Molecular Physics: An International Journal at The Interface Between Chemistry and Physics

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

This article was downloaded by: [RMIT University]

On: 17 August 2015, At: 05:48


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: 5 Howick Place,
London, SW1P 1WG

Molecular Physics: An International Journal at the


Interface Between Chemistry and Physics
Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/tmph20

Accurate computations of the rovibrational spectrum


of the He–HF van der Waals complex
a a b b
Jose Luis Cagide Fajin , Berta Fernandez , Aleksandra Mikosz & David Farrelly
a
Faculty of Chemistry, Department of Physical Chemistry , University of Santiago de
Compostela , E-15782 Santiago de Compostela, Spain
b
Department of Chemistry and Biochemistry , Utah State University , Logan, Utah 84322,
USA
Published online: 21 Aug 2006.

To cite this article: Jose Luis Cagide Fajin , Berta Fernandez , Aleksandra Mikosz & David Farrelly (2006) Accurate
computations of the rovibrational spectrum of the He–HF van der Waals complex, Molecular Physics: An International Journal
at the Interface Between Chemistry and Physics, 104:09, 1413-1420, DOI: 10.1080/00268970500480984

To link to this article: http://dx.doi.org/10.1080/00268970500480984

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Molecular Physics, Vol. 104, No. 9, 10 May 2006, 1413–1420

Accurate computations of the rovibrational spectrum of the


He–HF van der Waals complex
JOSE LUIS CAGIDE FAJINy, BERTA FERNANDEZ*y, ALEKSANDRA MIKOSZz and
DAVID FARRELLYz
yFaculty of Chemistry, Department of Physical Chemistry, University of Santiago de Compostela,
E-15782 Santiago de Compostela, Spain
zDepartment of Chemistry and Biochemistry, Utah State University, Logan, Utah 84322, USA

(Received 28 September 2005; in final form 18 November 2005)

The rovibrational spectrum of the He–HF van der Waals complex is calculated from an
accurate intermolecular potential energy surface. This is obtained by fitting a considerable
number of interaction energies evaluated at the Coupled Cluster Singles and Doubles level
including connected triple corrections and with an augmented correlation consistent polarized
Downloaded by [RMIT University] at 05:48 17 August 2015

valence quintuple zeta basis set extended with a set of 3s3p2d1f1g mid-bond functions. The
basis set was selected after a systematic study carried out at four intermolecular geometries.
The potential is characterized by two linear minima, i.e., He–HF and He–FH, with distances
from the He atom to the HF centre of mass of 3.1662 Å and 2.9989 Å and binding energies
of 43.844 and 26.169 cm1, respectively. These results are compared to data available
in the literature. An analytic fit to this potential energy surface is presented and used to
compute several low-lying rovibrational energy states using coupled channel methods. These
results are compared with fixed-frame diffusion Monte-Carlo calculations using a method
developed specifically for linear molecules.

Keywords: Rovibrational spectrum; CCSD(T) intermolecular potential energy surface; He–HF van der
Waals complex

1. Introduction In recent years a battery of experimental and


theoretical methods has been directed towards van der
The advent of helium nanodroplet isolation spectro- Waals complexes involving He and other rare-gas atoms
scopy (HENDI) and its variants [1, 2] has fuelled interest [1–4]. Prominent examples are rare gas complexes with
in van der Waals complexes consisting of clusters of hydrogen halides [5]. It is, therefore, natural to consider,
4
He atoms seeded either with single molecules or small specifically, the rotational structure and dynamics of
molecular complexes, e.g., ammonia-dimer [3]. It has complexes consisting of 4He atoms and hydrogen halide
become clear that the rotational and solvation dynamics molecules (designated HX where X ¼ F, Cl, Br, I).
of such systems are profoundly affected by whether Besides making contact with experimental studies, a
or not the droplet is a superfluid. In particular, the longer range goal of this work is to uncover and
properties of seeded 4He clusters contrast strongly with understand systematic differences in the rotational
the more classical solvation dynamics observed in dynamics of 4He droplets doped with the homologous
fermionic 3He clusters in that rotationally resolved sequence of molecules HF, HCl, HBr, HI and their
spectra, albeit with altered molecular constants, are dimers, trimers etc. The two extreme limits of this
routinely obtained. Thus, superfluid 4He droplets sequence represent cases where it is expected that the He
constitute ‘‘quantum solvents’’ which serve as excellent density will in the one case (HF) adiabatically decouple
‘‘gentle matrices’’ in which to perform spectroscopic from the molecular rotation or, in the other (HI),
studies [2]. adiabatically follow the rotor [2–4]. The availability of
accurate potential energy surfaces (PESs) for He–HX
*Corresponding author. Email: qfberta@usc.es complexes is clearly critical to understanding how

Molecular Physics
ISSN 0026–8976 print/ISSN 1362–3028 online ß 2006 Taylor & Francis
http://www.tandf.co.uk/journals
DOI: 10.1080/00268970500480984
1414 J. L. C. Fajin et al.

(and if ) such predicted transitions actually take place. In 1990 Lovejoy et al. [14] recorded the near-infrared
In addition, the effect of the He environment on rovibrational spectra of the He–HF and the HeDF
hydrogen bonding and H-atom tunneling dynamics complexes in a slit supersonic expansion. To analyze
within (HX)n, n ¼ 2, 3, . . . , complexes is a topic of the data they use a Lennard-Jones(m,6) potential. The
considerable interest [6, 7] which also necessitates the authors also checked the effect of the potential
availability of accurate PESs. anisotropy by using the surfaces of Rodwell et al. [10]
As an initial step towards such a systematic study, in but they got discrepancies with respect to the experi-
this paper an accurate PES is developed for the He–HF mental results. Even after modifying these potentials
van der Waals complex. This surface is then used to in order to get better agreement certain discrepancies
compute several low-lying rotational energy levels using still remained. An equilibrium dissociation energy of
a variety of accurate quantum approaches including (25.03  0.02) cm1 was obtained for the ground state
a variant of the rigid-body diffusion Monte Carlo and an equilibrium distance Re of (3.113  0.001) Å. The
(DMC) method [8]. DMC approaches are among the ground state dissociation energy D0 was estimated
few methods which are currently available for studying as (7.14  0.10) cm1.
droplets which consist of a large number of He Smith and Rabitz [15] studied the effect on the
atoms doped with a single, or a few, molecules. Here inelastic cross sections, rate constants and rotational
a fixed frame diffusion Monte Carlo (FFDMC) [4] energy level populations of infinitesimal functional
method is developed which is specifically applicable to variations in the rigid rotor He–HF surface. The surface
Downloaded by [RMIT University] at 05:48 17 August 2015

linear rotors such as HF. used was the analytic representation of the rigid rotor
He–HX complexes are characterized by two linear Rodwell et al.’s [10] potential proposed by Tennyson
minima whose relative stability depends on the partic- et al. [11]. They found that these observables were
ular complex [6, 8, 9]. In particular, for the He–HF sensitive to the variation of many Legendre components
van der Waals complex Rodwell et al. [10] used a hybrid of the potential.
In 1993 Slee et al. [16] presented the Iterative Secular
Hartree Fock SCF plus Damped Dispersion (HFD)
Equation (ISE) method which is based on the secular
method and a gaussian basis set to obtain the potential
equation-perturbation theory approach of Hutson and
energy surface for He–HF complex. They considered the
Le Roy [17]. This method was used together with
variation of the H–F distance by using distances of
Hutson’s potential surface [9c] to obtain the rovibra-
1.4827, 1.6078, 1.7328, 1.9180 and 2.1032 a0. The inter-
tional energy levels for He–HF, among other complexes.
action energies were fitted to a series of Legendre
These energy levels were close to those obtained
polynomials, and the potential surface was found to
by Lovejoy et al. [14]
have an absolute minimum with an equilibrium distance
Moszynski et al. [18, 19] used symmetry-adapted
between the He atom and the HF centre of mass perturbation theory (SAPT) and medium size basis
(denoted Re in the following) of 3.159 Å and an energy sets to compute the intermolecular potential energy
of 33.119 cm1. surfaces and the rovibrational spectra of the HeHF and
Later Tennyson et al. [11] used the linear combination HeDF complexes. The calculations were performed
of radial and angular momentum function products for three different HF distances, 1.6078, 1.7328 and
(LC-RAMP) method that derives from the close- 1.9180 a0. The global minimum is in He–HF linear
coupling approach in body-fixed coordinates, to obtain configuration, with Re ¼ 6.16 a0 and De ¼ 39.68 cm1.
the rovibrational levels of the He–HF complex. These A local minimum is located in the He–FH configuration,
levels were calculated with the potential of Rodwell et al. characterized by a He FH distance of 5.59 a0 and an
[10] The correlation between the HF vibrational coor- energy of 36.13 cm1. This potential energy surface
dinate and the complex modes was estimated to be less predicts a dissociation energy of 7.38 (7.50) cm1 for
than 0.02 cm1. He–HF (He–DF) and compares very well with the
Boughton et al. [12] measured the total differential empirical potential of Lovejoy and Nesbitt [14] although
scattering cross-section of the He–HF complex at two around the T-shaped configuration the two surfaces are
collision energies. They used the Coupled States (CS), slightly shifted and the agreement is worse. Later
IOS and single channel spherical potential methods Moszynski et al.’s potential was slightly modified [20]
to predict the cross section from different potentials. and used to calculate the integral and differential cross
These potentials are based on modifications of those sections for elastic and inelastic He–HF scattering.
proposed by Rodwell et al. [10] and of Barker et al. [13] The results were compared with previous experimental
They concluded that the potential minimum is char- data given accurate scattering cross sections.
acterized by Re ¼ (3.11  0.03) Å and an energy of In 2003, Stoecklin et al. [21] using the Brueckner
(24.0  1.7) cm1. Coupled Cluster Doubles method with a perturbative
Accurate computations of the rovibrational spectrum 1415

estimation of the triple contribution (BCCD(T)) method this vector makes with the HF bond axis, taking as  ¼ 0
and the aug-cc-pVQZ basis set extended with bond the linear He–HF configuration. The geometry of the
functions calculated the potential energy surface for the HF molecule is kept fixed at the r0 value of 0.925597 Å
He–HF complex. The calculations were carried out for obtained from the microwave spectra [24].
459 geometries with three different H–F distances, 1.50, Interaction energies were calculated using the super-
1.7328 and 2.12 a0. After fitting the potential, the surface molecular model and corrected for basis bet superposi-
presented two minima. The global minimum is located tion error (BSSE) by invoking the counterpoise method
in the linear He–HF conformation, with Re ¼ 5.95 a0 of Boys and Bernardi [25]. The BSSE correction is
and it has an energy of 43.70 cm1. The local minimum known to be essential in order to get accurate interaction
is in the He–FH geometry with the He atom at 5.65 a0 energies for these systems. By using the CCSD(T)
from the HF centre of mass and an energy of method and Dunning’s augmented correlation consis-
25.88 cm1. The barrier between the two minima tent polarized valence basis sets extended with an
is given by an almost T-shaped complex geometry with additional set of 3s3p2d1f1g mid-bond functions we
R ¼ 6.10 a0 and an energy of 16.7532 cm1. The obtained very good results in previous studies on similar
surface was used to determine the quenching cross complexes (see Ref. [23b] and Ref. cited therein).
section and rate coefficient of HF molecules in collisions Therefore, we use this method and these bases to carry
with He atoms by performing close coupling calcula- out the interaction energy calculations. The mid-bond
tions. In a subsequent work, Reese et al. [22] used this functions [26] are well known to very efficiently improve
Downloaded by [RMIT University] at 05:48 17 August 2015

surface to study the rotational transitions induced basis set convergence, they are located in the middle of
in these collisions. the van der Waals bond, and denoted 33211. The
In previous work, highly accurate intermolecular exponents of these functions are: 0.90, 0.30, 0.10 for the
potential energy surfaces have been computed using s and the p functions; 0.60, 0.20 for the d functions, and
the coupled cluster singles and doubles model including 0.30 for the g and f functions.
connected triple corrections (CCSD(T)) and augmented In the study of NeN2 van der Waals complex [27], we
correlation consistent basis sets extended with a set of showed that to obtain accurate van der Waals complex
3s3p2d1f1g mid-bond functions (denoted 33211) [23]. rovibrational spectra it is essential to carefully select
In the present study we evaluate the He–HF van der an adequate basis set. Considering this, we start the
Waals complex ground state PES and the corresponding calculations with a systematic basis set convergence
rovibrational levels. Considering the good results previ- study using the xaug-cc-pVXZ (x ¼ , d; X ¼ D, T, Q, 5)
ously provided by the CCSD(T) method [23], we use set of bases, with and without the 33211 set of mid-bond
it here. In order to select an adequate basis set, and functions. We carry out the calculations at four geome-
knowing the importance of properly checking basis set tries: two linear selected close to the PES minima
convergence (see for example the analysis in Ref. [23b]), available in the literature (He–HF: R ¼ 3.26 Å,  ¼ 0 ;
we decide to start by carrying out a systematic basis set He–FH: R ¼ 2.9581 Å,  ¼ 180 ) [18], a T-shape geom-
study within the series of bases xaug-cc-pVXZ-33211 etry (R ¼ 3.6609 Å,  ¼ 90 ) and a geometry in the
(x ¼ ,d; X ¼ D, T, Q, 5, 6). positive energy region (R ¼ 2.79 Å,  ¼ 0 ). The results of
The paper is organized as follows: In section 2 this study are presented in Table 1.
we describe the evaluation of the PES and comment Taking into account the size of the interaction energy,
on the results. In section 3, we compute the rovibra- we consider an error limit of 0.5 cm1 in the analysis. As
tional spectrum using standard methods. The fixed- expected, we can see that the presence of the mid-bond
frame diffusion Monte-Carlo method is described in functions accelerates basis set convergence, and even at
section 4 and ground state energies are computed. The the quintuple zeta level these functions are relevant
excited states are obtained using the fixed-node in the basis set. We have to go up to the augmented
approach. In the last section, we provide a summary quintuple zeta level in order to get the convergence-limit
and give our concluding remarks. stability. The double augmentation of this basis set is
not needed, but in the calculations performed without
mid-bond functions it is. It is worthwhile to point
2. Intermolecular potential energy surface out the large improvement in the results at the double
zeta level when using mid-bond functions.
We select a set of 110 intermolecular geometries to carry The He–HF configuration is significantly more
out the ab initio interaction energy calculations. The R, stable than the He–FH, and all bases agree on this
 coordinates are use to describe these geometries; R result, although differing considerably in the relative
is the length of the He position vector whose origin energy value. Considering the above, to carry out the
is located at the HF center of mass and  the angle calculations we select the aug-cc-pV5Z-33211 basis set.
1416 J. L. C. Fajin et al.

Table 1. Basis set study. Energies are given in cm1. See text We use the DALTON [28] and ACES II [29]
for the definition of the different geometries. programs. The interaction energy results are displayed
BASIS HeHF HeFH T-Shape Positive in table 2.
The complex’s PES is constructed by fitting the
aug-cc-pVDZ-33211 41.399 26.431 12.869 15.434
aug-cc-pVTZ-33211 42.182 25.784 12.156 14.250
ab initio single point interaction energies to the ana-
aug-cc-pVQZ-33211 42.356 25.856 12.024 11.684 lytic function V(R,), originally suggested by Bukowski
aug-cc-pV5Z-33211 42.467 25.969 12.023 10.566 et al. [30] and that has been used in several similar
daug-cc-pVDZ-33211 40.773 26.602 12.871 16.279 potentials [23g, 31, 32]. The function is the sum of
daug-cc-pVTZ-33211 42.082 25.748 12.202 14.224 two terms, a short range term Vsh and an asymptotic
daug-cc-pVQZ-33211 42.386 25.918 12.056 11.518 term Vas:
daug-cc-pV5Z-33211 42.487 26.013 12.046 10.446
aug-cc-pVDZ 31.335 9.156 7.741 69.152
VðR, Þ ¼ Vsh ðR, Þ þ Vas ðR, Þ ð1Þ
aug-cc-pVTZ 38.598 20.206 10.195 23.574
aug-cc-pVQZ 41.524 23.230 11.148 12.896
aug-cc-pV5Z 42.207 24.633 11.547 11.022 where
daug-cc-pVDZ 33.929 14.422 8.996 63.499
daug-cc-pVTZ 39.372 23.523 11.265 21.683 Vsh ðR,Þ ¼ GðR,ÞeDðÞBðÞR ð2Þ
daug-cc-pVQZ 42.083 25.120 11.876 12.431
daug-cc-pV5Z 42.392 25.634 11.957 10.782
Downloaded by [RMIT University] at 05:48 17 August 2015

Table 2. Intermolecular energies in cm1, R in Å and  in degrees.

R  E R  E R  E
7.0000 0.0 0.387 5.0000 60.0 2.243 2.8500 30.0 9.484
6.0000 0.0 1.053 4.5000 60.0 4.364 2.7500 30.0 20.197
5.5000 0.0 1.878 4.0000 60.0 8.890 2.6000 30.0 111.126
5.0000 0.0 3.567 3.5000 60.0 17.274 2.5000 30.0 224.032
4.5000 0.0 7.278 3.0000 60.0 17.319 6.0000 60.0 0.705
4.2500 0.0 10.664 2.9500 60.0 14.184 5.5000 60.0 1.224
4.0000 0.0 15.789 2.9000 60.0 9.638 2.4000 150.0 171.941
3.8500 0.0 19.977 2.8500 60.0 3.320 2.3000 150.0 311.593
3.7500 0.0 23.300 2.7500 60.0 16.502 2.5000 15.0 282.719
3.6500 0.0 27.044 2.6000 60.0 75.369 2.5000 45.0 172.538
3.5000 0.0 33.269 2.5000 60.0 147.092 2.5000 135.0 109.412
3.2600 0.0 42.468 5.5000 90.0 1.080 2.5000 165.0 67.788
3.2500 0.0 42.738 5.0000 90.0 1.966 5.5000 30.0 1.614
3.1500 0.0 44.017 4.5000 90.0 3.796 5.0000 30.0 3.018
3.0749 0.0 42.591 4.0000 90.0 7.635 4.5000 30.0 6.035
3.0000 0.0 37.820 3.6608 90.0 12.024 4.0000 30.0 12.781
2.9500 0.0 31.962 3.5000 90.0 14.442 3.5000 30.0 26.208
2.9000 0.0 23.200 3.0000 90.0 12.156 3.0000 30.0 29.567
2.8900 0.0 21.029 2.9500 90.0 8.908 2.9500 30.0 25.241
2.8749 0.0 17.476 2.9000 90.0 4.368 2.9000 30.0 18.809
2.8600 0.0 13.656 2.8500 90.0 1.807 3.5000 150.0 15.735
2.8500 0.0 10.696 2.7500 90.0 20.883 3.0000 150.0 22.306
2.8000 0.0 6.598 2.6000 90.0 76.290 2.8500 150.0 15.771
2.7000 0.0 61.081 2.5000 90.0 142.970 2.8000 150.0 10.942
2.5000 180.0 60.674 6.0000 120.0 0.634 2.7500 150.0 4.106
2.6000 180.0 17.117 5.5000 120.0 1.087 2.6500 150.0 18.068
2.7000 180.0 7.059 5.0000 120.0 1.968 2.6000 150.0 35.015
2.8000 180.0 19.449 4.5000 120.0 3.786 2.5000 150.0 86.364
2.9000 180.0 24.833 4.0000 120.0 7.633 5.5000 150.0 1.109
2.9580 180.0 25.970 3.5000 120.0 14.794 5.0000 150.0 1.991
3.0000 180.0 26.189 3.0000 120.0 15.114 4.5000 150.0 3.808
3.0100 180.0 26.181 2.9500 120.0 12.459 4.0000 150.0 7.729
3.2500 180.0 22.231 2.9000 120.0 8.581 5.0000 180.0 1.100
3.5000 180.0 16.163 2.8500 120.0 3.157 5.5000 180.0 1.114
3.7500 180.0 11.226 2.7500 120.0 14.014 6.0000 180.0 0.658
4.0000 180.0 7.741 2.6000 120.0 65.676 31641 30.9 33.773
4.5000 180.0 3.800 2.5000 120.0 129.372
Accurate computations of the rovibrational spectrum 1417

and Table 3. Parameters of the analytic PES fitted to the ab initio


interaction energies.
X
2
Cl6 0
Vas ðR, Þ ¼ f6 ðBðÞRÞ  P ðcos Þ ð3Þ Parameter Value Parameter Value
l¼0,2...
R6 l
b0 13.35140347292370 g11 0.48875737093263
DðÞ, BðÞ, and GðR, Þ are expansions in Legendre b1 0.09056257250910 g12 0.66202397351030
polynomials (P0l ); b2 0.36222620495196 g13 0.61187512997162
b3 0.10752025627937 g14 0.19932910113773
X
5 b4 0.10729093395687 g15 0.02744036687153
BðÞ ¼ bl P0l ðcos Þ, ð4Þ b5 0.09867415096543 g20 1.04324447891595
l¼0 d0 2.53146116070142 g21 0.09402772497221
d1 0.07421267974884 g22 0.20171778104736
X
5 d2 0.08276071628834 g23 0.16229578225988
DðÞ ¼ dl P0l ðcos Þ, ð5Þ d3 0.01498993894000 g24 0.03785716244561
l¼0 d4 0.04330076241291 g25 0.00646574608438
d5 0.03802078606627 g30 0.10441433253613
X
5   g00 4.57186377381132 g31 0.00497315571833
GðR, Þ ¼ g0l þ g1l R þ g2l R2 þ g3l R3 P0l ðcos Þ ð6Þ g01 0.75361611194433 g32 0.02153063761904
l¼0 g02 0.72139454377371 g33 0.01500658270722
g03 0.78271512913456 g34 0.00145052999200
Downloaded by [RMIT University] at 05:48 17 August 2015

and g04 0.29669542735975 g35 0.0000000000000


g05 0.02516881784584 C06 0.01401238000863
6
x  xk =k!
g10 3.73235058315367 C26 0.00793751597814
f6 ðxÞ ¼ 1  e k¼o ð7Þ

is the Tang-Toennies damping function and x ¼ BðÞR. HeHF contour


5.5
bl, dl, gkl and Cl6 are adjustable parameters. The fitted
values of the corresponding PES parameters are pre-
5
sented in table 3.
−3
The fit of the ab initio values gives a standard error of
0.0106 cm1 and a maximum error of 0.3235 cm1. The 4.5
R (Angstrom)

geometry corresponding to this error is R ¼ 6.0 Å and


 ¼ 0 . A contour plot of the PES is presented in 4
figure 1. Two minima are found in the PES, the absolute
minimum has energy of 43.844 cm1 and it is located 3.5
at the He–HF linear configuration with the He atom at a −15
−42
distance of 3.1662 Å with respect to the HF center of 3 −24
mass. The surface has a second minimum at the He–FH
linear configuration, with the He at a distance
2.5
of 2.9989 Å with respect to the HF center of mass and 0 20 40 60 80 100 120 140 160 180
at an energy of 26.169 cm1. A saddle point with T (degrees)
energy of 16.882 cm1 is located at R ¼ 3.2277 Å and Figure 1. Contour plot in Jacobi coordinates of the
 ¼ 96.222 . This gives an energy barrier between the aug-cc-pV5Z-33211.
global and the local minima of 26.962 cm1.
In table 4, these results are compared to those Taking into account the good performance of the
previously available. They are considerably different method and basis set we use, we can consider the dif-
from those in Ref. [10, 12, 14], and the SAPT [18], but ferences with respect to the results in Ref. [10, 12, 14, 18]
very close to those of previous theoretical calculations large, and suggest that these studies need revision.
with the BCCD(T) model [21]. The SAPT surface is
characterized by a difference between the He–HF and
He–FH minima of only 3.55 cm1, whereas we get a 3. Rovibrational levels and fixed frame diffusion
difference of 17.675 cm1 (in good agreement with Ref. Monte Carlo calculations
[21]). The SAPT potential differs also considerably from
the experimental results of Ref. [14] (up to 14.65 cm1 in Using the analytical fit to the potential several low-lying
the absolute minimum). rovibrational energy levels were calculated using the
1418 J. L. C. Fajin et al.

Table 4. Stationary points. Comparison to previous results.

HeHF HeFH Saddle point


R(Å) E(cm1) R(Å) E(cm1) R(Å)  ( ) E(cm1)
This work 3.1662 43.844 2.9989 26.169 3.2277 96.222 16.882
CCSD(T)
HFD[10] 3.159 33.119
Exp. [12] 3.11  0.03a 24.1  1.7
Exp. [14] 3.113  0.001 25.0  0.2
SAPT[18] 3.2597 39.68 2.9581 36.13 3.2539 90.0 19.028
BCCD(T)[21] 3.1486 43.70 2.9899 25.88 3.2280 95.0 16.7532
a
This is an Rm value.

Table 5. Rovibrational energy levels, given in cm1 and labeled by the set of quantum numbers (J,Ka,Kc).

Level This Work CCSD(T) This Work BOUND LC-RAMP[11] ISE[16] SAPT[19] L-J(m,6)[14]
HeHF (0,0,0) 6.716 6.720 6.393 7.349 7.380 7.347
Downloaded by [RMIT University] at 05:48 17 August 2015

(1,0,1) 5.964 5.968 5.632 6.574 6.608 6.572


(2,0,2) 4.484 4.487 4.128 5.045 5.085 5.043
(3,0,3) 2.326 2.233 1.924 2.814 2.861 2.812
(4,0,4) 0.003 0.040 0.011
HeDF (0,0,0) 7.501 7.517
(1,0,1) 6.731 6.743
(2,0,2) 5.212 5.216
(3,0,3) 2.991 2.984
(4,0,4) 0.173 0.153

TRIATOM program suite [33] and also BOUND [34] the rotor with the z-axis coinciding with the HF
which is a close coupling approach. These calculations bond, i.e.,
are reported in table 5 and excellent agreement is
Hmon ¼ BðJ2x þ J2y Þ: ð9Þ
obtained between the two methods. In the BOUND
calculations the potential was re-expanded in a basis of Here, B is the rotational constant. In rigid-body, DMC
Legendre functions up to order eight. Of course, these (RBDMC) rotations and translations are treated exactly
methods are not readily extended to complexes contain- on the same footing [4, 8], i.e., the components of the
ing more than a single He atom and other approaches total angular momentum Jx, Jy Jz, are thought of as
must be used. In particular, quantum Monte Carlo generating rotations about the x- and y- and z-axes just as
methods [8] have been used successfully to study a large the momentum operators px, py, pz generate translations
variety of seeded He clusters. DMC methods, often along these axes. Of course doing this ignores the fact
coupled with importance sampling, have been used that angular momentum components do not commute
extensively to study the properties of seeded He with each other. However, for small enough rotational
droplets [4]. Here we develop a variant of DMC that is moves in the RBDMC method the errors introduced
designed to treat linear rotors, or collections of linear by this approximation to the Green function are gener-
rotors, embedded in He nanodroplets. ally negligible. A complication is apparent however; the
In the center-of-mass frame, the Hamiltonian for a Hamiltonian used in RBDMC is, of necessity, defined in
rigid linear rotor such as HF interacting with a He atom the laboratory frame (otherwise Coriolis terms arise
is given by which are hard to treat in DMC [4]). However, the rotor
Hamiltonian is diagonal only in a frame that rotates with
2 2
h the molecule. This problem can be avoided by recogniz-
H¼ r þ VðR, Þ þ Hmon ð8Þ ing that the monomer Hamiltonian can be expressed in
2
terms of the orbital angular momentum operator, i.e.,
where  is the reduced mass and the monomer
Hamiltonian, Hmon, is diagonal in a frame fixed on Hmon ¼ BL2 ð10Þ
Accurate computations of the rovibrational spectrum 1419

where L ¼ (Lx,Ly,Lz). In this procedure, rotational [6] A. Sarsa, Z. Bacic, J. W. Moskowitz, and K. E. Schmidt,
diffusive moves are made around the center-of-mass Phys. Rev. Lett. 88, 123401 (2002).
[7] H.-C. Chang and W. Klemperer, J. Chem. Phys. 104, 7830
axes which are always aligned with the laboratory axes. (1996).
The ground state energy obtained using this method is [8] V. Buch, J. Chem. Phys. 97, 726 (1992).
E ¼ 6.718  0.059 cm1 in excellent agreement with [9] (a) J. M. Hutson and B. J. Howard, Mol. Phys. 45, 769
other methods reported in table 5. An expanded (1982); (b) C. M. Lovejoy and D. J. Nesbitt, Chem. Phys.
discussion of this algorithm, including the incorporation Lett. 147, 490 (1988); (c) J. M. Hutson, J. Chem. Phys. 89,
4550 (1988); (d) J. M. Hutson, J. Chem. Phys. 91, 4448
of importance sampling, together with application to
(1989); (e) G. Chalasinski, M. M. Szczesniak, and
the calculation of ground and excited states of larger B. Kukawska-Tarnawska, J. Chem. Phys. 94, 6677
clusters, doped with HF, HCl, HBr and HI molecules (1981); (f ) G. Chalasinski, S. M. Cybulski,
will be described elsewhere. M. M. Szczesniak, and S. Scheiner, J. Chem. Phys. 91,
7048 (1989).
[10] W. R. Rodwell, L. T. Sin-Fai-Lam, and R. O. Watts,
Mol. Phys. 44, 225 (1981).
4. Summary and conclusions [11] J. Tennyson and B. T. Sutcliffe, J. Chem. Phys. 79, 43
(1983).
The potential energy surface of the He–HF van der [12] C. V. Boughton, R. E. Miller, P. F. Vohralik, and
Waals complex is obtained by fitting 110 interaction R. O. Watts, Mol. Phys. 58, 827 (1986).
energies obtained with the CCSD(T) method and the [13] J. A. Barker and A. Pompe, Aust. J. Chem. 21, 1683
(1968).
Downloaded by [RMIT University] at 05:48 17 August 2015

aug-cc-pV5Z basis set extended with a set of 3s3p2d1f1g [14] C. M. Lovejoy and D. J. Nesbitt, J. Chem. Phys. 93, 5387
mid-bond functions. This basis set is selected after (1990).
a systematic study carried out at four intermolecular [15] M. J. Smith and H. Rabitz, Chem. Phys. 150, 361 (1991).
geometries. The potential has two linear minima, i.e. [16] T. Slee and R. J. Le Roy, J. Chem. Phys. 99, 360
He–HF and He–FH, with distances from the He atom (1993).
[17] J. M. Hutson and R. J. Le Roy, J. Chem. Phys. 83, 1197
to the HF centre of mass of 3.1662 and 2.9989 Å,
(1985).
respectively; and with binding energies of 43.844 and [18] R. Moszynski, P. E. S. Wormer, B. Jeziorski, and
26.169 cm1, respectively. The low-lying rovibrational A. van der Avoird, J. Chem. Phys. 101, 2811 (1994).
energy levels of the complex are computed using [19] R. Moszynski, B. Jeziorski, A. van der Avoird, and
different methods and good agreement is obtained P. E. S. Wormer, J. Chem. Phys. 101, 2825 (1994).
[20] R. Moszynski, F. de Weerd, G. C. Groenenboom, and
among these.
A. van der Avoird, Chem. Phys. Lett. 263, 107 (1996).
[21] T. Stoecklin, A. Voronin, and J. C. Rayez, Chem. Phys.
Acknowledgements 294, 117 (2003).
[22] C. Reese, T. Stoecklin, A. Voronin, and J. C. Rayez,
A. & A. 430, 1139 (2005).
This work has been supported by the European [23] (a) C. R. Munteanu, J. L. Cacheiro, B. Fernández, and
Research and Training Network NANOQUANT, J. Makarewicz, J. Chem. Phys. 121, 1390 (2004);
contract No. MRTN-CT-2003-506842; by the Spanish (b) J. L. Fajı́n, J. L. Cacheiro, and B. Fernández,
Ministerio de Educación y Ciencia and FEDER J. Chem. Phys. 121, 4599 (2004).
[24] D. Webb and K. N. Rao, J. Mol. Spect. 28, 121 (1968).
(CTQ2005-01076 project); by the Petroleum Research [25] S. F. Boys and F. Bernardi, Mol. Phys. 19, 553 (1970).
fund, administered by the American Chemical Society; [26] F. M. Tao and Y. K. Pan, J. Chem. Phys. 97, 553
and by the US National Science Foundation though (1992).
grant 0202185. We acknowledge computer time from [27] C. R. Munteanu, J. L. Cacheiro, and B. Fernández,
CESGA. J. Chem. Phys. 120, 9104 (2004).
[28] T. Helgaker, H. J. Aa. Jensen, P. Jørgensen, J. Olsen, K.
Ruud, H. Ågren, A. A. Auer, K. L. Bak, V. Bakken,
O. Christiansen, S. Coriani, P. Dahle, E. K. Dalskov,
References T. Enevoldsen, B. Fernández, C. Hättig, K. Hald,
A. Halkier, H. Heiberg, H. Hettema, D. Jonsson,
[1] J. P. Toennies and A. F. Vilesov, Annu. Rev. Phys. Chem. S. Kirpekar, R. Kobayashi, H. Koch, K. V. Mikkelsen,
49, 1 (1998). P. Norman, M. J. Packer, T. B. Pedersen, T. A. Ruden,
[2] C. Callegari, K. K. Lehmann, R. Schmied, and G. Scoles, A. Sánchez, T. Saue, S. P. A. Sauer, B. Schimmelpfennig,
J. Chem. Phys. 115, 10090 (2001). K. O. Sylvester-Hvid, P. R. Taylor, O. Vahtras, DALTON –
[3] M. Behrens, U. Back, R. Frochtenicht, M. Hartmann, and an electronic structure program, release 1.2 (2001).
M. Havenith, J. Chem. Phys. 107, 7179 (1997). [29] J. F. Stanton, J. Gauss, J. D. Watts, M. Nooijen, N.
[4] E. Lee, D. Farrelly, and K. B. Whaley, Phys. Rev. Lett. Oliphant, A. S. Perera, P. G. Szalay, W. J. Lauderdale,
83, 3812 (1999). S. A. Kucharski, S. R. Gwaltney, S. Beck, A. Balkova,
[5] K. Nauta and R. E. Miller, J. Chem. Phys 113, 10158 D. E. Bernholdt, K. K. Baeck, P. Rozyczko, H. Sekino,
(2000). C. Hober, R. J. Bartlett, J. Almlöf, P. R. Taylor,
1420 J. L. C. Fajin et al.

T. Helgaker, H. J. Aa. Jensen, P. Jørgensen, J. Olsen, [32] T. B. Pedersen, J. L. Cacheiro, B. Fernández, and
ACES II is a program product of the Quantum Theory H. Koch, J. Chem. Phys. 117, 6562 (2002).
Project, University of Florida, 1996, 32611 . [33] J. Tennyson, S. Miller, and C. R. Le Sueur, Comput.
[30] R. Bukowski, J. Sadlej, B. Jeziorski, P. Jankowski, Phys. Commun 75, 339 (1993).
K. Szalewicz, S. A. Kucharski, H. L. Williams, and [34] J.M. Hutson, BOUND computer code, version 4 (1992),
B. M. Rice, J. Chem. Phys. 110, 3785 (1999). distributed by the Collaborative Computational
[31] R. R. Toczylowski and S. M. Cybulski, J. Chem. Phys. Project No. 6 of the Science and Engineering Research
112, 4604 (2000). Council (UK).
Downloaded by [RMIT University] at 05:48 17 August 2015

You might also like