Journal of Modern Physics, 2011, 2, 472-480
doi:10.4236/jmp.2011.26057 Published Online June 2011 (http://www.SciRP.org/journal/jmp)
Copyright © 2011 SciRes. JMP
Study of the Double Nonlinear Quantum Resonances
in Diatomic Molecules
Gustavo López, Jorge Gomez Tejeda Zanudo
Departamento de Físic a, Uni versidad de Guad al ajara, Guadalajara, Mexico
E-mail: gulopez@udgserv.cencar.udg.m x, jgtz.fis@gmail.com
Received February 23, 2011; revised March 29, 2011; accepted April 19, 2011
Abstract
We study the quantum dynamics of diatomic molecule driven by a circularly polarized resonant electric field.
We look for a quantum effect due to classical chaos appearing due to the overlapping of nonlinear reso-
nances associated to the vibrational and rotational motion. We solve the Schrödinger equation associated
with the wave function expanded in term of proper stationary states, nlm
(vibrational angular
momentum states). Looking for quantum-classic correspondence, we consider the Liouville dynamics in the
two dimensional phase space defined by the coherent-like state of vibrational states. We consider the rela-
tionship between the overlapping of the classical resonances and the mixing of the quantum states, and it is
found some similarities when the quantum dynamics is pictured in terms of number and phase operators.
Keywords: Nonlinear Resonance, Diatomic Molecule, Quantum Resonances
1. Introduction
The study of quantum dynamics in the interval of pa-
rameters where classical chaotic behavior occurs is what
we call “Quantum Chaos, Chaology, or Quantum Mani-
festation of Chaos” [1] which deals with some type of
quantum manifestation of the classical chaos, mainly
associated with the statistical properties of near neighbor
levels of energy of the system [1]. In contrast, for quan-
tum systems associated to non chaotic classical ones, it is
mostly believed that classical dynamical behavior must
occur for large quantum numbers or high value of the
action variable [2] (Rydberg states). In particular, studies
of dynamical chaos in atomic and molecular systems has
been of great theoretical and practical interest [3-12]
since not enough integrals of motion are found either in
the classical or in its quantum system. Different ap-
proaches and studies have been used for the classical
[13-15] and quantum (quasi-classical region) [16-18]
cases, and most of them are based on the Morse potential
as the inter-atomic interaction [19]. On the other hand,
the classical study of the dynamics of atomic and mo-
lecular systems has shown that, under certain conditions,
these systems are capable of exhibiting a chaotic behav-
ior, even in the case the system has few degrees of free-
dom. In what follows of this introduction and to have a
better perspective of the problem, we summarize what
Berman et al. [20] did for the classical part of the prob-
lem.
It is known that the dynamics of a diatomic molecule
in a resonant circularly polarized electric field can be
modeled by the Hamiltonian





00
,; ,; ,=,,; ,
,,,
vib rot
int
HI ppHIHpp
HI t
 


(1)
where describes the vibration of the molecule
along its axis in terms of the action variable I and its
conjugated angle variable
()
0vib
H
, describes the rota-
tion of the molecule around its transversal axis of sym-
metry in terms of spherical coordinates (
()
0rot
H
,,r
), and
=
int
H
tdE describes the interaction of the electric
dipole moment (d) of the molecule with and external
electric field (
tE). These term are given explicitly by

() 2
0
2
() 2
022
,= ,
1
,; ,=,
2sin
vib eee
rot
HI IxI
p
Hppp
r
 
 




(2)
and
G. LÓPEZ ET AL.
473


0
00
2
=sincos
2
cos .
eff
int e
Ee I
Ed t
t



 

(3)
For the derivation of these expressions, the motion of
the molecule has been done with respect the center of
mass

11 221 2
=Rm mmmrr and relative 21
=
rrr
coordinates associated to the diatomic molecule, where
is its reduced mass. The parameter e
x
is defined as

2() 2
0
=2 vib
ee
x
dH dI
. The electric field has been
chosen as , and the magni-
tude of the electric dipole moment has been given by
 
0
=cos,sin,0tEt t

E
0
=2c
eff e
dd eIos
, where 0eff , being
eff the effective charge of the molecule and 0 repre-
sents the point of the minimum on the Morse potential
[19] which simulates the atom-atom interaction in the
diatomic molecule and has been taken up to second order.
The average small vibration oscillation around the equi-
librium point o is just
0
r
r
=de
e
r

2π22
0e
=2 cos
rd I
,
with e
representing the angular frequency of the os-
cillation of the molecule at first order. The dynamical
system described by this Hamiltonian close to resonance
(e
), and under the condition ,


has the
following constant of motion
=constant= ,pI k
(4)
The Hamiltonian (1.1) can be written in a more suit-
able form through a change of variable defined by the
generatrix function
2,,;,,;=Fnkp
 
12n
kp


=
 
 , which are given by
=,=,t
 
 (5)
=12,=,=,=,
I
npppkI

t
 (6)
and the Hamiltonian in this coordinates is written as


22
2
2
11 111
,;,;, sin cos
22 222
sin
eee W
HNPKnxnpk nn
 

 
  

 
 


 , (7)
where the variable “n” is assumed to have continuous
values, and the following definitions have been made:
2
0
=2r2
, and 0
=2
eff e
WEe
. This Hamilto-
nian depends only in the conjugate variables (, )n
and
(,)p
since
is an ignorable variable, and therefore,
k is a constant of motion. In this way, the Hamilton equ-
ations with this variables define the four dimensional
classical dynamical system

2
2
3
d1
=sinsin,
d22
d121
=2 sinc
d22
41/2
sin
d2 11
=coscoscos,
d222
sin
d=2 .
d
eee
nW
n
W
xnknn
pW
kn n
p


os,
 




 
 

 

 



(8)
This system has its critical points at ,
=0p=π2
,
=πm
with , and with the
roots of a third order polynomial. In the example given
by the reference [20], the parameters associated to the
diatomic molecule GeO [17] are used,
mZ=i
nn =1i,2,3
1
=985.8cm,()= 15cm,
ee

1
1
1
=2.2 cm,=0.48 cm,
ee
x

(9)
0=3.28dD, ,
0=1.62rÅ= 13.1 amu
where the units have been selected such that
1
3
==1cm
2b
kT
and . Note that these =0.956TK
parameters correspond to be close to resonance. Then,
there is one center points at =π
and which
forms the first resonances of the system (for W = 0.05
cm–1 and ), and there is other resonance located
near 2
12n
=0k
1.5n
which is due to
degree of freedom
rather that a critical point of the system. The resonances
overlapping Chirikov’s criterion [13] for appearing of
chaos was verified at , and total chaotic
behavior is observed after , see Figure 1.
This result suggests that classical chaotic behavior ap-
pears within the first two exited states of the associated
quantum system. However, the correspondence principle
[2] tells us that the quasi-classical behavior of a quantum
system is gotten for very large quantum numbers.
Therefore, one would not expect any quasi-classical be-
havior for ground and first exited states of the quantum
system.
1
= 0.177W
=1 cmW cm
.03 1
Copyright © 2011 SciRes. JMP
474 G. LÓPEZ ET AL.
(a)
(b)
(c)
Figure 1. Poincaré map for θ = π/2 with θ > 0, k = 0, θ0 = 1,
and p0 = 0 and for: (a) W = 0.048 cm–1; (b) W = 0.68 cm–1; (c)
W = 1.03 cm–1.
In this way, in this paper we study the quantum be-
havior of this system in the region of parameters where
this classical chaotic dynamics appears. This behavior
could be important in the study of diatomic molecular
clouds during the star born formation or supernova wind
shock studies from dying stars [25,26]. We solve the
associated Schrödinger equation, assuming the wave
function is a linear combination of the stationary states
with time depending coefficients and solving numeri-
cally the resulting equations for these coefficients, and
picture the expectation values of the quantum variables
in a phase space-like to look for a similarity with the
classical behavior.
2. Quantum Dynamics
2.1. Quantum Hamiltonian
Our goal is to solve the Schrödinger equation,
 
ˆ
=itHt
t

,
(10)
where ˆ
H
is the Hermitian operator associated to the
Hamiltonian,

2
22
22
0
0
1
=2sin
2sin cos.
2
eee
eff
e
p
HIxI p
r
Ee It
 



 



(11)
For this propose, the following operators are assigned
to the observables,
2
22 2
2
1ˆ
,,
2
ˆ
ˆˆ
=,,
sin
Iaa
p
Lp L

 

 







(12)
where a and represent the usual ascend and
descend operators of the quantum harmonic oscillator. If

=aa
n and
lm are the basis for the harmonic oscil-
lator and the angular moment operators such that
††
=,= 1
ˆ
==,,=1
an nnan nn
aannnnn aa


,
,
(13)
and

22
ˆ=1,
ˆ=,
z
Llmll lm
Llmmlmlml
 
(14)
the action of ˆ
is defined in terms of the phase opera-
tor [21-23] as
Copyright © 2011 SciRes. JMP
G. LÓPEZ ET AL.
Copyright © 2011 SciRes. JMP
475
 
ˆˆˆ
ˆˆ ˆˆ
1
e:=,e:=e =,
11
ˆˆ
cos:=e e,sin:=e e
22
iii
ii ii
aa
aa aa

 




1
.
(15)
From Equation (11) and the definitions (12) - (14), we
get the quantum Hamiltonian of the form 0
ˆˆ ˆ
=
H
HW
,
where 0
ˆ
H
and are given by
ˆ
W
2
††
0
2
2
11
ˆ=22
ˆ,
eee
Haaxaa
L


 



These operators have the following properties, (17)
ˆˆ
ˆˆ ˆ
ˆˆ
,cos=sin,,sin=cos,
e=1,e =1.
ii
nin i
nn nn

ˆ

 
 

(16)
and
  
1 1
ˆˆ ˆˆ ˆ
ˆˆˆ ˆˆ
ˆ ˆ
=sincos cossin sincos cossin sin.
22 2
W
Wnt tt tn
  
 
 

  
 

 
 
(18)
To construct the Hermitian operator , we have
used the fact that for any operator
ˆ
W
ˆ
A
, the operator
ˆˆ 2AA is Hermitian. Using the commutation relations
of Equation (16), one gets the commutativity of the vibra-
tional and rotational operators,
,
and we see that

,,
g

ˆˆ
, =fa

a0
ˆˆˆ
=
z
kL n
1
2
is a constant of
motion, that is

0
2
ˆ=,
11
=1
22
nl
nlee e
H nlmEnlm
Enxn ll

 ,

 
 

(20)
where there is a (2 1)l
degeneration due to quantum
number “m”. On the other hand, since the relation
1, >0
nl nl
EE
must be satisfied, the quantum vibra-
tional number “n” is bounded [19] in the following way
1
ˆ
ˆˆˆ
ˆ
,=, =
2
z
Hk HLn

 

 


0
(19) 11
0
22
e
nx,
 
(21)
which is the quantum analogue of Equation (4) and has
the physical meaning that external electric field’s pho-
tons excite the rotational and vibrational degree of free-
dom with the same number of quanta. The main reason
for choosing the phase operator as Equation (15) was to
be able to get this quantum constant of motion correctly.
where []
x
means the integer part of the real number “x”.
Therefore, the spectrum is finite. Let us propose the so-
lution of Equation (10) of the form
 
max
= . (22)
itEnl
nlm
nlm
tDtenlm
Now, substituting this equation in (10) and using the
orthogonality relation =nnll mm
nlm nlm

 
 , we get
the system of equations for the coefficients as
2.2. Time Evolution Equations
Using the number states n and the spherical harmonic
states lm , we see that 0
ˆ
H
is diagonal in the basis


max
;
ˆ
=,
it EE
nl nl
nlmnlmnlm nlm
nlm
iDeD W

  
(23)
=nlm n lm, and the eigenvalues are given by
where the matrix elements ,ˆ
=
nlm nlm
WnlmWn
  lm are given by


,
11
ˆˆ ˆ
ˆ
ˆˆ
=coscossin cos
22 2
11
ˆˆ ˆ
ˆ
ˆˆ
sinsinsin sin.
22
nlm nlmW
Wnnnnnnlm
nnn nnnlmtlm
 
 
 


 
 



 


tlm
From the expressions A1 to A7 of the Appendix, this matrix elements can be written as


 
11 1111
,=1212
421
llitnmit nm
lm lm
nlm nlmllnmnm
lmlm
cc
W
Wnnee
clc l

 
 

 


 




.
21
(24)
Substituting this expression in Equation (23) and using
the same dimensionless variables defined in the intro-
duction, we get the time evolution equation of the coeffi-
cients as
476 G. LÓPEZ ET AL.
 
 
,( ),( ),( ),()
1, 1,11, 1,1
1, 11, 1
,( ),(),
1, 11, 1
1, 1,1
ee
=1212 1212
4(21) (23)
ee
12 3212 32
(2 1)
ii
nl nl
lm lm
nlmn lmn lm
lm lm
i i
nl nl
lm lm
nlm
lm
cc
W
iDnnDn nD
cl cl
cc
nnD nn
cl

 

 
   
 
 

 
 
 

 
(),()
1, 1,1,
(2 1)nlm
lm
D
cl

 
where we have made the definitions =d dDD
,
,( ),( )1,1
=
nln lnl
EE
 

=EE, and
,( ),( )1,1nln lnl
 . The selection rules de-
duced from (26) are

=1, =1and=1.ln m  (25)
Note that the last two terms of these expressions are a
consequence of the constant of motion (19). The time
evolution of the coefficients in Equation (27) and the
selection rules in Equation (25) are similar to the electric
dipole transitions in an atom, except with the extra selec-
tion rule of n. Furthermore, suppose we are initially in a
given state

00 0
0=nlm and we set the frequency
to be such that it is almost in resonance with the fre-
quency of an allowed transition, say
f
ff
nlm (that is
,
00
nl nl
ff
EE
,
). For this case and neglecting the
non-resonant transitions, the equations of motion (28)
becomes
0
=and=
ii
r
ff
iDe DiDeD



0
,
r

(26)
where
and are defined as
r


0,,
00
=121242
flmlmf
ff
nnWccl
 1
and ,,
=
rnl nl
ff oo
EE
. In matrix notation, Equa-
tion (26) is written as
0
0
d=
d0
i
i
0
f
f
D
e
iD
e

 

 


 
D
D
which in terms of the Pauli operators becomes
0
dˆˆ
=cossin
d
with =,
rx ry
f
i
D
D
,
 




and this one is of the form
dˆ
=,
d
ˆˆˆ
with= cossin,
at
atrxry
iH
H

 

3
(27)
which is the Schrödinger equation for a two level atom
introduced in a circularly polarized electromagnetic field
[5].
2.3. Numerical Results
We solve numerically Equation (28), considering only
the coefficients nlm for . We use the same
parameters used in the classical numerical calculations,
Equation (9), which implies to have a close resonant
transition between the states
D,nl
100 and 211 with
10,( ),( )=0.82

. Higher order of excitation are not
considered since we want to see what happen to the
states closely related with the classical ones, where clas-
sical chaos appears. In this way, we are not interested
here in the quasi-classical region (very highly exited
states) but in the deep quantum region (first few states),
where quantum-classical correspondence is not expected.
The results of the numerical simulations are shown in
Figures 2(a) and (b) which represent the typical oscilla-
tion between the population in the two resonant states.
The Figure 2(a) shows the transition probabilities for
small values of W before there is a considerable mixing
of states (observe that 2
211 <0.5c). Figure 2(b) shows
the same probabilities but for the value
1
=1.03 cmW
which should correspond to have clas-
sical chaotic behavior, in the classical dynamics (note
that the value of W for classical chaos to appear is
1
=0.177 cm
ch
W
).
We see also that the classical value of the closed clas-
sical resonance suggests overlapping between quantum
states in n = 1 and n = 2, as we precisely observe in our
simulations, which is consequence of the resonant transi-
tion frequency between the states 100 and 211.
2.4. Quantum Phase Space Pictures
In this section we try two different approaches to see a
better relation between the quantum and classical dy-
namics. Here, we are interested in the behavior of ex-
pectation values of the dynamical variables rather than
the statistical properties in the phase space due to the
Schrödinger wave function (Wigner [27], Husimi [28], or
Glauber-Sudarshan [29,30] distribution functions) since
these values are the ones we want to compare. The first
and most used approach [21-23], is to use the phase
space representation in terms of the expectation value of
the dimensionless canonical variables ˆ
X
and
ˆ
P
††
ˆˆ
=,=
22
aa aa
XP
i

.
(28)
Copyright © 2011 SciRes. JMP
G. LÓPEZ ET AL.
477
(a)
(a)
Figure 2. Time evolution of the total probability, the probability amplitude of the state 100 (upper line) and the state 211
(lower line) for different values of the perturbation: (a) W = 0.048 cm–1, solid; W = 0.19 cm–1, dashed; W = 0.68 cm–1, bold; (b)
W = 1.03 cm–1.
The results of the numerical simulations of this ap-
proaches is presented in Figure 3, where we used the
same parameters of expression (15) and W = 1.03 cm–1.
For the initial state wave function of the system we chose
a poisson-like distribution in the coefficients with
the maximum in value in the state . This initial state
is determined by the following coefficients
0nl
D
100
D
000 010020
100 110120
200 210220
1.5 0.2 0.05
=,=,=
12 1212
80.40.
=,=,=
12 1212
1.5 0.2 0.05
=,=,=
12 1212
DDD
DD D
DDD
1
.
(29)
Copyright © 2011 SciRes. JMP
478 G. LÓPEZ ET AL.
Figure 3. Phase space like picture with the expectation values variable
X
and
P
for the initially poission-like distribu-
tion, Equation (29), using the same parameters as with the classical case.
The reason to use this initial state is based in the prop-
erties of the coherent states of the harmonic oscillator.
This selection not only gives a well defined initial value
of the expectation values, but also permits a further study
in terms of the Liouville dynamics for both the classical
and quantum case [24]. Although the dynamics of each
variable seem to be stable and similar to each other, the
phase space representation does not seem to give any
picture alike to the classical dynamics of the system (see
reference [20]), i.e., the phase space in terms of the ca-
nonical variables ˆ
X
and ˆ
P shows no resem-
blance to the classical dynamics, and this happen for
other different initial states.
In the second approach, we will use the phase space
representation in terms of the expectation value of

ˆ
arg ei
, as the angle, and the expectation value of the
number operator . In polar coordinates, the
n
ˆˆ
n
will
correspond to the radius and

ˆ
arg ei
to the angle of
the phase space of this set of variables, and
na
a
ˆ=
ˆ
e=
ia
a
3. Conclusions and Comments
We have studied the quantization of a diatomic molecule
by solving the Schrödinger equation with the known
Hamiltonian of the diatomic molecule with a circularly
polarized resonant rf-field, written in spherical coordi-
nates (rotations) and angle-action variables (vibrations).
The wave function was expanded in a finite combination
of a proper stationary basis with time dependent coeffi-
cients, and the system of equations for these coefficients
was obtained. Using the same parameters as in the clas-
sical case, a near resonant transition between the states
100 and 211 is gotten, which correspond to the
closer integer numbers for where the classical non
linear resonances appeared, 1 and 2
nn21.5n
. Us-
ing a poisson-like distributed initial wave function in the
quantum numbers, with maximum in the resonant state
100 , we try two different approaches to see the quan-
tum phase space expectation value dynamics and com-
pare it with the classical case. The usual approach, using
the canonical variables ˆ
X
and , fails to provide any
intuitive picture of the classical case. On the other hand,
the approach using the expectation value of
ˆ
P
ˆ
i
e
and
suggests some resemblance and relationship with the
classical case. Therefore, we have here the following
situation, on one hand, the correspondence principle tells
us that we must have the quasi-classical behavior (clas-
sical chaos) for this quantum system at very large quan-
tum number. However, classical chaotic behavior is ob-
tained just for the associated first states of the quantum
system, implying that quasi-classical chaotic behavior
can not be obtained here. So, as one could expect for this
case, quantum dynamics does not follow the classical
one.
ˆ
n
1a
. The expectation values of these va-
riables represent the classical analogous of the variables
on Figure 1 above. In the upper left plot of Figure 4 it is
shown the time evolution of n
which resemblances to
the classical case in terms of the main and different fre-
quencies with which it oscillates. For the numerical si-
mulations results presented in these figures we used the
same parameters as in the Subsection 2.3, and the same
initial state (29). The phase space picture in terms of the
operators and ˆ
ˆ
ni
e
seems to have a little bit resem-
blance with the classical results, perhaps because the
dynamics of resembles the classical part. Also, the
sudden slow changes of
ˆ
n
ˆ
i
e
arg
seem to suggest some
kind of relation with the resonances of the classical
case.
Copyright © 2011 SciRes. JMP
G. LÓPEZ ET AL.
Copyright © 2011 SciRes. JMP
479
Figure 4. Phase space like picture with the same parameters as in the classical case in terms of the expectation values ˆ
n
and
ˆ
i
arg e
.
4. Acknowledgements
We want to thank Professor Gennady P. Berman for his
comments and suggestions about this subject, and
CONACYT for its support with the grand number
0104129.
5. References
[1] E. R. Linda, “The Transition to Chaos,” Springer-Verlag,
New York, 2004.
[2] A. Messiah, “Quantum Mechanics I,” John Wiley & Sons,
New York, 1964.
[3] A. J. Lichtenberg and M. A. Liberman, “Regular and
Stochastic Motion,” Springer-Verlag, Berlin, 1983.
[4] G. Casati, B. V. Chirikov, D. L. Shepelyansky and I.
Guarnery, “Relevance of Classical Chaos in Quantum
Mechanics: The Hydrogen Atom in Monochromatic
Field,” Physics Reports, Vol. 154, No. 2, 1983, pp.
77-123. doi:10.1016/0370-1573(87)90009-3
[5] P. Lobastie, M. C. Bordas, B. Tribollet and M. Boyer,
“Stroboscopic Effect between Electronic and Nuclear
Motion in Highly Excited Molecular Rydberg Sates,”
Physicsl Review Letters, Vol. 52, No. 19, 1984, pp. 1681-
1684. doi:10.1103/PhysRevLett.52.1681
[6] J. Chevaleyre, C. Bordas, M. Boyer and P. Labastie,
“Stark Multiplets in Molecular Rydberg States,” Physicsl
Review Letters, Vol. 57, No. 24, 1986, pp. 3027-3030.
doi:10.1103/PhysRevLett.57.3027
[7] C. Bordas, P. F. Brevet, M. Boyer, J. Chevaleyre, P. La-
bastie and J. P. Perrot, “Electric-Field-Hinhered Vibra-
tional Autoionization in Molecular Rydberg States,”
Physicsl Review Letters, Vol. 60, 1988, p. 917.
doi:10.1103/PhysRevLett.60.917
[8] M. Lombardi, P. Labastie, M. C. Bordas and M. Boyer,
“Molecular Rydberg States: Clasical Chaos and Its Cor-
respondence in Quantum Mechanics,” Journal of Chemi-
cal Physics, Vol. 89, No. 6, 1988, 3479-3490.
doi:10.1063/1.454918
[9] M. Lombardi and T. H. Seligman, “Universal and Non-
universal Statistical Properties of Levels and Intensities
for Chaotic Rydberg Molecules,” Physical Review A, Vol.
47, No. 5, 1993, pp. 3571-3586.
doi:10.1103/PhysRevA.47.3571
[10] J. J. Kay, S. L. Coy, V. S. Petrovic, B. M. Wong and R.
W. Field, “Separation of Long-Range and Short-Range
Interactions in Rydberg States of Diatomic Molecules,”
Journal of Chemical Physics, Vol. 128, No. 19, 2008, p.
194301. doi:10.1063/1.2907858
[11] D. Sugny, L. Bomble, T. Ribeyre, O. Dulieu and M.
Desouter-Lecomte, “Rotovibrational Controlled-Not
Gates Using Optimized Stimulated Raman Adiabatic
Passage Techniques and Optimal Control Theory,”
Physical Review A, Vol. 80, 2009, ID 042325.
doi:10.1103/PhysRevA.80.042325
[12] A. Ruiz, J. P. Palao and E. J. Heller, “Nearly Resonant
Multidimensional Systems under a Transient Perturbative
Interaction,” Physical Review E, Vol. 80, 2009, ID
066606. doi:10.1103/PhysRevE.80.066606
[13] B. V. Chirikov, “A Universal Instability of Many-Dimen-
sional Oscillator Systems,” Physics Reports, Vol. 52, No.
5, 1979, pp. 263-379.
doi:10.1016/0370-1573(79)90023-1
[14] É. V. Shuryak, “Nonlinear Resonances in Quantum Sys-
tems,” Soviet Physics-JETP, Vol. 44, 1976, p. 1070.
[15] R. P. Parson, “Vibrational Adiabaticity and Infrared Mul-
tiphoton Dynamics,” Journal of Chemical Physics, Vol.
88, No. 6, 1987, pp. 3655-3666.
doi:10.1063/1.453865
[16] P. S. Dardi and K. Gray “Classical and Quantum Me-
chanics Studies of HF in an Intense Laser Field,” Journal
480 G. LÓPEZ ET AL.
of Chemical Physics, Vol. 77, No. 3, 1982, 1345-1353.
doi:10.1063/1.443957
[17] G. P. Berman and A. R. Kolovsky, “Quantum Chaos in a
Diatomic Molecule Interacting with a Resonant Field,”
Soviet Physics-JETP, Vol. 68, 1989, p. 898.
[18] G. P. Berman and A. R. Kolovsky, “Quantum Chaos in
Interactions of Multilevel Quantum Systems with a Co-
herent Radiation Field,” Soviet Physics-USP, Vol. 35, No.
4, 1992, p. 303.
[19] P. M. Morse, “Diatomic Molecules According to the
Wave Mechanics II. Vibrational Lavels,” Physical Re-
view, Vol. 34, No. 1, 1929, pp. 57-64.
[20] G. P. Berman, E. N. Bulgakov and D. D. Holm, “Nonlin-
ear Resonance and Dynamical Chaos in a Diatomic
Molecule Driven by a Resonant RF Field,” Physical Re-
view A, Vol. 52, 1995, p. 3074.
doi:10.1103/PhysRevA.52.3074
[21] L. Susskind and J. Glogower, “Quantum Mechanical
Phase and Time Operator,” Physics, Vol. 1, 1964, pp.
49-61.
[22] A. Lahiri, G. Ghosh and T. K. Kar, “Action-Angle Vari-
ables in Quantum Mechanics,” Physical Letters A, Vol.
238, No. 4-5, 1998, 239-243.
doi:10.1016/S0375-9601(97)00926-2
[23] P. Carruthers and M. M. Nieto, “Phase and Angle Vari-
ables in Quantum Mechanics,” Reviews of Modern Phys-
ics, Vol. 40, No. 2, 1968, pp. 411-440.
doi:10.1103/RevModPhys.40.411
[24] G. J. Milburn, “Quantum and Classical Liouville Dy-
namics of the Anharmonic Oscillator,” Physical Reviews
A, Vol. 33, No. 1, 1980, pp. 674-685.
doi:10.1103/PhysRevA.33.674
[25] M. Burton, “IC-443-The Interaction of a Supernova Rem-
nant with a Molecular Cloud,” Royal Astronomical Soci-
ety, Vol. 28, 1987, pp. 269-276.
[26] R. Chevalier, “Supernova Remnants in Molecular-
Clouds,” Astrophysical Journal 511, 798 (1999).
doi:10.1086/306710
[27] E. Wigner, “On the Quantum Correction for Thermody-
namic Equilibrium,” Physical Review, Vol. 40, No. 5,
1932, pp. 749-759.
doi:10.1103/PhysRev.40.749
[28] K. Husimi, “Some Formal Properties of the Density Ma-
trix,” Journal of the Physical Society of Japan, Vol. 22,
No. 4, 1940, pp. 264-314.
[29] R. J. Glauber, “Coherent and Incoherent States of the
Radiation Field,” Physical Review, Vol. 131, No. 6, 1963,
pp. 2766-2788.
doi:10.1103/PhysRev.131.2766
[30] E. C. G. Sudarshan, “Equivalence of Semiclassical and
Quantum Mechanical Descriptions of Statistical Light
Beams,” Physical Review Letters, Vol. 10, No. 7, 1963,
277-279.
Copyright © 2011 SciRes. JMP