
m1 +m2
m1 +m2
(19)
∂
√
e−|γ| . By uswith Fm1 ,m2 (γ) = (−1)
m1 !m2 ! ∂γ m1 ∂γ ∗m2
ing the fact that the derivative of any analytic function
f (z, z ∗ ) with respect to z are independent of z ∗ and viceversa and, by employing the Rodrigues formula for the
associated Laguerre polynomials, it can be shown that
2
Fm1 ,m2 (z) =
(20)
q
(−1)m1 m1 ! e−4|z|2 (2z)m2 −m1 Lm2 −m1 (4|z|2 ), m2 ≥ m1
m1
q m2 !
(−1)m2 m2 ! e−4|z|2 (2z ∗ )m1 −m2 Lm1 −m2 (4|z|2 ), m2 ≤ m1 .
m2
m1 !
Here, we change the variable z = γ/2 again, so that
we come back to the expression in terms of the original variable. Note that the Wigner function given by
(19) along with (20) is completely general and can be
applied to any quantum state irrespective of pure or
mixed. However, it is slightly difficult to perform numerical computation with the expression we have, since
it involves four sums. In what follows, we shall split
the expression into some parts in order to reduce the
computation time. Notice that the expression inside the
double summation in (19) are the elements of an infinite dimensional matrix,
P∞therefore, we rewrite the Wigner
function as W (z) =
m1 ,m2 =0 Φm1 ,m2 , with Φm1 ,m2 =
2
e2|z| Cm1 ,m2 Fm1 ,m2 (z). Thus, one can decompose the
matrix W (z) in terms of the diagonal Φm,m , upper offdiagonal Φm1 ,m2 and lower off-diagonal Φm2 ,m1 elements
as
∞
∞
∞
X
X
X
(Φm1 ,m2 + Φm2 ,m1 ) .
Φm,m +
W (z) =
m=0
m1 =0 m2 =m1 +1
(21)
6
Figure 3. Wigner functions for the Rosen-Morse squeezed states for (a) ξ = 0 (b) ξ = 0.6 (c) α = 100, ξ = 0.8 (d) α = 100, ξ = 0.95.
The second term can be written in terms of the upper
off-diagonal elements only as
Φm1 ,m2 + Φm2 ,m1 = 2Re [Φm1 ,m2 ] , ∀ m2 > m1 .
(22)
By replacing (22) in (21), we obtain the final expression
of the Wigner function, which reduces the computation
time substantially. Let us now apply the scheme to the
squeezed states of our system and compute the Wigner
function numerically. The results of the computation for
the trigonometric Rosen-Morse case are demonstrated in
Fig. 3. Interestingly, we observe the same behavior as in
the case of the Mandel parameter, i.e. there are some very
small negative regions in panel (a) in spite of the squeezing
parameter ξ being zero as per our expectation. Although
the negative regions are not visible properly in the figure,
because they are dominated by the positive peaks and,
with certain rescaling they would have been visible. However, there is no way to test whether this negativity is
due to the nonclassical property of the generalized coherent states, or due to the truncation of the infinite series.
But, intuitively one may argue that it can cause because
of the combined effects of these two. It should not occur
only because of the truncation of the series, since in Fig. 2
we have noticed that the generalized coherent states may
also possess little amount of nonclassicality. In any case,
when we increase the squeezing parameter, as in panel
(b), the negativity increases, as well as the distribution
becomes squeezed in real z axis. The rest of the figures as
depicted in panel (c) and (d) are more interesting, since
the negative peaks become stronger as we increase the
squeezing parameter ξ and, thus, move towards the more
squeezed region.
V.
CONCLUDING REMARKS
We have proposed a scheme for the generalization of
squeezed states along with an example of the trigonometric Rosen-Morse potential on which our protocol has been
applied. Subsequently, we verify our results by studying the nonclassical properties of the system by utilizing
several standard techniques; such as, quadrature squeezing, photon number squeezing and Wigner function. All
of them agree that the states emerged out of the generalized system are always nonclassical and the degree of nonclassicality increases with the increase of the squeezing of
the corresponding system. Although, the proposed generalization contains slightly complicated expression due to
which all of the nonclassical properties have to be studied
numerically, however, it is conceivable that the generalization appears with consistent results without any irregularity or shortcomings. Moreover, the generalization comes
7
out with a relatively closed expression and, thus, it is legitimate that a further simplified version should follow up
in near future subjecting to more intense investigation.
odd powers of n, so that for even case we obtain
ξα2n−1 [k(2n) + g(2n, 1) − g(2n + 1, 1)]
+
Acknowledgements: K.Z. would like to thank V.H.
and Centre de Recherches Mathématiques for kind hospitality, also acknowledges the support of CONACyT scholarship 45454 and the FRQNT international internship
award 210974. S.D. acknowledges the INSPIRE Faculty
research Grant (DST/INSPIRE/04/2016/001391) by the
Department of Science and Technology, Govt. of India.
V.H. acknowledges the support of research grants from
CRSNG of Canada.
n−1
X
(24)
(−ξ)m−1 α2n−2m−1 [g(2n + 1, m + 1)
m=1
−g(2n, m + 1) − k(2n)g(2n − 1, m)] = 0,
and for the odd case it turns out to be
ξα2n [k(2n + 1) + g(2n + 1, 1) − g(2n + 2, 1)]
+ (−ξ)n [k(2n + 1)g(2n, n) − g(2n + 2, n + 1)]
+
n−1
X
(25)
(−ξ)m−1 α2n−2m [g(2n + 2, m + 1)
m=1
−g(2n + 1, m + 1) − k(2n + 1)g(2n, m)] = 0.
APPENDIX
This section contains a detailed proof of the solution
(9) of the recurrence relation (6). First, we substitute the
solution (9) into the recurrence relation (6), so that we
obtain
[ n+1
2 ]
X
(−ξ)m αn−2m+1 g(n + 1, m)
(23)
m=0
[n/2]
−
X
(−ξ)m αn−2m+1 g(n, m)
m=0
[ n−1
2 ]
+ k(n)
X
(−1)m ξ m+1 αn−2m−1 g(n − 1, m) = 0,
As per the requirement, all the coefficients of α in both
of the cases must vanish. Hence, we end up with some
simple relations. For the even case (24) results to
g(2n + 1, 1) = g(2n, 1) + k(2n),
(26)
g(2n + 1, m + 1) = g(2n, m + 1) + k(2n)g(2n − 1, m),
(27)
whereas for the odd case they become
g(2n + 2, n + 1) = k(2n + 1)g(2n, n),
g(2n + 2, 1) = g(2n + 1, 1) + k(2n + 1),
g(2n + 2, m + 1) = g(2n + 1, m + 1)
+ k(2n + 1)g(2n, m).
(28)
(29)
(30)
where g(n, m) is given by (10). Subsequently, we collect
the terms of same power of α and separate the even and
Now, (26), (29) and (27), (30) are the equivalent sets of
equations, apart from a change of parameter. Therefore,
effectively we are left with three identities (26), (27) and
(28) to prove, which are indeed straightforward by following the definition of g(n, m) from (10).
[1] D. F. Walls, Squeezed states of light, Nature 306, 141–146
(1983).
[2] R. Loudon and P. L. Knight, Squeezed light, J. Mod. Opt.
34, 709–759 (1987).
[3] M. C. Teich and B. E. A. Saleh, Squeezed state of light,
Quantum Opt.: J. Euro. Opt. Soc. Part B 1, 153 (1989).
[4] Y. Yamamoto and H. A. Haus, Preparation, measurement and information capacity of optical quantum states,
Rev. Mod. Phys. 58, 1001 (1986).
[5] C. M. Caves, Quantum-mechanical noise in an interferometer, Phys. Rev. D 23, 1693 (1981).
[6] H. Vahlbruch et al.,
Coherent control of vacuum
squeezing in the gravitational-wave detection band,
Phys. Rev. Lett. 97, 011101 (2006).
[7] S. S. Y. Chua et al.,
Quantum squeezed light in
gravitational-wave detectors, Class. Quantum Grav. 31,
183001 (2014).
[8] N. C. Menicucci et al.,
Universal quantum computation with continuous-variable cluster states,
Phys. Rev. Lett. 97, 110501 (2006).
[9] M. Ban, Quantum dense coding via a two-mode squeezedvacuum state, J. Opt. B: Quantum Semiclass. Opt. 1, L9
(1999).
[10] J. P. Dowling, Quantum optical metrology–the lowdown
on high-N00N states, Contemp. Phys. 49, 125–143 (2008).
[11] V. Giovannetti, S. Lloyd and L. Maccone, Advances in
quantum metrology, Nature Photonics 5, 222–229 (2011).
[12] M. Riedel et al., Atom-chip-based generation of entanglement for quantum metrology, Nature 464, 1170–1173
(2012).
[13] H. Vahlbruch et al., Quantum engineering of squeezed
states for quantum communication and metrology, New
J. Phys. 9, 371 (2007).
[14] R. Schnabel et al., Quantum metrology for gravitational
wave astronomy, Nature Commun. 1, 121 (2010).
[15] P. M. Anisimov et al., Quantum metrology with two-mode
squeezed vacuum: parity detection beats the Heisenberg
limit, Phys. Rev. Lett. 104, 103602 (2010).
m=0
8
[16] J. Aasi et al., Enhanced sensitivity of the LIGO gravitational wave detector by using squeezed states of light,
Nature Photonics 7, 613–619 (2013).
[17] M. Hillery, Quantum cryptography with squeezed states,
Phys. Rev. A 61, 022309 (2000).
[18] S. L. Braunstein and P. Van Loock, Quantum information with continuous variables, Rev. Mod. Phys. 77, 513
(2005).
[19] U. L. Andersen, G. Leuchs and C. Silberhorn, Continuousvariable quantum information processing, Laser Photon. Rev. 4, 337–354 (2010).
[20] Y. Yamamoto et al.,
Generation of numberphase minimum-uncertainty states and number states,
J. Opt. Soc. Am. B 4, 1645–1662 (1987).
[21] S. Dey, A. Fring and V. Hussin,
Nonclassicality versus entanglement in a noncommutative space,
Int. J. Mod. Phys. B 31, 1650248 (2017).
[22] V. V. Dodonov, Nonclassical states in quantum optics: a squeezed review of the first 75 years, J. Opt. B:
Quant. Semiclas. Opt. 4, R1 (2002).
[23] A. O. Barut and L. Girardello,
New “coherent”
states associated with non-compact groups,
Commun. Math. Phys. 21, 41–55 (1971).
[24] D. Stoler, Generalized coherent states, Phys. Rev. D 4,
2309 (1971).
[25] M. Arik and D. D. Coon, Hilbert spaces of analytic functions and generalized coherent states, J. Math. Phys. 17,
524–527 (1976).
[26] A. Perelomov, Generalized coherent states and their applications, Springer-Verlag: Berlin (1986).
[27] V. I. Man’ko, G. Marmo, E. C. G. Sudarshan and
F. Zaccaria, f -oscillators and nonlinear coherent states,
Phys. Scr. 55, 528 (1997).
[28] S. T. Ali, J. P. Antoine and J. P. Gazeau, Coherent states,
wavelets and their generalizations, Springer-Verlag: New
York (2000).
[29] S. Sivakumar, Studies on nonlinear coherent states,
J. Opt. B: Quantum Semiclas. Opt. 2, R61 (2000).
[30] Y. Xia and G. Guo, Nonclassical properties of even and
odd coherent states, Phys. Lett. A 136, 281–283 (1989).
[31] R. L. M. Filho and W. Vogel, Nonlinear coherent states,
Phys. Rev. A 54, 4560 (1996).
[32] S. Mancini, Even and odd nonlinear coherent states,
Phys. Lett. A 233, 291–296 (1997).
[33] S. Dey, q-deformed noncommutative cat states and their
nonclassical properties, Phys. Rev. D 91, 044024 (2015).
[34] G. S. Agarwal and K. Tara, Nonclassical properties of
states generated by the excitations on a coherent state,
Phys. Rev. A 43, 492 (1991).
[35] T. M. Duc and J. Noh, Higher-order properties of photonadded coherent states, Opt. Commun. 281, 2842–2848
(2008).
[36] O. Safaeian and M. K. Tavassoly, Deformed photon-added
nonlinear coherent states and their non-classical properties, J. Phys. A: Math. Theor. 44, 225301 (2011).
[37] S. Dey and V. Hussin, Noncommutative q-photon-added
coherent states, Phys. Rev. A 93, 053824 (2016).
[38] G. S. Agarwal and A. Biswas, Quantitative measures of
entanglement in pair-coherent states, J. Opt. B: Quantum
Semiclas. Opt. 7, 350 (2005).
[39] C. T. Lee, Photon antibunching in a free-electron laser,
Phys. Rev. A 31, 1213 (1985).
[40] D. J. Wineland, J. J. Bollinger, W. M. Itano and
D. J. Heinzen, Squeezed atomic states and projection
noise in spectroscopy, Phys. Rev. A 50, 67 (1994).
[41] M. V. Satyanarayana, Generalized coherent states and
generalized squeezed coherent states, Phys. Rev. D 32,
400 (1985).
[42] S. L. Braunstein and R. I. McLachlan, Generalized squeezing, Phys. Rev. A 35, 1659 (1987).
[43] X. Ma and W. Rhodes, Multimode squeeze operators and
squeezed states, Phys. Rev. A 41, 4625 (1990).
[44] J. Katriel and A. I. Solomon, Generalized q-bosons and
their squeezed states, J. Phys. A: Math. Gen. 24, 2093
(1991).
[45] C. F. Lo and R. Sollie, Generalized multimode squeezed
states, Phys. Rev. A 47, 733 (1993).
[46] M. M. Nieto and D. R. Truax, Squeezed states for general
systems, Phys. Rev. Lett. 71, 2843 (1993).
[47] S. Seshadri, S. Lakshmibala and V. Balakrishnan, Geometric phases for generalized squeezed coherent states,
Phys. Rev. A 55, 869 (1997).
[48] D. A. Trifonov, Generalized uncertainty relations and coherent and squeezed states, J. Opt. Soc. Am. A 17, 2486–
2495 (2000).
[49] F. Hong-Yi and C. Hai-Ling, New approach for calculating Wigner functions of generalized two-mode squeezed
state and squeezed number state via entangled state representation, Commun. Theor. Phys. 36, 651 (2001).
[50] N. Alvarez-Moraga and V. Hussin, Generalized coherent
and squeezed states based on the h(1) ⊕ su(2) algebra,
J. Math. Phys. 43, 2063 (2002).
[51] L. C. Kwek and D. Kiang, Nonlinear squeezed states,
J. Opt. B: Quantum Semiclass. Opt. 5, 383 (2003).
[52] A-S F. Obada and G. M. Abd Al-Kader, A class of nonlinear squeezed coherent states, J. Opt. B: Quantum Semiclass. Opt. 7, S635 (2005).
[53] E. Shchukin, Th. Kiesel and W. Vogel, Generalized minimum -uncertainty squeezed states, Phys. Rev. A 79,
043831 (2009).
[54] M. M. Nieto and D. R. Truax, Holstein-Primakoff/ Bogoliubov Transformations and the Multiboson System,
Fortsch. Phys. 45, 145–156 (1997).
[55] H-C. Fu and R. Sasaki,
Exponential and Laguerre
squeezed states for su(1, 1) algebra and the CalogeroSutherland model, Phys. Rev. A 53, 3836 (1996).
[56] H. Wang, D. Goorskey and M. Xiao, Enhanced Kerr nonlinearity via atomic coherence in a three-level atomic system, Phys. Rev. Lett. 87, 073601 (2001).
[57] A. Gambetta et al., Real-time observation of nonlinear
coherent phonon dynamics in single-walled carbon nanotubes, Nature Phys. 2, 515 (2006).
[58] Y. Yan, J.-P. Zhu and G.-X. Li, Preparation of a nonlinear coherent state of the mechanical resonator in an
optomechanical microcavity, Opt. Exp. 24, 13590–13609
(2016).
[59] S. Dey, A. Fring and V. Hussin, A squeezed review on
coherent states and nonclassicality for non-Hermitian systems with minimal length, In: J. P. Antoine, F. Bagarello,
J. P. Gazeau (eds) Coherent States and Their Applications, Springer Proc. Phys. 205, 209–242, Springer, Cham
(2018).
[60] M. Angelova, A. Hertz and V. Hussin, Squeezed coherent
states and the one-dimensional Morse quantum system,
J. Phys. A: Math. Theor. 45, 244007 (2012).
[61] S. Dey and V. Hussin, Entangled squeezed states in noncommutative spaces with minimal length uncertainty relations, Phys. Rev. D 91, 124017 (2015).
9
[62] V. I. Man’ko, G. Marmo, S. Solimeno and F. Zaccaria,
Physical nonlinear aspects of classical and quantum qoscillators, Int. J. Mod. Phys. A 8, 3577–3597 (1993).
[63] N. Rosen and P. M. Morse, On the vibrations of polyatomic molecules, Phys Rev. 42, 210 (1932).
[64] C. B. Compean and M. Kirchbach, The trigonometric Rosen–Morse potential in the supersymmetric quantum mechanics and its exact solutions, J. Phys. A:
Math. Gen. 39, 547 (2005).
[65] R. J. Glauber, Photon correlations, Phys. Rev. Lett. 10,
84 (1963).
[66] E. C. G. Sudarshan,
Equivalence of semiclassical
and quantum mechanical descriptions of statistical light
beams, Phys. Rev. Lett. 10, 277 (1963).
[67] C. Aragone, G. Guerri, S. Salamo and J. L. Tani, Intelligent spin states, J. Phys. A: Math. Nucl. Gen. 7, L149
(1974).
[68] L. Mandel, Sub-Poissonian photon statistics in resonance
fluorescence, Opt. Lett. 4, 205–207 (1979).
[69] K. Zelaya, O. Rosas-Ortiz, Z. Blanco-Garcia and
S. Cruz y Cruz
Completeness and nonclassicality
of coherent states for generalized oscillator algebras,
Adv. Math. Phys. 2017, 7168592 (2017).
[70] E. Wigner, On the quantum correction for thermodynamic equilibrium, Phys. Rev. 40, 749 (1932).