You are on page 1of 6

Solid State Communications 146 (2008) 169174

www.elsevier.com/locate/ssc
Charge degeneracy for a ThomasFermi hydrogen molecule:
Boundunbound transition
Jo ao B. Diniz
a,c
, Andr e L.A. Penna
a,c,
, S ergio L. Garavelli
b,c
, Fernando A. Oliveira
a,c
a
Instituto de Fsica - Universidade de Brasilia, CP 04455, 70919-970 Brasilia DF, Brazil
b
Departamento de Fsica - Universidade Cat olica de Brasilia, CP 04455, 70919-970 Brasilia DF, Brazil
c
International Center for Condensed Matter Physics, CP 04455, 70919-970 Brasilia DF, Brazil
Received 3 December 2007; received in revised form 21 January 2008; accepted 23 January 2008 by F. De la Cruz
Available online 6 February 2008
Abstract
We study the relationship between quantum screening effects and charge symmetry in the hydrogen molecule. In this model, the charge
symmetry of the molecule is broken in the presence of the ThomasFermi potential exp(qr)/r. We argue that screening acts as an external eld
that promotes charge degeneracy in the system. In the presence of screening (q = 0), the charge symmetry of the H
+
2
molecule is broken and
thus the effective nuclear charge splits into two, (, ), where is the dual charge. The effective nuclear charge is then related to the
binding energy of the electron, while the dual charge is responsible for the screening effect in the wavefunction. Naturally, both the energy and
the wavefunction are modied by this charge degeneracy. We obtain an analytical formulation for the molecular energy with a charge degeneracy
effect. To test the efciency of our model, we analyze the behavior of the molecules energy as a function of the ThomasFermi parameter q and
calculate the critical parameter q

, as well as the critical dual charge

for which the boundunbound transition occurs. We observe an interesting


super-freezing screening phenomenon for the ThomasFermi hydrogen molecule.
c 2008 Elsevier Ltd. All rights reserved.
PACS: 71.30.+h; 71.15.-m
Keywords: D. Electronic states; D. Charge degeneracy; D. Boundunbound transition; D. Metalinsulator transition
1. Introduction
The screening effect in molecules and solids is one of
the most important and complex phenomena that originates
from electronelectron interaction. In condensed matter, the
screening phenomenon occurs when atoms and molecules in
a crystal lattice are immersed in a gas of electrons, which
alters the electric permittivity of the medium. In such cases,
the electron gas can modify the form of the intra-atomic (or
inter-atomic) potential energy between atoms (or between the
nuclear charge and the electron of an atom), and thus can locally
change the binding energy of these systems [1]. Derivation of
a detailed model that can deal with screening phenomena for

Corresponding author at: Instituto de Fsica - Universidade de Brasilia, CP


04455, 70919-970 Brasilia DF, Brazil. Tel.: +55 (61) 3307 2571; fax: +55 (61)
3307 2572.
E-mail address: penna.andre@gmail.com (A.L.A. Penna).
a realistic system is a very complicated task. In fact, an exact
theory of such screening requires the computation of all the
orders of interactions of the electron gas system. Considering
this difculty, a simpler approach for the study of screening
effects is to consider a ThomasFermi potential to model such
interactions [216]. For such systems, a typical starting point
involves describing the interaction between the elementary
charges and the potential
V(r) =
e
2
r
exp(qr), (1)
which represents the interaction between an electron and a
proton of absolute charge e shielded by a gas of electrons with
density N. We then denote the ThomasFermi parameter by
q = (6e
2
N/k
B
T
f
)
1/2
, where k
B
and T
f
are the Boltzmann
constant and the Fermi temperature, respectively. This potential
is used in condensed matter physics as the ThomasFermi
0038-1098/$ - see front matter c 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ssc.2008.01.035
170 J.B. Diniz et al. / Solid State Communications 146 (2008) 169174
potential [1], in plasma theory as the DebyeH uckel potential,
and in nuclear physics as the Yukawa potential [2].
Recently, screening concepts have found important appli-
cations in the study of macromolecules [17], the behavior of
metalinsulator transitions [1825], the properties of hydro-
gen under pressure [26,27], molecules adsorbed on metallic
surfaces [2834], and for describing many properties of alkali
atoms. Screening is relevant for the understanding of charged
polymer gels [3537], in the synthesis of charged nanostruc-
tures and nanostructure devices [38,39], and in several interest-
ing applications of modern solid-state physics in general [40
45]. Boundunbound transitions have been associated with
metalinsulator transitions since Motts seminal work [18,19].
It is not surprising that many attempts to nd accurate solutions
for this potential have been reported [216].
There are many ways to account for screening effects. As
a starting point we use the method described by Garavelli
and Oliveira [11]. For the hydrogen atom, this procedure
naturally yields a screened wavefunction (SWF), and results in
an improvement in the energy of the system. In this method,
we can start with an iterative Schr odinger equation in the
momentum space

k+1
( p) =
2
p
2
2E
k+1
_

V( p

p)

k
( p

) d
3
p

, (2)
where

k
( p) and

V( p) are the Fourier transforms of the
wavefunction and potential equation (1), respectively; k =
1, 2, . . . is the order of the iteration. Note that the potential

V( p) is xed, whereas the wavefunction


k+1
( p) is iterated
for each k. For the screened hydrogen atom, it is reasonable
to suppose that the energy E
k
is associated with an effective
nuclear charge
k
viewed by the electron, which assumes the
form

2
k
= 2E
k
. (3)
We can start with the wavefunction of the ground state in
the momentum space,

1
( p) = a
1
/( p
2
+
2
1
)
2
, which is the
solution for the unscreened hydrogen atom (q = 0), where a
1
is a constant. After a few iterations, we then obtain an analytical
expression for Eq. (2) as

k+1
( p) =
a
k
( p
2
+
2
k
)( p
2
+
2
k
)
, (4)
where a
k
is a constant. We introduce a newparameter,
k
, which
is related to the nuclear charge of the system,
k
, according to

2
k+1
= (
k
+ q)(
k
+ q). (5)
Here we note a quantitative gain: we start out from an iterative
process with complicated integrals and end up with the simple
algebraic recurrence relationships in Eqs. (4) and (5). This
means that after many iterations the k index can be dropped.
The inverse Fourier transform of Eq. (4) yields the SWF of the
ground state, namely
(r) =
e
r
e
r
( )r
= e
r
g(r), (6)
where = is the screening parameter of the screening
factor g(x) = (1 e
x
)/x. Note that screening acts as
an external eld that breaks down the charge symmetry and
creates the order parameter . Note that the parameter
becomes a new charge in theory and provides screening in
the wavefunction. In the absence of screening (q = 0), we can
conjecture that the charges and are symmetrical, which
might correspond to charge degeneracy in the system. We
believe that charge symmetry is important for understanding the
quantum screening effects in atomic and molecular structures.
In the present study, we investigate the relationship between
quantum screening effects and charge degeneracy in the
screened H
+
2
molecule. To understand the screening effect,
we argue that each hydrogen atom of the molecule H
+
2
has
a hidden charge degeneracy due to the Coulomb potential,
which establishes a charge symmetry in the molecular structure.
This motivated us to search for charge degeneracy (or charge
symmetry breaking) for the ThomasFermi molecule. This
means that in the presence of screening (q = 0), the effective
nuclear charge is degenerated (, ), giving rise to a
dual charge
1
responsible for screening in the wavefunction
of the system. Our main goal is to investigate novel aspects
due to the charge degeneracy of this molecular system. In this
communication, we show how charge degeneracy affects the
molecular energy and the effective nuclear charge of the system.
The boundunbound transition is analyzed, and we obtain its
critical parameter q

, as well as its critical dual charge

.
Finally, we discuss a super-freezing screening phenomenon
observed for a region where (q) 0 for q = 0 in the
ThomasFermi hydrogen molecule.
2. Charge degeneracy
Symmetry plays a fundamental role in describing important
natural phenomena in physics, which prompted us to explore
how the charge symmetry of the unscreened hydrogen molecule
is related to the quantum screening phenomenon. We start
with the hydrogen atom in the absence of screening (q = 0),
whose charge interaction is then mediated by a Coulomb-
type potential. In this case, charge symmetry dictates that the
electron sees only a well-dened and symmetrical effective
nuclear charge for the whole system. In a Coulomb potential,
the hydrogen atom exhibits a hidden charge degeneracy that
establishes a charge symmetry for the electron in any quantum
state. On the other hand, in the presence of a ThomasFermi
potential, there is a nuclear charge degeneracy manifested that
gives rise to charge symmetry breaking in the atomic system.
This suggests that the ThomasFermi parameter q acts as an
external eld in breaking down the charge symmetry of the
atom to create a quantum screening phenomenon. The charge
degeneracy then promotes the splitting of the nuclear charge
into two charges: an effective nuclear charge and a dual
charge , which modify the energy and wavefunction of the
1
The dual term refers to degeneracy (, ), which is a pair (or dual).
We then denote as the dual charge.
J.B. Diniz et al. / Solid State Communications 146 (2008) 169174 171
electron. Hence, charge symmetry is recovered when =
as q = 0, i.e., both the charges and are symmetrically
indistinguishable, or not degenerate. In this case, the electron
sees only the nuclear charge of the molecule. In fact,
it elucidates the hidden charge degeneracy in the hydrogen
molecule. This symmetry can be seen for the hydrogen atom
by taking the limit into Eq. (6), where we recover the
unscreened wavefunction (UWF) for the ground state. On the
other hand, for q = 0, the model exhibits charge degeneracy,
which means that = , and so the electron sees two
effective charges and with different numerical values. In
this sense, the effective nuclear charge is directly related to
the binding energy of the molecule, whereas the dual charge
is responsible for the screening in the wavefunction by
changing the binding energy of the system.
In this model, we assume that when q increases the nuclear
charge degenerates and and depart from each other. The
charge degeneracy allows us to dene the order parameter,
= , which can be called the screening parameter
of the wavefunction. Note that this parameter depends on the
dual charge , as well as on the effective nuclear charge .
Thus, the screening parameter depends on the existence of
charge degeneracy in the system. If = 0, charge symmetry is
recovered ( = ) in the model (no screening).
We are now ready to study the ThomasFermi H
+
2
molecule with a charge degeneracy effect. We xed the
protons at the positions (0, 0, R/2) and (0, 0, +R/2), while r
denotes the electron position. In this framework, the molecular
wavefunction is given by the combination, (r) = (|r +
R
2
|) + (|r
R
2
|), where is the screened atomic function,
dened by Eq. (6), and R is the distance between the ions. In
this case, V(|r +
R
2
|) and V(|r
R
2
|) are the electronproton
interactions, while V(|R|) represents the protonproton
interaction, according to the denition of Eq. (1). Since we
consider the protons xed, i.e. no vibrations, we have no extra
contributions for the dielectric constant. As a consequence, the
protonproton interaction is only modulated by the electronic
screening, in the same way as the electronproton interaction.
We begin by writing the Hamiltonian for a screened
molecular system by considering the ThomasFermi potential
equation (1)

H =

2
2m
+ V
_

r +
R
2

_
+ V
_

r
R
2

_
V(|R|). (7)
We note that both Hamiltonian and wavefunction are invariant
to the transformation r r. We then start out with the
Hamiltonian operator

H to calculate the functional energy E
of the molecular system
E =
(r)|

H|(r)
(r)|(r)
=
(|r+
R
2
|)|

H|(|r+
R
2
|)+(|r+
R
2
|)|

H|(|r
R
2
|)
(|r+
R
2
|)|(|r+
R
2
|)+(|r+
R
2
|)|(|r
R
2
|)
. (8)
Here we shall use atomic units. From Eq. (8) we obtain the
energy
E = K +U +
e
q R
R
, (9)
where the kinetic energy and the electronic contribution to
the potential energy, which can be analytically computed, are
respectively given by
K =
1
N
_
e
R
2
_
4
R
+
2
R
_

e
R
2
_
4
R
+ +
2
R
_
+

2
2
_
,
U =

N
_
2A(q)
[A(q) A(q)]e
q R
q R
+ B(q) + B(q) + C(, ) + C(, )
_
,
where = + , N is the normalization constant
N =

2

+ e
R
_
4
R
+

_
+ e
R
_

4
R
+

_
,
and
A(q) = ln
_
(2 + q)(2 + q)
( + q)
2
_
,
B(q) =
e
q R
q R
{Ei ((2 + q)R)
+ Ei ((2 + q)R) 2Ei (( + q)R)},
C(, ) =
2e
R
R
_
ln
_
(2 + q)( + q)
(2 + + q)q
_
+ Ei (( + q)R) Ei (q R)
_

2e
R
R
{Ei ((2 + q)R) Ei (( + q)R)}.
Here, Ei (x) is the exponential integral function. Note that
calculation of the energy, Eq. (9), results in a lengthy but
analytical equation. Using iterative and variational procedures,
we now analyze the behavior of the energy (9) and of
parameters (effective nuclear charge), (dual charge), and
(screening parameter).
3. Numerical results
We dene an iterative process using Eqs. (3) and (9). This
method allows us to obtain the energy and the parameters
and as functions of q. Furthermore, we compare the
results obtained from the iterative method with those from the
variational procedure, which consists of minimizing Eq. (9) in
relation to the parameters and .
In Fig. 1, we present the energy E(R) as a function of the
inter-atomic distance R for (a)(c) q = 0. Curve (a) shows
the iterative process, which has a minimum at R
eq
= 2.330
and binding energy of E
a
= 0.5761. Curve (b) shows the
results of the variational process for the UWF ( = 0), which
has a minimum at R
eq
= 2.033 and binding energy of E
b
=
0.5865, in good agreement with Coulson value E
coulson
=
0.5648 [46]. Curve (c) shows the variational process for SWF
172 J.B. Diniz et al. / Solid State Communications 146 (2008) 169174
Fig. 1. Energy as a function of inter-protonic distance R (atomic units). Curves
(a)(c) for q = 0. Curves (d)(f) for q = 0.5. Curves (a), (e): iterative process;
curves (b), (d): variational process for the UWF; curves (c), (f): variational
process.
for q = 0. It should be noted that the curves (c) and (b)
are identical (same values for R
eq
and E
c
). As expected, the
variational result is better than the iterative result. However,
we should bear in mind that the energy used in the variational
process (Eq. (9)) was obtained using the iterative process. The
minimization process yields better values of and than the
iterative process, which may be improved to provide a new
wavefunction and energy. Despite its limitation, the iterative
process is better than the variational procedure for the UWF as
q increases. For curves (d)(f), q = 0.5. Again the variational
energy in curve (f) is lower than the UWF, represented by curve
(e) and the iterative result shown in curve (d).
In Fig. 2, we plot the binding energy E
b
as a function of
the ThomasFermi parameter q. Curve (a) is for the variational
process, curve (b), for the iterative process while curve (c) is
for the variational process for the unscreened function ( = 0).
Note that for very small q the variational processes with the
SWF and UWF yield better results than the iterative process for
the screened function. For values of q > 0.08, both iterative
and variational process with the screened functions yield better
results than the UWF, which shows the importance of screening
in the wavefunction, even for small q. The critical value q

for
the existence of the bound state, i.e., E(q

) = 0, is q

= 1.3964
for the variational procedure, q

= 1.3670 for the iterative


process, and q

= 1.1510 for the UWF. The critical dual


charge

in the boundunbound transition computed using


variational and iterative methods is

= 1.9176 and

=
2.2108, respectively. We recognize that the variational critical
parameters q

and

are more precise results (the best result


means a higher q

). In both the variational and iterative cases,


the SWF is far better than the UWF.
In Fig. 3 we plot the radius of equilibrium R
eq
as a function
of q for the three process. Curve (a): the variational process;
curve (b): iterative process; curve (c): UWF. Notice that as the
ThomasFermi parameter q increases so also does the inter-
atomic distance, i.e., the screening, which in fact facilitates
the ion separation. As we approach the critical region, R
eq
grows to reach the critical value: Curve (a) R

eq
= 3.9713 at
q

= 1.3964; curve (b) R

eq
= 4.3905 at q

= 1.3641; curve
Fig. 2. Binding energy as a function of the ThomasFermi parameter q (atomic
units). Curve (a): variational process; curve (b): iterative process; curve (c):
variational process for the UWF.
Fig. 3. Inter-protonic distance R as a function of the ThomasFermi parameter
q (atomic units). Curve (a): variational process; curve (b): iterative process;
curve (c): variational process for the UWF.
(c) R

eq
= 4.3117 at q

= 1.2184, we should remember that


the critical value for the atomic system is q

= 1.1906. For a
given q a better result means a large binding energy and a small
radius. However when we compare curves (b) and (c) we notice
that in the region 0.08 < q < 0.46, R
c
< R
b
, nevertheless
E
b
< E
c
. This is due to the fact that the set of parameters
, , R are enough to yield good energy, and not only R. In
any circumstance however curve (a) yields better E and R
eq
as expected. The nite value of our R

eq
is a balance between
its fast growth at the boundunbound transition and the system
aversion to the atomic limit
R R

eq
and E
H
> 0 > E
H
+
2
. (10)
In Fig. 4, we plot the charge and the screening parameter
as functions of q. This plot is critical in elucidating the
difference between the methods and the subtle physics within.
Curves (a) and (b) represent for the iterative and variational
procedures, respectively. Note that in both curves is an
increasing function of q. However, for the iterative method
the increase is continuous, whereas for the (more precise)
variational method it starts at a small value and the screening
becomes effective only for q 0.6 or 1.0. It is interesting
J.B. Diniz et al. / Solid State Communications 146 (2008) 169174 173
Fig. 4. Effective nuclear charge and screening parameter as functions of
the ThomasFermi parameter q (atomic units). (a), (b) Curves for obtained
using the iterative and variational processes, respectively. (c)(e) Curves for
obtained using the variational process for the UWF, the variational process and
the iterative process, respectively.
that only for charges smaller than the bare atomic charge does
the screening parameter become non-null for the variational
method. The smoothness of curve (a) is once again a limitation
of the recurrence relation given by Eqs. (3) and (5). Curves
(c)(e) display for the unscreened function and the variational
and the iterative methods, respectively. In curve (c), decreases
with q, with a nite critical

= 0.614 at the boundunbound


transition. In curve (d), rapidly decreases for q 0.6 and
there is synchronization between the change in behavior of
and . As soon as reaches the atomic charge, increases due
to the decrease in , which in turn decreases due to the increase
in . This super-freezing phenomenon for is well understood
in terms of the Bohr radius a
H
=
1
. For large , there is
a small Bohr radius for which the screening is not effective.
Note that this phenomenon of a rapid increase in screening
was not observed in the ThomasFermi atom [11]. It may be
present in more complex molecules as well. Finally, curves (d)
and (e) both show

= 0 at the boundunbound transition,


which means that the transition occurs when the electron
sees no charge at all. This relation is simple, intuitive, and
straightforward: no effective charge implies no bounding. This
result is very important for the interpretation of metalinsulator
transitions.
4. Conclusions
We studied the quantum screening phenomenon in the
hydrogen molecule in terms of charge symmetry aspects.
We demonstrated that the unscreened hydrogen molecule
has a hidden charge symmetry that breaks down in the
presence of the ThomasFermi parameter q. We used this
concept to understand the quantum screening effect in this
molecule. Moreover, the model can be extended for more
complex molecular structures. Our objective was rst to show
that a charge degeneracy effect occurs in simple molecular
structures as a starting point before considering more advanced
applications in condensed matter physics.
Using this charge degeneracy model, we obtained an
analytical expression for the ground state of a screened
H
+
2
molecule. This led to the wavefunction (r), and an
analytical expression for the energy, Eq. (9), with no adjustable
parameters. Screening in the potential breaks the charge
symmetry of the hydrogen atom, i.e., (, ) for q = 0.
This splitting means that a new interpretation of the charges
is required. We keep Eq. (3) as the nuclear charge that the
electron sees, in analogy to the unscreened hydrogen atom,
while becomes the dual charge. Moreover, as q increases,
decreases, as expected for an effective charge. The splitting
introduced by Eq. (5) is forced into the wavefunction, which
responds to the nuclear charge and to the new screening
parameter = . This qualitative improvement in the
wavefunction yields a quantitative gain with a better ground-
state energy. We observed that is an increasing function of q,
and for q = 0 we obtained = 0, as expected.
We also reported a super-freezing phenomenon in the
ThomasFermi H
+
2
molecule that has no counterpart in the
literature, and that the variational procedure yields a good
description of the entire screened process. The combination
of iterative and variational methods yields a simple, clear and
precise solution that can be used in more complex electronic
systems.
Acknowledgments
We are grateful for nancial support from Conselho Na-
cional de Desenvolvimento Cientco e Tecnol ogico-CNPq
(Brazil), Coordenac ao de Aperfeicoamento de Pessoal de
Nvel Superior-CAPES (Brazil), Fundac ao de Empreendi-
mentos Cientcos e Tecnol ogicos-FINATEC (Braslia), and
Financiadora de Estudos e Projetos-FINEP (Brazil). We would
like to thank Dr Mendeli Vainstein for careful reading of the
manuscript and suggestions for improvements.
References
[1] N.W. Ashcroft, N.D. Mermin, Solid State Physics, Saunders College
Publishing, 1976.
[2] H. Yukawa, Proc. Phys. Math. Soc. Jpn. 17 (1935) 48; 19 (1937) 1084.
[3] L. Hulthen, K.V. Laurikainen, Rev. Mod. Phys. 23 (1951) 1.
[4] G.M. Harris, Phys. Rev. 125 (1962) 1131.
[5] C. Lovelace, D. Massom, Nuovo Cimento 26 (1962) 472.
[6] C.R. Smith, Phys. Rev. A 134 (1964) 1235.
[7] H.M. Schey, J.L. Schwartz, Phys. Rev. B 139 (1965) 1428.
[8] G.J. Iafrate, L.B. Mendelssohn, Phys. Rev. 182 (1969) 244.
[9] F.J. Rogers, H.C. Graboske Jr., D.J. Harwood, Phys. Rev. A1 (1970) 1577.
[10] R.G. Sachs, M. Goeppert-Mayer, Phys. Rev. 53 (1938) 991.
[11] S.L. Garavelli, F.A. Oliveira, Phys. Rev. Lett. 66 (1991) 1310.
[12] O.A. Gomes, H. Chacham, J.R. Mohallem, Phys. Rev. A 50 (1994) 228.
[13] V.I. Yukalov, E.P. Yukalova, F.A. Oliveira, J. Phys. A 31 (1998) 4337.
[14] J.M. Ugalde, C. Sarasola, X. Lopez, Phys. Rev. A 56 (1997) 1642.
[15] E.R. Vrscay, Phys. Rev. A 33 (1986) 1433.
[16] C. Stubbins, Phys. Rev. A 48 (1993) 220.
[17] P.L. Taylor, B.-C. Xu, F.A. Oliveira, T.P. Doerr, Macromolecules 25
(1992) 1694.
[18] N.F. Mott, Rev. Modern Phys. 40 (1968) 667.
[19] N.F. Mott, MetalInsulator Transitions, Taylor and Francis, London,
1990.
[20] A.J. Peter, K. Navaneethakrishnan, Phys. Status Solidi B 220 (2000) 897.
[21] A.J. Peter, Solid State Commun. 129 (2004) 169.
[22] I.B. Spielman, W.D. Phillips, J.V. Porto, Phys. Rev. Lett. 98 (2007)
086401.
174 J.B. Diniz et al. / Solid State Communications 146 (2008) 169174
[23] D.M. Ceperley, B.J. Alder, Phys. Rev. B 36 (1987) 2092.
[24] A. Ferraz, F.A. Oliveira, M.A. Amato, Solid State Commun. 64 (1987)
1321.
[25] G.F. Chapline Jr., Phys. Rev. B 6 (1972) 2067.
[26] H.K. Mao, P.M. Beel, R.J. Hemley, Phys. Rev. 55 (1985) 99.
[27] R.J. Hemley, H.K. Mao, Phys. Rev. Lett. 61 (1988) 857.
[28] J. Diaz-Valdes, et al., Nucl. Instrum. Methods B 254 (2007) 69.
[29] N.R. Arista, Nucl. Instrum. Methods B 164 (2000) 108.
[30] R. Dez Mui no, A. Salin, Phys. Rev. B 62 (2000) 5207.
[31] I.S. Bitensky, V.Kh. Ferleger, I.A. Wojciechowski, Nucl. Instrum.
Methods B 125 (1997) 69.
[32] I.A. Wojciechowski, et al., Nucl. Instrum. Methods B 143 (1998) 473.
[33] Y. Susuki, Phys. Rev. A 56 (1997) 2918.
[34] J.E. Vald es, et al., Phys. Rev. A 68 (2003) 064901.
[35] J.P. Cotton, B. Farnoux, G. Jannink, R. Ober, J. Phys. A 6 (1973) 951.
[36] Y. Rabin, P. Pekarski, R. Bruisnma, Europhys. Lett. 24 (1993) 145.
[37] X. Schlagberger, J. Bayer, J.O. R adler, R.R. Netz, Europhys. Lett. 76
(2006) 346.
[38] M.H. Sorensen, N.A. Mortensen, M. Brandbyge, Appl. Phys. Lett. 91
(2007) 102105.
[39] M.L. Cohen, S.G. Louie, A. Zettl, Solid State Commun. 113 (2000) 549.
[40] F. Chen, W.D. Kirkey, M. Furis, M.C. Cheung, A.N. Cartwright, Solid
State Commun. 125 (2003) 617.
[41] C. Ciuti, G. Bastard, Solid State Commun. 133 (2005) 537.
[42] L.V Butov, Solid State Commun. 127 (2003) 89.
[43] M. Kellogg, J.P. Eisenstein, L.N. Pfeiffer, K.W. West, Solid State
Commun. 123 (2002) 515.
[44] D. Iannuzzi, M. Lisanti, J.N. Munday, F. Capasso, Solid State Commun.
135 (2005) 618.
[45] E. Moore, B. Gherman, D. Yaron, J. Chem. Phys. 106 (1997) 4216.
[46] R. Mcweeny, Coulsons Valence, third edition, Oxford University Press,
1979.

You might also like