Sie sind auf Seite 1von 8

High resolution infrared spectroscopy and ab initio calculations of HCNH 2 /D 2 binary

complexes
D. T. Moore, M. Ishiguro, L. Oudejans, and R. E. Miller

Citation: The Journal of Chemical Physics 115, 5137 (2001); doi: 10.1063/1.1394743
View online: http://dx.doi.org/10.1063/1.1394743
View Table of Contents: http://scitation.aip.org/content/aip/journal/jcp/115/11?ver=pdfcov
Published by the AIP Publishing




















This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
129.72.2.27 On: Fri, 15 Nov 2013 05:20:16
High resolution infrared spectroscopy and ab initio calculations
of HCNH
2
D
2
binary complexes
D. T. Moore,
a)
M. Ishiguro,
b)
L. Oudejans, and R. E. Miller
Department of Chemistry, University of North Carolina, Chapel Hill, North Carolina 27599
Received 10 May 2001; accepted 25 June 2001
High-resolution infrared laser spectroscopy has been used to study HCNH
2
and HCND
2
complexes in the gas phase. The experimental results are compared with ab initio calculations that
are also reported here. The latter calculations reveal two prominent minima on the potential surface,
one corresponding to a T-shaped complex with the H
2
at the hydrogen end of the HCN and the
other a linear complex with the H
2
H-bonded to the nitrogen. The latter minimum is the global
minimum on the surface, in agreement with the fact that this structure is observed experimentally for
both o-H
2
and p-D
2
. 2001 American Institute of Physics. DOI: 10.1063/1.1394743
INTRODUCTION
The spectroscopy of weakly bound complexes has
emerged as a powerful approach for the quantitative study of
intermolecular forces. The tremendous advances in recent
years have provided detailed structural and dynamical infor-
mation for a wide range of species.
1
In parallel with the
experimental advances, theoretical methods have developed
to the point where exact quantum methods can now be used
to calculate the spectra of many of these complexes directly
from the potential energy surface.
2
Due to the weakness of
the bonds associated with these complexes, the intermolecu-
lar vibrational motion is often highly anharmonic and the
associated spectra contain detailed information on the poten-
tial surface.
As a part of our ongoing study of cluster growth and
dynamics in liquid helium nanodroplets see companion
paper
3
we carried out the present high resolution gas phase
study of the binary H
2
/D
2
HCN complexes. Complexes
containing hydrogen are particularly interesting because the
low mass of the H
2
leads to large zero point energies and
hence weak binding. A few complexes containing hydrogen
have been studied previously, including the detailed spectro-
scopic and dynamical studies of H
2
HF Refs. 47 and
H
2
HCl Ref. 8 complexes as well as various isotopic
variations. In all of these cases, the H
2
is bound to the hy-
drogen end of the HXXF, Cl molecule. The correspond-
ing T-shaped conguration optimizes the electrostatic in-
teraction of the molecules, in that it is favorable for both the
dipolequadrupole and quadrupolequadrupole interactions.
This is most vividly illustrated in the case of H
2
HF, where
only the ortho complex is observed, given that these electro-
static interactions are missing for para hydrogen (J0),
where the quadrupole moment is averaged to zero. In con-
trast, the lower zero point energies associated with the
D
2
HF and H
2
HCl complexes make both the ortho and
para complexes sufciently stable to be observed
experimentally.
6,8
Detailed theoretical studies have also been
reported for the H
2
,D
2
HF systems, including a quantitative
ab initio potential surface, detailed bound state calculations,
photofragment nal state distributions and dissociation
energies.
914
Infrared spectra have also been reported for the
hydrogen-bonded CO-H
2
Refs. 1518 and H
2
OH
2
Ref.
19 complexes, in both cases the oxygen acting as the proton
acceptor.
Our rst observation of the HCNH
2
complex was in
helium droplets, as presented in a companion paper.
3
The
corresponding results indicated that ortho-H
2
and para-H
2
bind at opposite ends of the HCN molecule. In order to better
understand the helium results, we subsequently conducted
the gas phase infrared IR spectroscopy experiments pre-
sented in the current work. Most recently, Ishiguro et al.
measured the microwave spectra of these complexes also
presented in a companion paper
20
, which also indicate that
ortho-H
2
and para-H
2
bind at opposite ends of the HCN. In
the current IR study, only the HCNoH
2
complex was ob-
served. Ab initio calculations are presented that help to un-
derstand the preference for forming the ortho-H
2
complex.
The results agree with previous DFT calculations by Weso-
lowski and Weber,
21
that yielded a linear minimum with the
H
2
at the N-end of HCN. In addition, we nd a T-shaped
minimum on the potential energy surface, with the H
2
bind-
ing at the H-end of the HCN.
EXPERIMENTAL METHOD
Rotationally resolved infrared spectra have been ob-
tained for HCNH
2
and HCND
2
in the gas phase using the
optothermal detection method.
22,23
In both cases, the funda-
mental CH stretching vibration of the HCN near 3300
cm
1
was excited. The frequency region of interest was ac-
cessed using an F-center laser, operating on the RbCl:Li
crystal #3 and pumped by a krypton ion laser. A detailed
description of the tuning and calibration of this laser system
is given elsewhere.
23
In the present case, the absolute cali-
a
Author to whom correspondence should be addressed. Electronic mail:
remiller@unc.edu
b
Department of Chemistry, Faculty of Science, Kyushu University,
Fukuoka, 812-8581, Japan.
JOURNAL OF CHEMICAL PHYSICS VOLUME 115, NUMBER 11 15 SEPTEMBER 2001
5137 0021-9606/2001/115(11)/5137/7/$18.00 2001 American Institute of Physics
This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
129.72.2.27 On: Fri, 15 Nov 2013 05:20:16
bration of the spectrum was established using the R(0) tran-
sition of the HCN monomer, which was also observed in the
spectrum.
There are two basic congurations for the optothermal
method that are typically used to record infrared spectra.
With the bolometer positioned directly in the path of the
molecular beam, laser induced signals from weakly bound
complexes normally appear with the opposite phase to those
of stable monomers, since the complexes often dissociate
before reaching the detector. We have used this op out
mode of operation extensively to record spectra of many
complexes.
2426
For the purposes of studying the state-to-
state dissociation dynamics of other systems,
27,28
we have
also developed an apparatus that works in the op in
mode,
29,30
where the bolometer is placed off of the molecular
beam axis to detect the fragments that recoil from the mo-
lecular beam upon laser induced vibrational predissociation
of the complex. This method has the advantage that the bo-
lometer is not exposed to the direct molecular beam ux and
thus maintains its high sensitivity throughout the experiment.
In particular, the op-out method is problematic since hy-
drogen condenses on the bolometer operated at 1.6 K, re-
ducing its sensitivity during the course of the day. As a re-
sult, the op in mode was used in all of the experiments
reported here, with the bolometer positioned 4 from the mo-
lecular beam axis.
A spherical multipass cell
27,31
was used to optimize the
excitation efciency of the molecules in the beam. Since the
approximately 60 laser crossings are all at slightly different
angles with respect to the molecular beam, residual Doppler
broadening is observed in the spectra. Under these conditions
the instrumental linewidths were approximately 20 MHz,
substantially greater than the free running linewidth of the
laser approximately 2 MHz. Two electrodes placed on ei-
ther side of the photolysis volume can be used to apply a dc
electric eld to the complex. Large electric elds can be used
to orient the complexes in the laboratory frame using what is
sometimes referred to as the brute force method.
32,33
Un-
der these conditions the normal spectrum of a free rotor col-
lapses into a group of closely spaced pendular
transitions
34,35
near the vibrational origin. This electric eld-
induced Q branch is generally stronger than the zero eld
P- or R-branch transitions, facilitating the initial spectral
search. Once these transitions are found, the electric eld can
be switched off and a eld free spectrum recorded.
The complexes of interest were formed by expanding a
gas mixture composed of 0.5% of HCN and 25% hydrogen
or deuterium in helium through a 50 m diameter, room
temperature nozzle maintained at a total pressure of approxi-
mately 500 kPa. The expansion was sampled by a 400 m
diameter skimmer and the resulting molecular beam was fur-
ther collimated by a second, 2 mm diameter skimmer. As
noted above, the photofragments resulting from vibrational
predissociation of the complexes were detected by placing
the bolometer 4 from the molecular beam axis. Whenever
the laser was tuned through a transition associated with the
complex, photofragments were detected. The technique is
made background free by modulating the laser and using
phase sensitive detection of the bolometer signal.
FIG. 1. Schematic structures representing the two minima found by ab initio
calculations. Energies, frequencies, and structural parameters are summa-
rized in Tables IIV.
TABLE I. Ab initio energies for hydrogen-bonded structure 1 in Fig. 1.
Basis set
MP2 energies hartree
D
e
cm
1
HCN
a
H
2
a
H
2
HCN
6-311G(2d,2p) 93.222 578 1.162 829 94.385 974 124.4
6-311G(2df,2pd) 93.250 517 1.164 692 94.415 804 130.6
6-311G(3d f ,3pd) 93.254 614 1.164 992 94.420 268 145.3
a
Counterpoise corrected value see text.
TABLE II. Ab initio frequencies and selected structural data for structure 1.
Basis set
R
HB

a
R
COM

a
CH stretch
cm
1

Redshift
cm
1

b
B cm
1

c
6-311G2d,2p 2.497 4.126 3459.16 4.61 0.3863
6-311G2df,2pd 2.474 4.104 3473.44 6.39 0.3894
6-311G3df,3pd 2.440 4.071 3458.53 6.47 0.3941
a
As indicated in Fig. 1 for structure 1.
b
From free HCN optimized with the same basis set.
c
Represents average of calculated B and C rotational constants.
5138 J. Chem. Phys., Vol. 115, No. 11, 15 September 2001 Moore et al.
This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
129.72.2.27 On: Fri, 15 Nov 2013 05:20:16
AB INITIO CALCULATIONS
In parallel with the experimental work, ab initio calcu-
lations were performed using GAUSSIAN 98 Ref. 36 on an
IBM SP2 platform. Geometry optimizations were performed
at the MP2 level from various asymmetric starting structures,
with the H
2
placed near either end of the HCN. The follow-
ing basis sets were used: 6-311G2d,2p, 6-311
G2df,2pd, 6-311G3df,3pd. In all cases only the
two minima shown schematically in Fig. 1 where found. Fre-
quency calculations were performed for the nal geometries,
and no imaginary frequencies were found. The binding ener-
gies (D
e
) were computed by performing counterpoise cor-
rected partial optimizations for the monomer units in the
following fashion. Starting from the optimized complex
structure, the electrons and nuclei of the other monomer unit
were removed, while leaving the basis functions present.
Then a partial geometry optimization was performed leaving
the internal degrees of freedom of the remaining monomer
unit as variables. This treatment was used to correct for
basis-set superposition error.
37
All calculations with the H
2
near the H-end of the HCN
optimized to the T-shaped geometry indicated in Fig. 1a.
The MP2 energies for this structure and the corresponding
counterpoise corrected monomer energies are presented in
Table I. The frequency data and structural parameters are
presented in Table II. All calculations with the H
2
near the
N-end of the HCN optimized to the linear geometry shown
in Fig. 1b. The MP2 energies for this structure and the
corresponding counterpoise corrected monomer energies are
presented in Table III. The CH stretching frequencies and
structural parameters are presented in Table IV. The fre-
quency shifts due to complex formation were calculated for
the CH stretch using the frequencies not counterpoise cor-
rected from calculations on free HCN, optimized with the
same basis set as the complex. These results will be dis-
cussed in detail below, but for now we simply point out that
the linear geometry structure 2 is more strongly bound than
the T-shaped structure 1 by 25 cm
1
at all basis sets.
EXPERIMENTAL RESULTS
The search for the gas phase spectrum of HCNH
2
was
straightforward, given that the shift of the CH vibration is
expected to be small. It was made even easier by carrying out
the rst searches with a strong electric eld applied to the
interaction region, thus enhancing the transition intensity us-
ing the pendular state method.
34,35
Figure 2a shows a scan
of the region near the free CH stretching region of HCN,
in the presence of an 18.9 kV/cm electric eld. Only two
bands are seen in this spectrum; the lower frequency one
corresponds to the pendular spectrum of the free CH
stretch of the linear HCN dimer. The second band, appearing
at somewhat higher frequencies, only appeared when H
2
was
added to the gas mixture and is assigned to the HCNH
2
complex. Extensive searches were carried out for other
HCNH
2
isomers without success.
Having located the pendular state spectrum, a more de-
nitive assignment of this band to the HCNH
2
complex can
be made by recording the corresponding zero eld spectrum.
This spectrum is shown in Fig. 2b, with the closely spaced
transitions to the low frequency side of this spectrum arising
from the free stretch of the HCN dimer, discussed in detail
previously.
26
The transition marked with an asterisk is the
R(0) transition of the HCN monomer, located at
TABLE III. Ab initio energies for structure 2 in Fig. 1.
Basis set
MP2 energies hartree
D
e
cm
1
HCN
a
H
2
a
HCNH
2
6-311G2d,2p 93.222 686 1.162 841 94.386 231 154.5
6-311G2df,2pd 93.250 570 1.164 701 94.416 004 160.9
6-311G3df,3pd 93.254 601 1.164 965 94.420 355 173.2
a
Counterpoise corrected value see text.
TABLE IV. Ab initio frequencies and selected structural data for structure 2.
Basis set
R
HB

a
R
COM

a
CH stretch
cm
1

Redshift
cm
1

b
B cm
1

c
6-311G2d,2p 2.853 3.787 3462.59 1.41 0.4294
6-311G2df,2pd 2.848 3.781 3478.84 1.00 0.4304
6-311G3df,3pd 2.813 3.747 3464.36 0.63 0.4366
a
As indicated in Fig. 1 for structure 2.
b
From free HCN optimized with the same basis set.
c
Represents average of calculated B and C rotational constants.
FIG. 2. Experimental spectra of the CH stretch vibration in the
HCNorthoH
2
binary complex a with 18.9 kV/cm electric eld applied
and b zero eld condition. The two transitions in a are electric eld
induced pendular state transitions near the vibration origins of the
1
free-CH stretch HCN dimer and the HCNorthoH
2
complex. The eld
free spectrum of the HCNorthoH
2
in b consists of regular P and R
branches, in overlap with the R branch of the
1
HCN dimer at lower
frequencies.
5139
J. Chem. Phys., Vol. 115, No. 11, 15 September 2001 Infrared spectroscopy of HCNH
2
/D
2
complexes
This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
129.72.2.27 On: Fri, 15 Nov 2013 05:20:16
3314.412 43 cm
1
.
38
The remaining transitions are assigned
to the HCNH
2
complex, as shown in the gure. Table V
contains a list of the transition frequencies, along with the
corresponding residuals obtained from tting the spectrum to
a simple linear rotor energy level expression. The t to the
data is clearly excellent and the resulting molecular constants
are summarized in Table VI.
It is interesting to note that, as illustrated in the previous
studies of H
2
HCl Ref. 8 and HFH
2
,
4
a T-shaped com-
plex involving ortho-H
2
would give rise to a band
with a strong Q-branch. Based on the fact that we only ob-
serve P and R branch transition in the present spectrum, we
presume that the band is , excluding the T-shaped iso-
mer as the source of this spectrum. It is therefore tempting,
in analogy with the H
2
HCl results,
8
to attribute this spec-
trum to the para-H
2
HCN hydrogen-bonded complex, which
should produce a band. However, the small redshift
observed for this band 0.34 cm
1
from HCN monomer at
3311.48 cm
1
Ref. 38 indicates that the CH stretch is only
weakly affected by the hydrogen, suggesting that the H
2
is
bound at the nitrogen end of the HCN monomer. Further-
more, since we observe only one hydrogen related band in
the spectrum, it is reasonable to expect that it is associated
with the more abundant spin-species ortho-H
2
, particularly
since it interacts more strongly with the HCN than para-H
2
,
in analogy with the HX systems discussed above. These con-
siderations, along with the computational results presented
above and discussed below, lead us to conclude that the ex-
perimental spectrum corresponds to the linear HCN
ortho-H
2
complex shown in Fig. 1b.
Further support for this assignment comes from the ro-
tational constants determined from the t. First we note that
the rotational constant decreases slightly 0.3% upon vibra-
tional excitation. A somewhat larger percentage decrease is
observed in the case of the HCN monomer
38
0.7%, which
can be explained as being due to the lengthening of the CH
bond upon vibrational excitation. In contrast, vibrational ex-
citation of a hydrogen-bonded HX vibration generally leads
to an increase in the rotational constant due to the corre-
sponding increase in the strength and hence contraction of
the hydrogen bond. Indeed, the rotational constant for
H
2
HF, where the H
2
binds to the H atom of HF, increases
by 2.2% in the excited vibrational state.
4
Thus the small
decrease in the rotational constant again suggests that the H
2
is remote from the hydrogen on HCN.
It is interesting to note that the centrifugal distortion
constant is quite large for this complex, indicating that the
hydrogen molecule is quite weakly bound to the HCN mono-
mer. In fact, the van der Waals stretching frequency esti-
mated in the pseudodiatomic approximation, using the
relation
39

4B
3
D
is only 27.6 cm
1
. If we now assume that the assignment of
the spectrum to a linear complex is correct, we can use the
following equation for a rodball system to determine the
intermolecular distance:
40
I
complex
M
c
R
i
2

cos
2
1
2
I
HCN
,
M
c

m
HCN
m
H
2
m
HCN
m
H
2
,
where I
HCN
and I
complex
are the moments of inertia of free
HCN and the complex, respectively, R
i
is the distance be-
tween the centers of mass of the HCN and H
2
, M
c
is the
reduced mass of the complex as dened above, and cos
2

is the expectation value of the square of the cosine of the


HCN bending angle, determined by Ishiguro et al. to be
32.65.
20
The distance determined in this fashion using the
B constant given in Table VI is 3.961 .
Isotopic substitution is a standard approach for locating
the positions of certain atoms in a molecule and the above
structural assignment can be conrmed by studying the spec-
trum of the corresponding HCND
2
complex. Here again,
only a single band was observed, as shown in Fig. 3.
The corresponding R and P branch transitions are indicated
in the gure. The transition frequencies are listed in Table
VII and the corresponding molecular constants are summa-
rized in Table VI. This band is assigned to the HCN pD
2
complex by analogous arguments to those used for the ortho-
H
2
complex.
Comparing the resulting constants with those for the
HCNoH
2
complex reveals some interesting points. The in-
termolecular bond length calculated as above is 3.911 ,
signicantly shorter than for the oH
2
complex. Also, the vi-
brational redshift of HCN pD
2
is somewhat larger than for
TABLE V. A summary of observed transitions, frequencies, and residuals
from t Table VI to a linear rotor spectrum for the HCNortho-H
2
com-
plex. All frequencies are in cm
1
.
Transition Observed Obs.Calc.
P(3) 3308.5958 0.000 49
P(2) 3309.4286 0.000 16
P(1) 3310.2801 0.000 16
R(0) 3311.9972 0.000 01
R(1) 3312.8442 0.000 15
R(2) 3313.6742 0.000 15
R(3) 3314.4839 0.000 04
TABLE VI. A summary of molecular constants for HCNortho-H
2
and
para-D
2

a
. All constants are in cm
1
.
HCN(o-H
2
)
MW data for
HCN(o-H
2
)
b
HCN( p-D
2
)
B 0.430 6524 0.430 289 0.266 2724
B 0.429 3216 0.266 0014
D10
4
4.2025 4.085 0.9217
D10
4
3.3810 0.877

0
3311.13995 3311.12287
Redshift
c
0.3369 0.3541
a
Uncertainties in parentheses correspond to one standard deviation 1
from the least squares t in the last digits.
b
Reference 20.
c
From HCN monomer at 3311.476 83 cm
1
Ref. 38.
5140 J. Chem. Phys., Vol. 115, No. 11, 15 September 2001 Moore et al.
This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
129.72.2.27 On: Fri, 15 Nov 2013 05:20:16
the H
2
complex. Both of these effects are consistent with a
lower zero-point energy for the pD
2
complex. Furthermore,
the centrifugal distortion constant is a factor of 5 smaller
than for the H
2
complex, indicative of the stronger binding
for the pD
2
complex. In fact, the intermolecular stretching
frequency for HCN pD
2
estimated as above is 28.6 cm
1
,
which is larger than for HCNoH
2
; note that in the har-
monic oscillator limit this stretching frequency would be
smaller by a factor of (2)
1/2
for HCN pD
2
. Clearly, the pD
2
and oH
2
complexes sample different regions of the potential
energy surface. In light of this, we made several attempts to
construct effective one-dimensional 1D potentials that
could simultaneously t the B and D constants for both the
oH
2
and pD
2
complexes. Our lack of success in these en-
deavors led us to conclude that a multidimensional model is
needed to quantitatively describe the vibrational dynamics of
these systems.
It is now clear that the observed spectra arise from linear
complexes of HCN and ortho-H
2
or para-D
2
, bound at the
N-end of the HCN. The question that remains is why the
para-H
2
complex was not observed. Although para-H
2
is
present in the gas mixture at 1/3 the abundance of ortho-H
2
,
we do not believe this difference is sufcient to account for
the absence of this complex in the spectrum, particularly
since the missing species in the HCND
2
case is the more
abundant one 2:1 ortho:para. Recall that the spin designa-
tions for the boson species D
2
J oddpara, J evenortho
are opposite to those for H
2
. One possibility is that the para-
H
2
complex undergoes rapid vibrational predissociation, and
the transitions are simply too broad to observe. This is cer-
tainly a reasonable explanation, particularly if the para-H
2
is
bonded to the hydrogen end of the HCN and thus more di-
rectly coupled to the CH stretching vibration. Other pos-
sible explanations for why the para-H
2
complex is missing
from the spectrum will be explored in the next section.
All of the transitions observed for HCN pD
2
and
HCNoH
2
were found to have the same Gaussian linewidth
as the HCN monomer, namely 20 MHz. This can be attrib-
uted to Doppler broadening associated with the spherical
multipass cell. Since no additional broadening is observed
due to vibrational predissociation of the complex in the ex-
cited vibrational state, the lower limit on the predissociation
lifetime is estimated to be 160 ns. This long lifetime is con-
sistent with the fact that the H
2
is remote from the CH
stretching coordinate, making vibrational predissociation
slow.
41
COMPARISON OF EXPERIMENT AND AB INITIO
CALCULATIONS
The experimental results presented above can be justi-
ed in terms of the ab initio calculations, given that for all
basis sets the well depths (D
e
) are larger for the linear
structure compared to the T-shaped. Indeed, for the largest
basis set the well depths are 173.2 cm
1
and 145.3 cm
1
,
respectively. Therefore, it is perhaps not surprising that we
observe the formation of the nitrogen bound complex instead
of the hydrogen bonded one. However, in view of the large
zero-point energies associated with these complexes and the
signicant difference between the two nuclear spin species of
hydrogen, this simple argument cannot explain everything
we observe. The most important effect is that, as is often the
case in complexes with H
2
,
42
the anisotropy of the intermo-
lecular potential is not strong enough to signicantly mix the
rotational states of the hydrogen. Thus, for HCNH
2
, the
para and ortho complexes have the hydrogen in essentially
pure j 0 and j 1 states, respectively. To a good approxi-
mation, therefore, the effective intermolecular potentials for
para and ortho hydrogen can be obtained by averaging over
the appropriate free rotor wave function for the hydrogen;
calculations of this type are presented in the accompanying
paper.
3
The implications of this averaging are that the j 0
hydrogen is completely delocalized in angle at either the
hydrogen or nitrogen end of the HCN molecule, behaving
essentially like a rare-gas atom. For the j 1 species, the
presence of the HCN splits the degeneracy of the free-rotor
M-states, allowing different projections k1 and k0
of the H
2
angular momentum on the HCN molecular axis.
The observation of a band for the HCNoH
2
complex
indicates that it is in the k0 case the k1 case gives
rise to bands
6,8
, so the appropriate free-rotor wave
function for the averaging is cos(
H
2
). This wave function
has its highest density in the linear structure predicted to
be the minimum at the N-end of the HCN, and a node where
FIG. 3. Experimental spectrum of the CH stretching vibration in the
HCNparaD
2
binary complex. The spectrum of the HCNparaD
2
com-
plex consists of regular P and R branches, in overlap with the R branch of
the
1
free stretch HCN dimer at lower frequencies.
TABLE VII. A summary of observed transitions, frequencies and residuals
from linear rotor t Table VI to the spectrum for the HCNpara-D
2
. All
frequencies are in cm
1
.
Transition Observed Obs.Calc.
P(3) 3309.5336 0.000 02
P(2) 3310.0601 0.000 05
P(1) 3310.5906 0.000 05
R(0) 3311.6546 0.000 20
R(1) 3312.1832 0.000 20
R(2) 3312.7080 0.000 14
R(3) 3313.2256 0.000 11
5141
J. Chem. Phys., Vol. 115, No. 11, 15 September 2001 Infrared spectroscopy of HCNH
2
/D
2
complexes
This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
129.72.2.27 On: Fri, 15 Nov 2013 05:20:16
the H
2
axis is perpendicular to the HCN. Thus the rotational
averaging will result in a stronger interaction than for the j
0 species.
The absence of the para-H
2
complexes in our infrared
experiment can now be understood by taking into account
the mechanism of complex formation in the free jet expan-
sion. Close to the nozzle, in the warm part of the expansion,
complexes are formed by three-body collisions and then
cooled by two-body collisions. However, these two-body
collisions also tend to replace less stable complexes by more
strongly bound ones. For example, if a para-hydrogen com-
plex weakly bound collides with an ortho-H
2
molecule, it is
energetically favorable to form an ortho complex more
strongly bound and eject the para-hydrogen. Since there are
many two-body collisions downstream of where the com-
plexes are formed, this exchange process is expected to occur
with high efciency at the relative concentrations of ortho-
and para-hydrogen and -deuterium considered here. It is in-
teresting to note that in their recent microwave study, Ishig-
uro et al. did observe the hydrogen-bonded pH
2
HCN com-
plex, although the intensity was a factor of 5 lower than for
HCNoH
2
, instead of the factor of 3 that is expected from
the normal abundances.
20
Since the exchange process is de-
pendent on experimental factors such as gas composition,
stagnation pressure, and nozzle geometry, the ratio could be
even larger in our case, making the para-H
2
signals simply
too small to detect.
Let us now consider the calculated vibrational frequency
data. Although the magnitudes of the vibrational frequencies
are consistently overestimated by ab initio harmonic calcu-
lations, relative frequency shifts are often more faithfully
reproduced. For the largest basis set, the calculated fre-
quency shift is approximately an order of magnitude larger
for the T-shaped, H-bonded complex than for the linear one,
as expected. Also, the agreement between the calculations
and the experiment improves with the size of the basis set.
The best calculations give a shift of 0.63 cm
1
, compared
with the experimental values of 0.337 cm
1
and 0.354 cm
1
for HCNorthoH
2
and HCNparaD
2
, respectively. Com-
paring rotational constants and bond lengths is more prob-
lematic, given that the experimental values represent vibra-
tional averages, while the calculations are for the minimum
energy structures. Nevertheless, the agreement is reasonable,
the best basis set giving B0.4366 cm
1
for
HCNorthoH
2
, compared with the experimental ground
state value of B0.430 71 cm
1
. A somewhat larger value is
indeed expected from the calculation, since vibrational aver-
aging of an anharmonic potential will lengthen the van der
Waals bond, making the experimental value smaller.
It is interesting to examine the physical reasons for the
linear N-end bonding of the HCNH
2
complex, in contrast
with the T-shaped, hydrogen-bonded H
2
HF complex. As
pointed out in the introduction, the T-shaped structure of the
ortho-H
2
HF complex is favorable for both the dipole
quadrupole (dq) and quadrupolequadrupole (qq) inter-
actions. Since these electrostatic interactions are also domi-
nant in HCNH
2
, a linear geometry seems surprising,
especially since it involves a repulsive interaction between
the quadrupoles. The reason can be seen qualitatively from
the relative magnitudes of the electrostatic moments of HCN
and HF. The calculated MP2/6-311G3df,3pd dipole
moments atomic units are 0.718 and 1.190 for HF and
HCN, respectively, while the quadrupole moments are 1.728
and 1.669, respectively. Since the dipolequadrupole term is
relatively larger for HCNoH
2
, the larger angular factor in
the linear geometry
43
compensates for the repulsive
quadrupolequadrupole term. Using the calculated moments,
the electrostatic contribution to the intermolecular interaction
can be evaluated for the two minimum structures using the
standard equations for the dq and qq interactions.
43
At
the distances corresponding to the minima, these equations
give 71.1 cm
1
for structure 1 4.07 and 84.8 cm
1
for structure 2 3.74 , respectively. Thus the electrostatic
interaction accounts for approximately half of the total cal-
culated well depths, and favors the observed experimental
geometry. The remainder of the stabilization in the two wells
results from competition between repulsive interactions and
induction/dispersion forces, the latter being always attrac-
tive. For HCN, the dispersion interaction, which depends on
charge density, should be stronger for the nitrogen lone pair
than for the strongly polarized H atom, further favoring the
N-end geometry.
SUMMARY
We have presented the rst infrared gas phase spectra of
the complexes formed between H
2
/D
2
and HCN. The obser-
vation of bands in both cases, which are only slightly
redshifted from the free CH stretch of HCN, leads to the
conclusion that the most stable complexes have linear
HCN(J1)H
2
/D
2
structures. This is in agreement with the
ab initio calculations reported here, which in addition reveal
a somewhat shallower well at the hydrogen end, correspond-
ing to a T-shaped complex. In spite of extensive searches, we
nd no evidence for the formation of complexes with the
H
2
/D
2
at the hydrogen end of the HCN. The absence of para-
H
2
complexes is most likely due to thermodynamically fa-
vored exchanges with more strongly bound ortho-H
2
com-
plexes in the free jet.
In part, our motivation for studying this system in the
gas phase is related to an accompanying paper, which deals
with the formation of hydrogenHCN complexes in liquid
helium nano-droplets.
3
The interest in hydrogen-containing
complexes in helium arises because they are sufciently
weakly bound to consider the possibility that the helium sol-
vent modies the structure of the complex. Having accurate
molecular constants for the gas phase complex is clearly im-
portant for determining such effects. The results obtained
here will be compared with the helium droplet experiments
in the following paper.
ACKNOWLEDGMENTS
This work was supported by the National Science Foun-
dation CHE-99-87740. The authors also acknowledge the
donors of the Petroleum Research Fund, administered by the
American Chemical Society, for partial support of this re-
search. One of the authors M.I. thanks the nancial support
5142 J. Chem. Phys., Vol. 115, No. 11, 15 September 2001 Moore et al.
This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
129.72.2.27 On: Fri, 15 Nov 2013 05:20:16
of The Japan Society for the Promotion of Science. The com-
puter time was provided by the North Carolina Supercom-
puting Center.
1
D. J. Nesbitt, Chem. Rev. 88, 843 1988.
2
J. M. Hutson, Annu. Rev. Phys. Chem. 41, 123 1990.
3
D. T. Moore, M. Ishiguro, and R. E. Miller, J. Chem. Phys. 115, 5144
2001, following paper.
4
C. M. Lovejoy, D. D. Nelson, Jr., and D. J. Nesbitt, J. Chem. Phys. 87,
5621 1987.
5
K. W. Jucks and R. E. Miller, J. Chem. Phys. 87, 5629 1987.
6
C. M. Lovejoy, D. D. Nelson, Jr., and D. J. Nesbitt, J. Chem. Phys. 89,
7180 1988.
7
E. J. Bohac and R. E. Miller, J. Chem. Phys. 98, 2604 1993.
8
D. T. Anderson, M. D. Schuder, and D. J. Nesbitt, Chem. Phys. 239, 253
1998.
9
D. C. Clary and P. J. Knowles, J. Chem. Phys. 93, 6334 1990.
10
D. C. Clary, J. Chem. Phys. 96, 90 1992.
11
P. J. Krause and D. C. Clary, Mol. Phys. 93, 619 1998.
12
D. H. Zhang, J. Z. H. Zhang, and Z. Bacic, J. Chem. Phys. 97, 927 1992.
13
D. H. Zhang, J. Z. H. Zhang, and Z. Bacic, Chem. Phys. Lett. 194, 313
1992.
14
D. H. Zhang, J. Z. H. Zhang, and Z. Bacic, J. Chem. Phys. 97, 3149
1992.
15
A. R. W. McKellar, Chem. Phys. Lett. 186, 58 1991.
16
A. R. W. McKeller, J. Chem. Phys. 108, 1811 1998.
17
I. Pak, L. A. Surin, B. S. Dumesh, D. A. Roth, F. Lewen, and G. Win-
newisser, Chem. Phys. Lett. 304, 145 1999.
18
A. R. W. McKellar, J. Chem. Phys. 112, 9282 2000.
19
M. J. Weida and D. J. Nesbitt, J. Chem. Phys. 110, 156 1999.
20
M. Ishiguro, T. Tanaka, K. Harada, C. J. Whitham, and K. Tanaka, J.
Chem. Phys. 115, 5155 2001, this issue.
21
T. A. Wesolowski and J. Weber, Int. J. Quantum Chem. 61, 303 1997.
22
R. E. Miller, Acc. Chem. Res. 23, 10 1990.
23
Z. S. Huang, K. W. Jucks, and R. E. Miller, J. Chem. Phys. 85, 3338
1986.
24
P. A. Block, K. W. Jucks, L. G. Pedersen, and R. E. Miller, Chem. Phys.
139, 15 1989.
25
K. W. Jucks and R. E. Miller, J. Chem. Phys. 88, 2196 1988.
26
K. W. Jucks and R. E. Miller, J. Chem. Phys. 88, 6059 1988.
27
R. J. Bemish, E. J. Bohac, M. Wu, and R. E. Miller, J. Chem. Phys. 101,
9457 1994.
28
L. Oudejans and R. E. Miller, J. Phys. Chem. 99, 13670 1995.
29
R. J. Bemish, P. A. Block, L. G. Pedersen, W. T. Yang, and R. E. Miller, J.
Chem. Phys. 99, 8585 1993.
30
P. A. Block, L. G. Pedersen, and R. E. Miller, J. Chem. Phys. 98, 3754
1993.
31
E. J. Bohac, M. D. Marshall, and R. E. Miller, J. Chem. Phys. 96, 6681
1992.
32
H. J. Loesch and A. Remscheid, J. Chem. Phys. 93, 4779 1990.
33
B. Friedrich and D. R. Herschbach, Nature London 353, 412 1991.
34
J. M. Rost, J. C. Grifn, B. Friedrich, and D. R. Herschbach, Phys. Rev.
Lett. 68, 1299 1992.
35
P. A. Block, E. J. Bohac, and R. E. Miller, Phys. Rev. Lett. 68, 1303
1992.
36
M. J. Frisch, J. S. Binkley, and H. B. Schlegel, GAUSSIAN 98, Gaussian Inc.,
Pittsburgh, PA, 1998.
37
S. F. Boys and F. Bernardi, Mol. Phys. 19, 553 1970.
38
J.-I. Choe, T. Tipton, and S. G. Kukolich, J. Mol. Spectrosc. 117, 292
1986.
39
C. H. Townes and A. L. Schawlow, Microwave Spectroscopy Dover, New
York, 1975.
40
K. R. Leopold, G. T. Fraser, F. J. Lin, D. D. Nelson, Jr., and W. Klemperer,
J. Chem. Phys. 81, 4922 1984.
41
R. E. Miller, Science 240, 447 1988.
42
A. R. W. McKellar, Faraday Discuss. Chem. Soc. 73, 89 1982.
43
G. C. Maitland, M. Rigby, E. B. Smith, and W. A. Wakeham, Intermolecu-
lar Forces Clarendon, Oxford, 1981.
5143
J. Chem. Phys., Vol. 115, No. 11, 15 September 2001 Infrared spectroscopy of HCNH
2
/D
2
complexes
This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
129.72.2.27 On: Fri, 15 Nov 2013 05:20:16

Das könnte Ihnen auch gefallen