Sie sind auf Seite 1von 8

JOURNAL OF CHEMICAL PHYSICS VOLUME 114, NUMBER 12 22 MARCH 2001

Low-temperature dynamics and spectroscopy in exohedral rare-gas C60


fullerene complexes
A. Ruiz, J. Hernández-Rojas, J. Bretón, and J. M. Gomez Llorentea)
Departamento de Fı́sica Fundamental II, Universidad de La Laguna, 38205 Tenerife, Spain
共Received 14 August 2000; accepted 5 January 2001兲
The adatom dynamics in exohedral C60 fullerene complexes of rare-gas atoms are studied with a
three degrees of freedom model. The eigenvalue problem of the corresponding quantum
Hamiltonian is solved and the electric-dipole spectra for ArC60, NeC60, and HeC60 in the
low-temperature range from 5 to 40 K are simulated. The most important spectral features are
related to the degree of angular anisotropy in the adatom–C60 interaction. The ArC60 and NeC60
complexes present very simple spectra which can be assigned in terms of three-mode oscillators; the
corresponding motion takes place in the deep hexagon wells 共also in the pentagon wells for NeC60)
of the interaction potential. On the contrary, the HeC60 complex shows more complicated spectra
with important tunneling effects due to the smaller angular anisotropy of the interaction. The onset
of almost free internal rotation takes place in this complex at rather low energies, and this gives rise
to a low-frequency rotational band in the spectra at temperatures above T⬃15 K. © 2001
American Institute of Physics. 关DOI: 10.1063/1.1350918兴

I. INTRODUCTION large polarizability values are not longer observed.5 The


same techniques were applied to the RbC60 complex and
The discovery and synthesis of C60 and other fullerenes provided, in this case, a value of the barrier activation energy
has given rise to a whole class of new molecular species of 20 meV for the rotational motion of the alkali metal on the
known as fullerene complexes. The simplest of these systems C60 surface.6
are formed by an atom or small molecule trapped either in- The internal rotational–vibrational dynamics are, there-
side 共endohedral complexes, X@C60) or outside 共exohedral fore, interesting aspects of these molecular systems. Sup-
complexes, XC60) the fullerene cage. Examples of X are ported by the latter experimental results, one may imagine
alkali and other metal atoms, rare-gas atoms, small diatomic
the adatom having a very complicated motion with a radial
molecules such as CO, etc. While endohedral complexes are
共i.e., out-of-surface兲 stretching mode, and two angular 共i.e.,
extremely difficult to produce in macroscopic quantities, the
on-surface兲 modes which may be either bendings, when the
exohedral ones can be obtained by standard procedures and
adatom is trapped in the surface potential wells, or internal
constitute the building blocks of doped fullerites. Very inter-
rotations when it has enough energy to cross over the surface
esting properties and important applications have been either
found or suggested for these new materials. All this has barriers; all these motions may further interact with the
stimulated the experimental study of exohedral complexes in fullerene vibrations and rotations. As is well known, molecu-
gas phase. For instance, photoionization time-of-flight mass lar spectroscopy is one of the most direct experimental ways
spectroscopy has been used to analyze the abundance of al- of obtaining detailed information on this motion. To our
kali and alkaline-earth Xn C60 complexes 共n in the range knowledge, no experimental spectra have been reported so
0–500兲;1,2 low-energy molecular-beam techniques have pro- far, thus the development of reliable theoretical models to
vided information on the interaction between C60 and Na predict them becomes of great relevance in understanding

clusters;3 electron photo-detachment experiments on Nan C60 these dynamics and in guiding future experimental work.
have given information on the electronic structure and inter- These are main goals of this paper.
action of these complexes.4 The interaction between the fullerene molecule and the
Recent experiments that use molecular beam deflection adatom is a determining factor in the molecular dynamics.
techniques have even revealed information on the dynamics Ab initio data on this interaction is sparse and almost inex-
of the adsorbed atom on the fullerene surface5,6. With these istent for exohedral complexes.7 However, if we restrict the
methods the polarizability of KC60 has been measured as a class of complexes to those with closed-shell electronic
function of temperature in the range from 300 to 483 K; the structure, semiempirical methods can be used to construct
very high values found for such electronic property are at- reliable potential energy surfaces.8–11 This has motivated our
tributed to the almost free skating of the potassium atom on choice of the exohedral complexes of rare-gas atoms, for
the C60 surface at those temperatures. At low temperatures which the interaction is of van der Waals-type. These are
the K atom is trapped in the surface potential wells and the also very convenient molecular systems from the experimen-
tal point of view. As a matter of fact, in the last two decades
a兲
Electronic mail: jmgomez@ull.es many spectroscopic techniques have been used in the study

0021-9606/2001/114(12)/5156/8/$18.00 5156 © 2001 American Institute of Physics

Downloaded 24 Jul 2009 to 62.76.169.62. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
J. Chem. Phys., Vol. 114, No. 12, 22 March 2001 Exohedral rare-gas C60 fullerene complexes 5157

of similar 共in principle兲 aromatic rare-gas van der Waals p r2 l2


complexes.12 H⫽ ⫹ ⫹V 共 r, ␪ , ␾ 兲 . 共2兲
2␮ 2␮r2
Theoretical rotational dipole and Raman spectra have al-
ready been predicted by our group for endohedral fullerene
complexes such as Li⫹ @C60, Na⫹ @C60, CO@C60, This approximation neglects the effects of the fullerene rota-
LiH@C60, and LiF@C60. 9,13–16 The strong confinement con- tion on the adatom motion. As shown in Refs. 9 and 13,
straints existing in these molecular systems simplifies con- when the spectrum is dominated by radial and bending vi-
siderably their numerical treatment by allowing the separa- brational bands 共as will be the general case in this work兲 the
tion of the internal rotational motion of the guest species effect of the fullerene rotations is just a rotational broadening
from all other motions. This is not possible, in general, for of the spectral lines, which can be very well reproduced by
the exohedral complexes which will, therefore, require a si- coarse-graining the spectra with an appropriate convolution
multaneous treatment of the adatom rotations and vibrations. function.
In this work we present electric-dipole spectra for the The reliability of our theoretical analysis will depend
exohedral complexes HeC60, NeC60, and ArC60 in the tem- primarily on the accuracy of the C60-adatom effective inter-
perature range from 5 to 40 K. The analysis of these spectra action potential V. Symmetry constrains considerably the
will display the important effect of temperature and show form of V; namely, this belongs to the totally symmetric
how the adatom on-surface dynamics can change from bend- irreducible representation of the I h point group, which is the
ing vibrations to almost free rotations. In Sec. II we present representation and symmetry group of the fullerene elec-
the effective Hamiltonian for the rotational–vibrational mo- tronic and vibrational ground states used in the successive
tion of the adatom. Section III gives some details of the adiabatic approximations leading to Hamiltonian in Eq. 共2兲.
calculations and presents the structure of the eigenvalue We can take advantage of this symmetry by expanding the
spectrum for the effective Hamiltonian. Section IV presents angular dependence of the potential V in the spherical har-
and analyzes the electric-dipole spectra. The main conclu- monic basis Y L,M
sions are put together in Sec. V.
V 共 r, ␪ , ␾ 兲 ⫽2 ␲ 1/2 兺
L,M
c L,M 共 r 兲 Y L,M 共 ␪ , ␾ 兲 . 共3兲

II. AN EFFECTIVE ROTATIONAL–VIBRATIONAL


HAMILTONIAN Then only L subspaces that contain the totally symmetric
representation of the I h group can appear; these are L⫽0, 6,
The effective Hamiltonian describing the dynamics of an 10, 12, 16, etc. As will be seen later, this expansion is very
atom interacting with a C60 fullerene was derived in Refs. 9 convenient also for computational reasons.
and 13 共see also Ref. 17兲. In the body fixed frame it takes the As mentioned in the Introduction, ab initio data on the
form adatom–fullerene interaction is sparse and almost inexistent
1 2 p r2 l2 for exohedral complexes. However, in our case the total in-
H rv⫽ K ⫹ ⫹ ⫹V 共 r, ␪ , ␾ 兲 , 共1兲 teraction energy, which occurs between two closed shell sys-
2I c 2␮ 2␮r2
tems, is dominated by dispersion-repulsion terms. In such
where K⫽J⫺l is the fullerene angular momentum operator, case, there are in the literature several semiempirical expres-
which is the difference between the total angular momentum sions for V. 8,9,11 Here we will use in the potential Y LM ex-
J and the atom angular momentum l: I c ⫽ 32 60M R 2 is the pansion an expression of us which consists of two-body
fullerene moment of inertia, with R being the fullerene radius Lennard-Jones interactions between the fullerene C atoms
共3.55 Å兲 and M the C atom nuclear mass; p r is the adatom and the adatom.8,9 This form has been used in numerical
radial momentum operator and ␮ ⫽60M m/(60M ⫹m) its re- studies of the classical dynamics of atoms trapped inside and
duced mass; finally, V is the potential of the effective inter- outside the C60 fullerene, and reproduces accurately enough
action between the adatom and the fullerene, which depends the available experimental and ab initio data for both en-
on the adatom spherical coordinates r, ␪, ␾. dohedral and exohedral complexes. We then find that a trun-
Three approximations were performed to derive Eq. 共1兲: cation of the expansion in Eq. 共3兲 at L⫽16 gives very good
The usual Born–Oppenheimer separation of the electronic account of the interaction, specially of those features more
degrees of freedom, the adiabatic separation of the C60 vibra- relevant in our study such as main well depths, frequencies
tions, and the standard rigid-rotor approximation for the and barrier heights.
fullerene molecule. Figure 1 presents a view of the potential angular depen-
The relatively large value of the fullerene moment of dence at fixed r for the three rare-gas exohedral complexes.
inertia I c makes a full quantal treatment of the fullerene ro- We observe minima at the pentagon and hexagon centers of
tation a formidable task; the problem lies in the large number the C60 truncated icosahedrum, hexagon being always more
of J values involved in the rotational spectra, even at the low stable. Figure 2 displays the r-dependence of some of the
temperatures considered here 共J up to ⬃300兲. One must try Y LM expansion coefficients c L,M (r), and the total interaction
then a further approximation. This is the fullerene infinite- along the C 3 symmetry axis. As can be seen the isotropic
mass approximation discussed in Ref. 9, which is based on (L⫽0) term dominates the interaction at intermediate-large
the large ratio between fullerene and adatom masses. The distances and practically determines radial equilibrium posi-
effective Hamiltonian then becomes tions r e and radial vibrational frequencies. This particular

Downloaded 24 Jul 2009 to 62.76.169.62. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
5158 J. Chem. Phys., Vol. 114, No. 12, 22 March 2001 Ruiz et al.

FIG. 1. The adatom–fullerene interaction V(r, ␪ , ␾ ) as a function of the


polar angle ␪ for ␾ ⫽0 and r⫽r e . One of the C 5 symmetry axis has been
chosen along the z axis of the reference frame, and one of the neighbor C 2
symmetry axis lies on the zx plane. For ␪ ⫽0 the atom is on the pentagon
center; the deep wells occur on the hexagon centers, and maximum values
appear on the C atoms. Equilibrium positions r e are given in Table I.

behavior of the fullerene–atom interaction will simplify con- FIG. 2. Radial dependence 共narrow lines兲 of some of the coefficients
siderably the calculation of the spectrum, as will be seen c L,M (r) in the potential expansion 关Eq. 共3兲兴. The reference frame axes have
next. been chosen as in Fig. 1. Labels are 共L, M兲. The 共0, 0兲 coefficient is the
spherically isotropic contribution to the potential. The thick line gives the
values of the total interaction along the C3 symmetry axis and calculated
III. THE CALCULATION OF EIGENVALUES AND from the potential expansion truncated at L⫽16.
EIGENFUNCTIONS
Led by the differences in radial and angular frequencies
at the potential minima we will write the effective Hamil- The coupling term in Eq. 共4兲 has, consequently, the fol-
tonian in Eq. 共2兲 as the sum lowing expression:
H⫽H vib⫹H rot⫹H coup ,
where H vib and H rot are, respectively, pure radial and angular
共4兲
H coup⫽ 冉
l2 1 1

2 ␮ r 2 r 2e 冊
⫹V coup共 r, ␪ , ␾ 兲 , 共9兲
Hamiltonians given by
where
p r2
H vib⫽ ⫹V vib共 r 兲 , 共5兲
2␮ V coup共 r, ␪ , ␾ 兲 ⫽2 ␲ 1/2 兺 ⬘ 关 c LM 共 r 兲 ⫺c LM 共 r e 兲兴 Y LM 共 ␪ , ␾ 兲
L,M
with

V vib共 r 兲 ⫽2 ␲ 1/2 兺 c LM 共 r 兲 Y LM 共 ␪ 0 , ␾ 0 兲 共6兲


⫺2 ␲ 1/2 兺 ⬘ c LM 共 r 兲 Y LM 共 ␪ 0 , ␾ 0 兲 .
L,M
共10兲
L,M
We have produced a convenient basis set 兵 兩 n,l,m 典 其 for
and
the diagonalization of the effective Hamiltonian H, by direct
l2 product of the two eigenstate basis sets corresponding to the
H rot⫽ ⫹V rot共 ␪ , ␾ 兲 , 共7兲 reference Hamiltonians H vib( 兵 兩 n 典 其 ) and H rot( 兵 兩 l,m 典 其 ). This
2 ␮ r 2e
basis is then truncated so as to include all eigenstates of
with H vib⫹H rot with eigenenergies E vib⫹E rot⭐E cut , where E cut is
an energy cutoff which is chosen high enough to ensure con-
V rot共 ␪ , ␾ 兲 ⫽2 ␲ 1/2 兺 ⬘ c LM 共 r e 兲 Y LM 共 ␪ , ␾ 兲 .
L,M
共8兲 vergence of our low-temperature spectra. The angular mo-
mentum states 兩l, m典 were used in the diagonalization of H rot
In these equations ␪ 0 and ␾ 0 define a reference orientation, and the one-dimensional Hamiltonian H vib was diagonalized
which we have chosen along a C 3 symmetry axis 共going using a standard Numerov method. Then all the angular in-
through an hexagon center兲, and r e is the equilibrium radial tegrals required in the calculation of matrix elements were
position along such orientation; in the C 3 case, r e is the analytically obtained in terms of the 3 j symbols, and the
global-equilibrium distance of the potential surface. In Eq. radial ones were evaluated numerically. Symmetry is used to
⬘ excludes the L⫽0 constant term.
共8兲 the sum 兺 L,M separate the H rot and full H diagonalizations in blocks be-

Downloaded 24 Jul 2009 to 62.76.169.62. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
J. Chem. Phys., Vol. 114, No. 12, 22 March 2001 Exohedral rare-gas C60 fullerene complexes 5159

TABLE I. Some details of the calculations. Values of r e give the adatom-fullerene equilibrium distances along
the C 3 symmetry axis; they correspond to global minima of the interaction potential. The number of l values
included in the diagonalization of H rot is denoted n l 共all m values are included for each l兲. E cut gives a cut-off
energy; all calculated eigenstates of H rot⫹H vib with energies E rot⫹E vib⭐E cut are included in the diagonalization
of the effective Hamiltonian H 共this cut-off restriction was not used for the A symmetry blocks兲. The total
number of these states is denoted n t ; the number of states in each symmetry representation are given in
parentheses.

r e 共Å兲 nl E cut 共cm⫺1兲 n t (A g ⫹T g ⫹G g ⫹H g ⫹A u ⫹T u ⫹G u ⫹H u )

Ar 6.67 112 ⫺220 13 776(1098⫹1659⫹1620⫹2675⫹828⫹1959⫹1592⫹2345)


Ne 6.39 90 ⫺40 6966(574⫹792⫹816⫹1365⫹399⫹1044⫹816⫹1160)
He 6.21 72 100 3768(224⫹432⫹480⫹780⫹140⫹603⫹464⫹645)

longing to the different I h irreducible representations, so that associated to the bending motion on the fullerene surface.
the larger matrix found in our calculation has dimensions The pentagon potential wells occur at higher energies, so that
2675⫻2675. Table I presents more details of the computa- possible states influenced by them do not appear in Fig. 3
tion. neither will they contribute in the low temperature spectra.
Figures 3–5 display the low-energy eigenvalue spectrum The He complex has a completely different and more
of the effective Hamiltonian for each one of our three complicated structure. The analogues of the level sets in
fullerene complexes. The Ar complex shows the simplest ArC60 have split completely in HeC60, which indicates im-
spectrum with lower-energy levels grouped in sets of 20 and portant tunneling effects at low energies and the possibility
40 almost degenerate states; thus we have either one or two of nearly free internal rotational motion at higher energies.
states from each set for each one of the 20 hexagon wells of All these features are a consequence of the much lower ro-
the interaction potential. From this level structure and the tational barriers in the interaction potential and the small He
shape of the wave functions 共see an example in Fig. 6兲 we mass. Moreover, hexagon and pentagon potential wells are
conclude that at these lower energies the atom can get very close in energy 共see Fig. 1兲, and both participate then in
trapped in one of those hexagon wells; since the splitting the level structure. For instance, the lowest 20 states that
between the levels in a set, which is due to quantum tunnel- appear grouped in three sets in Fig. 5 correspond to states
ing, is so small, such trapping occurs in fact for quite long whose wave functions are localized on the hexagon wells,
times. In each well the motion corresponds to a three dimen- while the next two sets contain together 12 levels with wave
sional oscillator with a C 3 symmetry; the possible degenera- functions localized on pentagon wells of the interaction po-
cies in this symmetry group are one and two, in accordance tential surface 共the number of these wells is 12兲. The wave
with the numbers given previously. The high-frequency functions for higher levels present either localization on both
mode of this oscillator corresponds basically to the radial wells or a more delocalized structure, which is another indi-
vibration, while the other two, more strongly coupled, are cation of more or less hindered internal rotational motion.
The NeC60 system represents a situation intermediate be-
tween that of the other two complexes. Most energy levels

FIG. 3. Low-energy eigenvalue spectrum for the ArC60 complex. Labels on


the right indicate the number of states contained in each energy level. The n
value on the left gives the vibrational excitation of the radial motion and FIG. 4. Low-energy eigenvalue spectrum for the NeC60 complex. Notation
labels the lowest energy state in which such an excitation appears. is as in Fig. 3.

Downloaded 24 Jul 2009 to 62.76.169.62. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
5160 J. Chem. Phys., Vol. 114, No. 12, 22 March 2001 Ruiz et al.

FIG. 7. Electric dipole spectra of ArC60 at different temperatures. These


have been coarse grained with a temperature-dependent convolution func-
tion to simulate the broadening effect due to the overall fullerene complex
rotation. The first peak/band at ⬃13 cm⫺1 corresponds to transitions with
one-quantum excitation of the bending motion in the hexagon potential
FIG. 5. Low-energy eigenvalue spectrum for the HeC60 complex. The well wells. At higher temperatures such transitions can take place not only from
type in which wave functions are localized are given on the right for the two the ground level but also from excited ones; then mode-coupling and anhar-
lowest set of levels; higher states show localization in both hexagon and monicities in the bending motion make the narrow low-temperature peak
pentagon wells in each wave function. Labels on the left give the symmetry evolve into a broader band. The peak at ⬃37 cm⫺1 corresponds to transi-
of the states in those two level sets. tions with one-quantum excitation of the radial motion. Some low-intensity
harmonics are also observed.

can be closely grouped in sets of 20 and 40 states on one


hand, and 12 and 24 states on the other; the former corre-
spond, as in ArC60, to trapping in the hexagon wells, and the and
latter to trapping in the pentagon wells; the two well types 1
are here, as in HeC60, close in energy 共see Fig. 1兲. Tunneling I共 ␻ 兲⫽
Z 兺
i, j
exp关 ⫺ ␤ E j 兴 兩 di j 兩 2 ␦ 关共 E i ⫺E j 兲 /ប⫺ ␻ 兴 . 共13兲
effects are unimportant at the lower energies included in Fig.
4 and more important at higher energies. In Fig. 6 we present In these equations C is a constant which depends on the
two of the low-energy wave functions for this complex, one density of molecules, molecule parameters, and electromag-
localized on the hexagon potential wells and the other on the netic constants;9 Z is the partition function, E i the eigenval-
pentagon wells. ues of Hamiltonian in Eq. 共4兲, di j are matrix elements of the
effective electric dipole moment operator d⫽r, with r being
IV. THE ELECTRIC-DIPOLE SPECTRA the atom position vector operator per unit charge with re-
spect to the fullerene center. We are assuming a linear de-
For the calculation of the low-temperature dipole spectra
pendence of the dipole moment on the position vector opera-
we have used the usual expression for the absorption coeffi-
tor.
cient as a function of the angular frequency ␻
For temperatures in the range from 0 to 40 K our spectra
␣ 共 ␻ 兲 ⫽CS 共 ␻ 兲 , 共11兲 are well converged. Figures 7–9 display these spectra coarse
where grained, as discussed in Sec. II, with a temperature-
dependent convolution function representing the fullerene
S 共 ␻ 兲 ⫽ប ␻ 关 1⫺exp共 ␤ ប ␻ 兲兴 I 共 ␻ 兲 , 共12兲 rotational broadening. Here again the ArC60 complex pre-
sents the simplest spectrum 共Fig. 7兲. It is a vibrational spec-
trum whose lines can be assigned to transitions between the
stretching and bending states in each well. At the lowest
temperature the relevant transitions originate from the
ground level set. When temperature increases transitions
from excited levels also appear; their frequencies change
slightly due to potential anharmonicity, which gives rise to
broader vibrational bands.
The NeC60 共Fig. 8兲 complex show similar spectra, with
vibrational transitions. However, both pentagon and hexagon
FIG. 6. Stereographic projections of the angular dependence of the wave potential wells contribute in this case, although the transi-
functions for three A g states; the one in ArC60 shows localization on the
tions from the latter turn out to be more intense. At higher
hexagon potential wells; the other two for NeC60 present localization on
either hexagon or pentagon wells. The change in sign of the wave functions temperatures, states closer to the potential barriers, where
is indicated using different line types. anharmonicity is larger, contribute; mode-coupling and tun-

Downloaded 24 Jul 2009 to 62.76.169.62. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
J. Chem. Phys., Vol. 114, No. 12, 22 March 2001 Exohedral rare-gas C60 fullerene complexes 5161

FIG. 8. Electric dipole spectra for NeC60 at different temperatures, and


coarse grained as in Fig. 7. The main peak at ⬃12 cm⫺1 in the spectrum at
T⫽5 K corresponds to a transition with one-quantum excitation of the bend-
ing motion in the hexagon potential wells, from the ground level set; its left
satellite peaks have similar origin but from excited bending states. At higher FIG. 10. Stereographic projections of the angular dependence of the wave
temperatures mode-coupling, anharmonicities and tunneling effects make functions corresponding to the two most relevant state pairs involved in the
these peak evolve into a broad bending vibrational band, to which one- bending transition peak observed in the HeC60 complex at ⬃16 cm⫺1 共see
quantum excitations of the bending motion in the pentagon potential wells Fig. 9兲. Top figures are for final states in the G g quadruplet. Bottom figures
now also contribute. The intense peak at ⬃35 cm⫺1 corresponds to transi- give the corresponding initial approximate eigenstates obtained as described
tions with one-quantum excitation of the radial motion in the hexagon well; in the Appendix. The change in sign of the wave functions is indicated using
its small satellite corresponds to the analog transition in the pentagon well. different line types. As we see the transition can be assigned to a two-quanta
excitation of a bending mode orthogonal to the internal rotational motion
that connects the five hexagon wells around each pentagon. This transition
neling effects are also noticeable in this energy region, and brings the adatom very close to the top of the hexagon–pentagon potential
all this makes the vibrational bands broaden considerably. barriers, where it couples with the motion in the pentagon wells.
The most complicated spectra are shown by the HeC60
complex 共Fig. 9兲. Tunneling effects are already important at spectra. One is what looks like a pure rotational band in the
the lowest temperatures so that the vibrational transitions lowest frequency range. Its origin is in the internal rotational
seen in the two previous cases split now in many lines. When motion that takes place above the surface potential barriers.
temperature increases, two main features characterize the A confirmation of this interpretation is given by the similar-
ity between this band and the rotational spectrum of a true
free rotor with a rotational constant calculated at the mini-
mum of the spherical potential term c 00(r) 共see the insets in
Fig. 9兲.
The other relevant feature of the HeC60 spectra at higher
temperatures is the sharp peak around ⬃16 cm⫺1 which can
be assigned dominantly to a set of transitions from 10 states,
very close in energy, to the same G g quadruplet. The wave-
functions corresponding to these initial states present local-
ization on the hexagon wells, while those corresponding to
the final G g states are localized on both pentagon and hexa-
gon wells. By choosing proper linear combinations of the
initial states, as performed in the Appendix, we were able to
reduce all these transitions to only four carrying all the in-
tensity. The new states obtained in this way, which can be,
therefore, taken as approximate eigenstates, present a par-
ticular localization structure, which is displayed in Fig. 10,
FIG. 9. Electric dipole spectra for HeC60 spectra at different temperatures, together with that of the G g final states. The localization on
and coarse grained as in Fig. 7. At the two lowest temperatures, bending the five hexagon wells around a pentagon which we can ob-
transitions split in many peaks due to strong tunneling between different
serve in this Fig. 10 is indicating that the motion involved in
hexagon and pentagon wells. Above T⫽15 K, a low-frequency internal ro-
tational band appears; this assignment is confirmed by the inset figures, the sharp spectral feature can be represented as a precession
which give the pure rotational spectrum for a He atom moving as a free 共strongly mediated by quantum tunneling兲 of the atom posi-
rigid-rotor on the C60 fullerene surface. The peak at ⬃16 cm⫺1 is assigned to tion vector r around the C 5 symmetry axis. This precession
a two-quanta excitation of a bending mode orthogonal to an internal rota-
can be accompanied by the oscillation of the r-C 5 angle, and
tional motion that connects the five hexagon wells around each pentagon.
The peak at ⬃53 cm⫺1 is a combination of the previous frequency and the it is the excitation of this bending motion which gives rise to
frequency of the radial vibrational motion. the narrow peak observed in the spectra. From the nodal

Downloaded 24 Jul 2009 to 62.76.169.62. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
5162 J. Chem. Phys., Vol. 114, No. 12, 22 March 2001 Ruiz et al.

structure of the final G g states we deduce that this transition takes place now at rather low energies, and this gives rise to
takes place with interchange of two quanta of the bending a low-frequency internal rotational band in the spectra at
motion. Since this analysis has been applied to the compo- temperatures above T⬃15 K. An intense bending peak also
nent of the dipole moment operator on the C 5 symmetry appears; if the laser field is chosen along the C 5 symmetry
axis, it corresponds then to the linear response of the system axis, this peak can be easily assigned to a two-quanta exci-
to an oscillating electric field along that axis direction. The tation of a bending mode orthogonal to the internal rotational
analyses for the other two orthogonal components of dipole motion that connects the five hexagon wells around each
moment operator are somewhat more involved. However, as pentagon.
discussed in the Appendix, one can demonstrate using sym- Concluding, our results demonstrate that low-
metry arguments the equivalence of all three components. temperature dipole spectra of rare-gas exohedral fullerenes
The higher-frequency peaks appearing in the HeC60 spectra complexes are a source of very detailed and relevant infor-
at ⬃53 cm⫺1 and ⬃69 cm⫺1 have the same bending origin, mation on the adatom dynamics and on its interaction with
but in these cases one and two quantum excitations of the the C60 fullerene.
radial vibrational motion also take place, respectively.

V. CONCLUSIONS ACKNOWLEDGMENTS

Recent gas-phase experiments on exohedral fullerene This work has been supported by a grant from the Span-
complexes are providing interesting data on the interaction ish ‘‘Dirección General de Enseñanza Superior’’ 共Grant No.
between the adatom and the C60 fullerene. This interaction PB97-1479-C02-01兲. A. R. thanks ‘‘Consejerı́a de Educa-
determines the adatom dynamics, which take place on the ción, Cultura y Deportes del Gobierno de Canarias’’ for a
C60 spherical microsurface. Rotational–vibrational spectra fellowship. We thank Dr. O. Roncero for providing us with
are a primary experimental source of information on such the code for the resolution of the radial Schrödinger equa-
dynamics. Since these are not yet available, we have proceed tion.
in this work to their theoretical evaluation and study for the
complexes of rare-gas atoms. Existing experimental tech-
APPENDIX
niques to obtain the spectra of similar aromatic rare-gas van
der Waals complexes could be in principle applied to the The most intense bending peak in the HeC60 spectrum
fullerene complexes, thus making our choice of particular was assigned to a group of transitions from a set of ten close-
interest. in-energy states (2T u ⫹G u ) to a G g quadruply degenerate
By adiabatic separation of the C60 vibrational motion state. In order to establish the dynamical origin of such peak
and freezing its slow rotations, we were able to reduce the we have proceed to find four approximate eigenstates by lin-
low-energy dynamics to a problem with three degrees of ear combinations of the original ten initial states, so that the
freedom, two angular ones representing the on-surface mo- transitions from these new initial states will carry all the
tion of the adatom on the C60 fullerene surface, and a radial intensity of the spectral peak.
one giving the out-of-surface vibration. The interaction po- Let us denote with ␹ i each one of the four G g final states
tential between the rare-gas atom and the fullerene is and with ␾ i each one of the ten initial states. For each com-
strongly constrained by the molecular symmetry; this to- ponent d ␣ ( ␣ ⫽x,y,z) of the dipole moment operator we de-
gether with its van der Waals nature make quite reliable its fine the four approximate eigenvalues as the linear combina-
estimation by semiempirical arguments. The eigenvalue tions
problem for the effective Hamiltonian has been solved by 10
1
expanding the angular dependence of the wave function in
spherical harmonics and the radial one using a Numerov
兩␺ j典⫽ 兺 兩 ␾ 典 c 共 j⫽1,4兲 ,
N j i⫽1 i i j
共A1兲
method. The eigenfunctions and eigenvectors obtained are
where
accurate enough to produce converged electric-dipole spectra
in the low-temperature range from 0 to 40 K. c i j ⫽ 具 ␾ i兩 d ␣兩 ␹ j 典 , 共A2兲
The ArC60 complex presents very simple spectra that can and 1/N j is the normalization constant with

冋兺 册
be assigned in terms of a three-dimensional oscillator with
10 1/2
C 3 symmetry; the corresponding motion takes place in the
deep hexagon wells of the interaction potential; tunneling N j⫽ 兩 c i j兩 2 . 共A3兲
i⫽1
between these wells is negligible at these low temperatures.
NeC60 has similar spectra; however, since the pentagon wells The dipole-moment matrix elements between these four
are now deep enough, trapping is also possible there at our new initial states and the ␹ i final states can be written then in
lowest temperatures; excitation of the vibrational motion that the following form:
takes place in these pentagon wells gives rise to new lines in
具 ␺ k 兩 d ␣ 兩 ␹ l 典 ⫽ 具 ␹ k 兩 D̂ ␣ 兩 ␹ l 典 共 k,l⫽1,4兲 , 共A4兲
the spectra, although much less intense. At higher tempera-
ture these spectra reflect the increase of anharmonicity, mode where the operator D̂ ␣ is
coupling and tunneling effects. Finally, the HeC60 complex 10
shows quite complicated spectra which reflect the important
tunneling effects. The onset of almost free internal rotation
D̂ ␣ ⫽ 兺 d ␣兩 ␾ i 典具 ␾ i兩 d ␣ .
i⫽1
共A5兲

Downloaded 24 Jul 2009 to 62.76.169.62. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp
J. Chem. Phys., Vol. 114, No. 12, 22 March 2001 Exohedral rare-gas C60 fullerene complexes 5163

Since the four final ␹ i states are degenerate, we can al- 6


Ph. Dugourd, R. Antoine, D. Rayane, E. Benichou, and M. Broyer, Phys.
Rev. A 62, 011201共R兲 共2000兲.
ways fix them so that D̂ ␣ is diagonal in this set. In this way 7
J. Cioslowski, Electronic Structure Calculations on Fullerenes and their
each new initial state ␺ i will be connected by the dipole Derivatives 共Oxford University Press, New York, 1995兲, and references
moment operator to just one final state ␹ i . The overall line therein; A. S. Hira and A. K. Ray, Phys. Rev. A 52, 141 共1995兲; S.
intensity of the four transition coincides then with that of the Patchkovski and W. Thiel, J. Chem. Phys. 106, 1796 共1997兲.
8
original transitions, i.e., J. Bretón, J. González-Platas, and C. Girardet, J. Chem. Phys. 99, 4036
共1993兲; J. Hernández-Rojas, J. Bretón, and J. M. Gomez Llorente, Chem.
4 10 4 Phys. Lett. 222, 88 共1994兲; S. Iglesias-Groth, J. Bretón, and C. Girardet,
兺 兩 具 ␺ j 兩 d ␣兩 ␹ j 典 兩 ⫽ i⫽1
j⫽1
2
兺 j⫽1
兺 兩 具 ␾ i兩 d ␣兩 ␹ j 典 兩 2. 共A6兲
9
Chem. Phys. 237, 285 共1998兲.
J. Hernández-Rojas, J. Bretón, and J. M. Gomez Llorente, Chem. Phys.
Lett. 237, 115 共1995兲; J. Chem. Phys. 104, 1179 共1996兲.
The three components of the dipole moment operator 10
A. Ruiz, J. Hernández-Rojas, J. Bretón, and J. M. Gomez Llorente, J.
belong to the same I h group representation (T u ). Besides, Chem. Phys. 109, 3573 共1998兲.
11
one can go from one component to the other using the rota- G. Cardini, P. Procacci, P. R. Salvi, and V. Schettino, Chem. Phys. Lett.
tion operators R x , R y , R z , which, since they belong to the 200, 39 共1992兲; J. Low Temp. Phys. 19, 562 共1993兲; L. Pang and F.
Brisse, J. Phys. Chem. 97, 8562 共1993兲; C. J. Williams, M. A. Whitehead,
T g representation, do not mix initial with final states. Then, and L. Pang, ibid. 97, 11652 共1993兲; A. L. R. Bug, A. Wilson, and G. A.
one can easily demonstrate that the analysis performed for Voth, ibid. 96, 7864 共1992兲; M. S. Son and Y. K. Sung, Chem. Phys. Lett.
one of the d ␣ components may be conveniently transformed 245, 113 共1995兲.
to analyze completely the other two. In this way, one can
12
S. Leutwyler and J. Jortner, J. Phys. Chem. 91, 5558 共1987兲; H. J. Neusser
and H. Krause, Chem. Rev. 94, 1892 共1994兲; Th. Brupbacher, J. Makare-
show, for instance, that each one of the three components wicz, and A. Bauder, J. Chem. Phys. 101, 9736 共1994兲; P. M. Maxton, M.
gives exactly the same contribution to the spectrum. W. Schaeffer, S. M. Ohline, W. King, V. A. Venturo, and P. M. Felker,
ibid. 101, 8391 共1994兲; R. Sussmann, R. Neuhausser, and H. J. Neusser,
1
T. P. Martin, N. Malinowski, U. Zimmermann, U. Näher, and H. Schaber, ibid. 103, 3315 共1995兲; W. Kim and P. M. Felker, ibid. 107, 2193 共1997兲;
J. Chem. Phys. 99, 4210 共1993兲. R. Neuhausser, J. Braun, H. J. Neusser, and A. van der Avoird, ibid. 108,
2
U. Zimmermann, N. Malinowski, U. Näher, S. Frank, and T. P. Martin, 8408 共1998兲.
13
Phys. Rev. Lett. 72, 3542 共1994兲. J. Hernández-Rojas, J. Bretón, and J. M. Gomez Llorente, J. Chem. Phys.
3
A. A. Scheidemann, V. V. Kresin, and W. D. Knight, Phys. Rev. A 49, 104, 5754 共1996兲.
14
R4293 共1994兲; V. V. Kresin, A. A. Scheidemann, and W. D. Knight, ibid. J. Hernández-Rojas, J. Bretón, and J. M. Gomez Llorente, J. Chem. Phys.
49, 2696 共1994兲; V. V. Kresin, V. Kasperovich, G. Tikhonov, and K. 105, 4482 共1996兲.
15
Wong, ibid. 57, 383 共1998兲. J. Hernández-Rojas, J. Bretón, and J. M. Gomez Llorente, J. Phys. Chem.
4
B. Palpant, A. Otake, F. Hayakawa, Y. Negishi, G. H. Lee, A. Nakajima, Solids 58, 1689 共1997兲.
and K. Kaya, Phys. Rev. B 60, 4509 共1999兲. 16
J. Hernández-Rojas, A. Ruiz, J. Bretón, and J. M. Gomez Llorente, Int. J.
5
D. Rayane, R. Antoine, Ph. Dugourd, E. Benichou, A. R. Allouche, M. Quantum Chem. 65, 655 共1997兲.
Aubert-Frécon, and M. Broyer, Phys. Rev. Lett. 84, 1962 共2000兲. 17
G. Brocks and D. van Koeven, Mol. Phys. 63, 999 共1988兲.

Downloaded 24 Jul 2009 to 62.76.169.62. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/jcp/copyright.jsp

Das könnte Ihnen auch gefallen