You are on page 1of 8

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/233762391

Rotational Spectra and van der Waals Potentials of Ne-Ar

Article  in  The Journal of Chemical Physics · January 1995


DOI: 10.1063/1.468904

CITATIONS READS
52 32

8 authors, including:

Fraser Gerald Francis J. Lovas


National Institute of Standards and Technology National Institute of Standards and Technology
86 PUBLICATIONS   2,884 CITATIONS    282 PUBLICATIONS   11,794 CITATIONS   

SEE PROFILE SEE PROFILE

R.D. Suenram T. Emilsson


University of Virginia APL Engineered Materials
398 PUBLICATIONS   8,326 CITATIONS    79 PUBLICATIONS   2,317 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Dimethyl carbonate View project

Remote sensing View project

All content following this page was uploaded by Francis J. Lovas on 03 June 2014.

The user has requested enhancement of the downloaded file.


Rotational spectra and van der Waals potentials of Ne–Ar
J.-U. Grabow, A. S. Pine, G. T. Fraser, F. J. Lovas, and R. D. Suenram
Molecular Physics Division, National Institute of Standards and Technology,
Gaithersburg, Maryland 20899
T. Emilsson, E. Arunan and H. S. Gutowsky
Noyes Chemical Laboratory, University of Illinois, Urbana, Illinois 61801
~Received 29 August 1994; accepted 11 October 1994!
The high sensitivity and resolution of Fourier-transform microwave spectroscopy using a pulsed jet
coaxial to a Fabry–Perot resonator have been exploited to measure pure rotational transitions of
several isotopomers of the weakly polar Ne–Ar van der Waals dimer in natural abundance.
Transitions of the most abundant isotopomer, 20Ne– 40Ar, could be observed with an excellent
signal-to-noise ratio with a single polarization pulse. The ground-state rotational constants for this
species yield a zero-point separation of R 0.360.7 pm. Simple model van der Waals potentials have
been fit to the microwave transitions for the various isotopomers, providing estimates of the
equilibrium spacing at the well minimum of R e 5348.0~2! pm. More elaborate potentials based on
ab initio calculations or on molecular-beam scattering cross sections and thermodynamic and
transport properties have also been tested. The induced electric dipole moment is estimated to be
m057.3~1.6!310233 C m @0.0022~5! D# by comparison of p/2 polarization pulses with a reference
molecule ~Ar–CO2! whose dipole moment is known from Stark effect splitting measurements.
Uncertainties in parentheses are one standard deviation. © 1995 American Institute of Physics.

I. INTRODUCTION depth, binding energy, and vibrational intervals are quite


model dependent. We have also estimated the induced zero-
The finite, but small, induced electric dipole moment of point dipole moment by comparison of p/2 polarization
the heteronuclear rare gas dimers gives rise to a pure rota- pulses with Ar–CO2 , and have obtained a measure of the
tional microwave spectrum. Jäger, Xu, and Gerry1 recently field gradient at the Ne nucleus from the hyperfine splittings
reported pure rotational microwave transitions of Ne–Xe, of 21Ne– 40Ar.
Ar–Xe, and Kr–Xe using a Balle–Flygare-type2 pulsed-jet These extremely accurate microwave frequency mea-
Fourier-transform spectrometer with a nozzle coaxial to the surements on the rare gas heterodimers provide a stringent
microwave resonator.3 The multiple isotopes of Xe, Kr, and test for any ab initio electronic structure calculations8 or
Ne observed in natural abundance enabled Jäger et al.1 to model potentials constructed from thermodynamic, transport
estimate the equilibrium bond lengths of these species using and molecular beam scattering cross section data.9,10 No
the substitution procedure of Costain.4 From the observed spectroscopic data have been incorporated into any of the
hyperfine splittings in the 131Xe–rare gas dimers, they ob- existing heterodimer potentials, although vacuum ultraviolet
tained the nuclear quadrupole coupling constants which pro- fluorescence measurements have been included for some
vide a measure of the induced charge distortion at the Xe homodimers.9,11,12 Since the Born–Oppenheimer potentials
nucleus. Also, they estimated the induced dipole moments of for all diatomics are one dimensional, their rotation–
these species from the p/2 polarization pulse conditions ref- vibration eigenvalues can be obtained numerically with high
erenced to the dipole moment of Ar–CO2 known from Stark precision and speed for comparison to the spectral measure-
effect measurements.5 More recently, this group also ob- ments. A knowledge of the van der Waals potentials for the
served the Ne–Kr and Ar–Kr dimers6 and Ne2 –Kr and rare gas dimers is basic to our understanding of the isotropic
Ne2 –Xe trimers.7 short-range overlap and exchange repulsion and long-range
Here, we report the observation of several isotopomers dispersion forces in all intermolecular interactions.
of the less polar Ne–Ar species using similar instrumenta-
tion. From the zero-point separations determined from the
II. EXPERIMENTAL METHODS
ground-state rotational constants for the isotopomers, we
found that the substitution procedure in this highly anhar- The Fourier-transform microwave ~FTMW! spectrom-
monic species does not produce an isotopically invariant eters at NIST and Illinois have previously been
equilibrium bond length, R e . Therefore, we fit the micro- described,13,14 but they have been modified somewhat to en-
wave transitions directly to a variety of simple analytical hance the sensitivity for the weakly polar species studied
potentials which yielded a more consistent estimate for R e . here. Primarily, the polarizing driver power has been in-
Preliminary fits of the potentials to the transitions of the creased with a low-noise solid-state amplifier ~1 W from 6 to
more abundant isotopes provided excellent predictions for 18 GHz! at NIST and traveling-wave tubes ~20 W from 4 to
the rarer species, eliminating the need to scan. The micro- 12 GHz, 10 W from 12 to 18 GHz! at Illinois. An 80 dB PIN
wave transitions reliably specify the radial position and the diode switch on the output of the TWTs suppressed the ex-
curvature of the potential well at its minimum, but the well cess noise during the free-induction decay. A fast single-pole

J.Downloaded¬22¬Sep¬2009¬to¬129.6.168.96.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright;¬see¬http://jcp.aip.org/jcp/copyright.jsp
Chem. Phys. 102 (3), 15 January 1995 0021-9606/95/102(3)/1181/7/$6.00 © 1995 American Institute of Physics 1181
1182 Grabow et al.: Rotational spectra of Ne–Ar

out! based on estimates of its frequency ~;11 627 MHz!


from empirical potentials.10,15 Once the conditions were op-
timized for the NIST spectrometer, we could observe this
transition on a single polarization pulse for the power spec-
trum shown in Fig. 1~a!. The corresponding voltage or am-
plitude spectrum has a signal-to-rms-noise ratio of ;14:1.
The spectra appear as Doppler doublets since the nozzle is
coaxial with the propagation vector of the radiation in the
standing-wave cavity. The true rest frequency is the average
of the Doppler pair.
It was then straightforward to estimate the J53–2 and
J51–0 transitions for this isotopomer and all the corre-
sponding transitions predicted for the 8.8% abundant
22
Ne– 40Ar species were found within a search window of a
few MHz. For the much less abundant isotopes, 36Ar, 38Ar,
and 21Ne, many thousands of pulses would have to be inte-
grated to obtain a usable signal-to-noise ratio; so we needed
to predict their frequencies within the cavity bandwidth
~;100 kHz! to obviate searches. This was accomplished by
potential modeling to be described later, with the result for
the 20Ne– 36Ar J53–2 transition shown in Fig. 1~b! requiring
about 64 000 polarization pulses. Spectra of the 21Ne– 40Ar
and 22Ne– 36Ar J51–0 transitions recorded at Illinois with

FIG. 1. FTMW power spectra of Ne–Ar recorded with the NIST apparatus:
~a! the J52–1 Doppler-doubled transition of 20Ne– 40Ar with a single po-
larization pulse using 4096 points digitized at intervals of 125 ns; ~b! the
J53–2 transition of 20Ne– 36Ar after 64 000 pulses using 1024 points at 125
ns ~356 min recording time!.

double-throw PIN diode switch feeds the cavity with the


polarizing pulse while protecting the detection system from
the high power reflection, and subsequently transmits the
free-induction-decay radiation through a low-noise micro-
wave amplifier to an image-rejection mixer for heterodyne
detection.3 The larger diameter mirrors at Illinois ~50 cm vs
40 cm at NIST! reduce diffraction losses at the lower fre-
quencies required for the J51–0 transitions of Ne–Ar. The
pumping system at Illinois is also larger, allowing a 10 Hz
gas pulse rate vs 3 Hz at NIST, which was valuable for the
multiple pulse integration required for the rarer isotopes. In
both cases a 1 mm pulsed solenoid valve was used; the gas
mixture was about 4% $1%% Ar in first-run Ne at a backing
pressure of 800 $270% kPa at NIST $Illinois%. These rather
different sample conditions probably reflect the sensitivity of
the expansion to the nozzle plunger tension and the channel
opening through the cavity mirror.

III. SPECTRAL ASSIGNMENTS AND ANALYSIS FIG. 2. FTMW power spectra of Ne–Ar recorded with the Illinois appara-
tus: ~a! the J51–0 Doppler-doubled, hyperfine-split transition of 21Ne– 40Ar
The J52–1 transition of the major isotopic species,
after 21 000 pulses using 512 points at 500 ns ~4.3 min recording time!; ~b!
20
Ne– 40Ar, was first observed at 11 652.3074~4! MHz ~un- the J51–0 transition of 22Ne– 36Ar after 240 000 pulses using 512 points at
certainties in parentheses are one standard deviation through- 300 ns ~50 min!.

J. Chem. Phys., Vol. 102, No. 3, 15 January 1995


Downloaded¬22¬Sep¬2009¬to¬129.6.168.96.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright;¬see¬http://jcp.aip.org/jcp/copyright.jsp
Grabow et al.: Rotational spectra of Ne–Ar 1183

TABLE I. Microwave transition frequencies, rotational constants, substitution structure coordinates and bond lengths for Ne–Ar isotopomers.

M ~Ne!–M ~Ar!a 20240 22240 20236 22236 20238

n~J51–0! ~MHz! 5 828.9326~5! 5 489.0125~5! 6 036.0797~15! 5 696.9714~20! 5 927.2332~20!


n~J52–1! ~MHz! 11 652.3074~4! 10 973.1708~5! 12 066.1478~15! 11 388.6637~20! 11 848.6960~20!
n~J53–2! ~MHz! 17 464.5156~5! 16 447.5767~6! 18 084.1299~20! 17 069.7523~20! 17 758.5610~20!

B 0 ~MHz! 2 914.9286~5!b 2 744.9100~4! 3 018.5398~12! 2 848.9249~18! 2 964.0965~18!


D J ~kHz! 231.01~13! 201.77~15! 249.79~50! 219.45~75! 239.80~75!
H J ~Hz! 270.8~7.0! 261.1~7.0! 287.2~21.0! 263.2~31.0! 279.2~31.0!

R 0 ~pm! 360.7017~2! 360.2686~2! 360.9613~7! 360.5343~11! 360.8253~11!


r s ~Ne! ~pm! 235.6135~2! 228.0114~2! 227.2598~7! 219.4219~11! 231.6142~40!c
r s ~Ar! ~pm! 117.9171~4! 125.4667~4! 126.3349~14! 134.1146~21! 121.9758~21!
R s ~pm! 353.5306~5! 353.4782~5! 353.5947~18! 353.5364~26! 353.5900~45!
R e ~pm! 346.360~12! 346.688~12! 346.228~40! 346.539~55! 346.355~92!
a
M /u519.992 439 1 for 20Ne; 21.991 383 7 for 22Ne; 39.962 383 1 for 40Ar; 35.967 545 6 for 36Ar; 37.962 732 2 for 38Ar. Reference 16.
b
1s uncertainties in parentheses are propagated from measurement precision, not least-squares, and do not include model inadequacies.
c
r s ~Ne!5r s ~Ar!M ~38Ar!/M ~20Ne!.

21 000 and 240 000 polarization pulses respectively are where r s is the coordinate of the substituted atom from the
shown in Fig. 2~a! and 2~b!. This has enabled us to observe center of mass, B 0 is the measured ground-state rotational
the hyperfine structure associated with the nonzero nuclear constant of the ‘‘parent’’ molecule, B 80 is for a singly substi-
spin ~I53/2! for 21Ne. Frequency calibration differences be- tuted species, and the ‘‘substituted’’ reduced mass is
tween the two spectrometers were less than 0.4 kHz for the
D m 5M •Dm/ ~ M 1Dm ! , ~4!
transitions measured in common.
where M is the total mass of the parent molecule and Dm is
A. Bond length
the change in mass of the substituted atom. Then the ‘‘equi-
The observed transition frequencies and resulting rota- librium’’ internuclear separation, for R s 5r s ~Ne!1r s ~Ar!, is
tional constants for the spinless Ne–Ar dimers are given in given by
Table I. Since only three transitions were observed for each
R e .2R s 2R 0 . ~5!
species, only B 0 , D J , and H J could be determined, so they
must be considered effective values ignoring higher-order This relation strictly holds only for diatomics with a small
centrifugal distortion. Note the H J values are negative in all vibrational dependence, DB, to the B value and with
cases, and the large ratio of H J /D J indicates that extrapola- DB/DB 8 5 (B e /B e8 ) 3/2 , which may not be appropriate for
tion to higher transitions is suspect. From the calculated B 0 , these highly anharmonic rare gas dimers. The r s coordinates
we give in Table I the zero-point separations, R 0 , using and the internuclear distances R s and R e are also given in
Table I for the various isotopic species. Curiously, the R s
m •R 20 5\ 2 / ~ 2B 0 ! , ~1!
values are more consistent among the isotopomers than the
where the reduced mass in terms of the atomic masses is R e ; as was also found for Ne–Xe by Jäger et al.1 We will
show later that these R e values are significantly lower than
m 5M ~ Ne! M ~ Ar! / @ M ~ Ne! 1M ~ Ar!# . ~2!
those obtained for simple model potentials fit to the same
The resulting ground-state R 0 vary by about 0.2% among the microwave data with nearly experimental precision.
isotopes due to the differing zero-point averages for ^1/R 2&.
Costain4 has given a prescription for an isotopic substitution B. Hyperfine interactions
structure that can lead to a more invariant ‘‘equilibrium’’ or
structural bond length. He defines The frequencies for the hyperfine transitions in the
21
Ne– 40Ar species are given in Table II along with the re-
D m •r 2s 5 ~ \ 2 /2 !@~ 1/B 80 ! 2 ~ 1/B 0 !# , ~3! sulting nuclear quadrupole coupling constants, eqQ.17 The
J51–0 F assignments are based on the expected 5:4 split-
ting ratio and are confirmed by the 2:3:1 relative
intensities.17 With these assignments, eqQ is negative, indi-
TABLE II. Hyperfine transition frequencies and quadrupole coupling con-
cating a negative field gradient at the Ne nucleus. Though
stant for 21Ne– 40Ar ~I53/2 for 21Ne!.
hyperfine structure was also partially resolved for the higher
J8 J9 F8 F9 n ~MHz! eqQ ~kHz! J transitions, we obtained reliable frequencies only for the
strongest DF5DJ, F5J1I components given in Table II.
1 0 1/2 3/2 5 650.7262~10!
1 0 5/2 3/2 5 650.7199~10! 230~2!
1 0 3/2 3/2 5 650.7128~10! C. Dipole moment
2 1 7/2 5/2 11 296.2574~25!
3 2 9/2 7/2 16 931.3818~25! An estimate of the zero-point induced dipole moment,
m0 , of Ne–Ar was made by optimizing the polarization

J. Chem. Phys., Vol. 102, No. 3, 15 January 1995


Downloaded¬22¬Sep¬2009¬to¬129.6.168.96.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright;¬see¬http://jcp.aip.org/jcp/copyright.jsp
1184 Grabow et al.: Rotational spectra of Ne–Ar

TABLE III. van der Waals potential parameters for Ne–Ar.

Potential HFD-Ba HFD-B* b LJ MO X-6 MS HL

ec ~cm21! 46.977 1 45.787 36 51.327 09 42.007 76 48.598 52 47.156 94 45.554 68


6ded 60.005 5 60.034 60.005 0 60.004 4 60.51 60.81
R e ~pm! 348.89 348.118 195 347.804 005 348.213 433 347.982 608 348.046 371 348.038 323
6d R e 60.000 82 60.004 8 60.000 9 60.000 69 60.041 60.041
a 9.692 905 67 11.225 411 8 12. 6.733 504 14.594 038 13.343 396 0.418 659
6da 60.001 0 60.002 1 60.22 60.071
b 22.273 808 51 21.507 555 8.027 47 1.846 545
6db 60.0048 61.39 60.32
se ~kHz! 27 155 2.81 21.39 3.99 2.64 2.60 2.68

c( e ,R e )
f
0.0470 0.9997 0.8153 0.5177 20.999 8 20.999 8
c~e,a! 0.1769 0.4501 20.999 9 20.999 9
c~e,b! 0.7139 21.000 0 21.000 0
c(R e , a ) 20.4252 20.5307 0.999 8 0.999 8
c(R e , b ) 20.6657 0.999 9 0.999 9
c~a,b! 1.000 0 1.000 0

f ~e,e!g 0.018 0.024 0.025 0.022 0.000 18 0.000 12


f (R e ,R e ) 0.019 0.024 0.023 0.021 0.000 35 0.000 36
f ~a,a! 0.040 0.022 0.000 17 0.000 36
f ~b,b! 0.013 0.000 07 0.000 07
f ( e ,R e ) 0.16 0.024 0.12 0.15 0.023 0.022
f ~e,a! 0.21 0.15 0.019 0.022
f ~e,b! 0.11 0.011 0.007
f (R e , a ) 0.19 0.14 0.023 0.029
f (R e , b ) 0.12 0.017 0.017
f ~a,b! 0.010 0.018

D 0h ~cm21! 34.03 32.89 38.51 29.09 35.74 34.28 32.69


n~v 5120!i ~cm21! 19.10 18.79 18.99 18.73 18.86 18.83 18.83
n1~J5120!j ~MHz! 4837.0 4816.1 4875.5 4790.5 4843.5 4831.3 4831.5
a
Barow and Aziz, Ref. 10. D c 51.44, c 6 51.097 818 26, c 8 50.342 846 23, c 10 50.301 039 22, c 12 50.744 832 25.
b
Adjusted to fit present data with the D c and c n coefficients fixed as in footnote a.
c
Significant figures of parameters necessary to reproduce calculation, see Refs. 32 and 33.
d
6 uncertainties in parameter units are one standard deviation of the model fits and do not reflect uncertainties in the fundamental constants or atomic masses.
e
Weighted least squares standard deviation, except rms deviation for fixed HFD-Ba potential.
f
Correlation coefficients between parameters in parentheses.
g
Freedom and co-freedom coefficients between parameters in parentheses, see Refs. 32 and 33.
h
Zero-point dissociation or binding energy.
i
v 51-0 vibrational interval.
j
J51-0 rotational transition in v 51 level.

power and pulse length, t, for the J52–1 transition at ab initio Hartree–Fock calculation of the induced dipole mo-
11 652.3074 MHz, resulting in the 2m0E t /\5p/2 condition ment function, m(R), in several rare gas heterodimers for
for the maximum signal strength. Since the absolute field comparison to far-infrared collision-induced absorption. For
distribution inside the cavity is not known, we calibrated the Ne–Ar at the zero-point separation, their m~R 0!'9.8310233
system for a p/2 pulse for the nearby J53–2 transition of C m'0.0029 D is in reasonable agreement with our mea-
Ar–CO2 at 10 958.6522 MHz. The dipole moment for surement, though we should compare the zero-point aver-
Ar–CO2 @m052.266~7!310231 C m50.067 93~20! D# is aged ^c0um(R)uc0&. This moment is dominated by repulsive
known independently from the Stark effect measurements of overlap effects with a partial ~'20%! compensation due to
Steed, Dixon, and Klemperer.5 In both cases, we measured dispersion.18 The Ne–Ar dipole moment is 3– 6 times
the p/2 polarizing power transmitted through the cavity smaller than for the other rare gas heterodimers previously
~40~1! mW and t56.5~5! ms for Ne–Ar and 0.25~1! mW and observed.1,6
t52.8~2! ms for Ar–CO2 in Ne carrier gas!. The relative
transmission factors ~input3output coupling! of the cavity
were within 10% at the two frequencies. The effective mo- IV. VAN DER WAALS POTENTIALS
ments for the transitions are corrected for their J dependence
by averaging [((J11) 2 2m 2 )/(2J11)(2J13)] 1/2 over all Barrow and Aziz10 have constructed an empirical poten-
m,17 yielding the Ne–Ar moment, m057.3~1.6!310233 C m tial for Ne–Ar based on ab initio Hartree–Fock calculations
50.0022~5! D. These uncertainties are dominated by our of Alrichs, Penco, and Scoles,8 with damped dispersion ac-
judgment on the p/2 polarization pulse duration. cording to the calculations of Kumar and Meath19 and Stan-
Birnbaum, Krauss, and Frommhold18 have reported an dard and Certain.20 Their potential, denoted HFD-B, is

J. Chem. Phys., Vol. 102, No. 3, 15 January 1995


Downloaded¬22¬Sep¬2009¬to¬129.6.168.96.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright;¬see¬http://jcp.aip.org/jcp/copyright.jsp
Grabow et al.: Rotational spectra of Ne–Ar 1185

V ~ R ! 5 e @ A exp~ 2 a x1 b x 2 ! 2F ~ x !~ c 6 /x 6 1c 8 /x 8 sis of the Barrow–Aziz potential. Because of the large num-


ber of significant potential parameters, even in this truncated
1c 10 /x 10 1c 12 /x 12 !# , ~6! series, it is not conveniently refined by fitting the limited
where x5R/R e and microwave data obtained here.
We have also fit our microwave data to a variety of
F ~ x ! 5exp@ 2 ~ D c /x21 ! 2 # for x,D c ; simple classical potentials with few parameters of the forms
F ~ x ! 51 for x.D c . ~7! below:
Lennard-Jones ~LJ!:
Two constraints apply so that the potential minimum is at
R5R e (x51) and the well depth is e, namely, V ~ R ! 5 e @ 1/x 12 22/x 6 # ; ~11!

a 52 b 1 @ F d 22F c D c ~ D c 21 !# / @ F c 21/F ~ 1 !# , ~8a! Morse ~MO!:

A5 @ F ~ 1 ! F c 21 # exp~ a 2 b ! , ~8b! V ~ R ! 5 e @ exp„2 a ~ 12x ! …22 exp„a ~ 12x ! …# ; ~12!

where F c 5c 6 1c 8 1c 10 1c 12 and F d 56c 6 18c 8 110c 10 Exponential-6 ~X-6!:


112c 12 . Barrow and Aziz10 then adjusted e, R e , b, D c , and V ~ R ! 5 @ e / ~ a 26 !#@ 6 exp„a ~ 12x ! …2 a /x 6 # ; ~13!
c 12 to fit the best available molecular beam scattering,21–23
virial coefficients,24 diffusion25 and viscosity26 data; their pa- Maitland–Smith ~MS!: 30

rameters are given in Table III.


V ~ R ! 5 @ e / ~ g 26 !#@ 6/x g 2 g /x 6 # ; g 5 a 1 b ~ x21 ! ;
We then add to V(R) the centrifugal potential, ~14!
V rot~ R ! 5\ 2 J ~ J11 ! / ~ 2 m R 2 ! , ~9! Hajigeorgiou–LeRoy ~HL!:31
to account for the rotational contribution to the energy and
V ~ R ! 5 e $ @ 12 ~ 1/x 6 ! exp~ 2 a z2 b z 2 !# 2 21 % . ~15!
solve the one-dimensional Schrödinger equation for the ei-
genvalues numerically using the Numerov–Cooley27 proce- In all these potentials the dimensionless radial x and z pa-
dure as implemented by LeRoy.28 By integrating over the rameters are defined as above in Eqs. ~6! and ~10!. The
range from 200 to 1200 pm in steps of 1 pm, we obtained Lennard-Jones has only two adjustable parameters, e and R e ,
convergence of the eigenvalues to 1 part in 108, comparable while the Morse and Exponential-6 have three ~adding a!
to our experimental precision. and the others four ~adding b!.
The Barrow–Aziz parameters for the HFD-B potential Results of the weighted least-squares fits to our micro-
yielded a J52–1 transition frequency for 20Ne– 40Ar about wave transitions are given in Table III. Of these potentials,
25 MHz below where it was found. Although this was ad- the least flexible Lennard-Jones gives the worst standard de-
equate to initiate a search for the main isotopes, the discrep- viation, s'21 kHz, and the e and R e parameters are highly
ancies were about 50 000 times our experimental accuracy correlated. The Morse potential yields s'4.0 kHz with rea-
and we needed much better predictions for the rarer isotopes. sonably uncorrelated parameters even though the long-range
Therefore, we adjusted some of the parameters to fit the ini- part of the potential is not very realistic for the dispersive
tial microwave results, eventually finding all of the isoto- van der Waals interaction. The Exponential-6 potential gives
pomer transitions listed in Tables I and II. Our final least- a good overall fit with s'2.6 kHz with uncorrelated param-
squares fits included all the transitions, with those involving eters, which is slightly better than the adjusted HFD-B* with
the rarer isotopes, 36Ar, 38Ar, or 21Ne, weighted by a factor of the same number of parameters. The modified Lennard-Jones
0.1 from the more abundant species because of a lower potentials of Maitland–Smith30 and Hajigeorgiou–LeRoy31
signal-to-noise ratio. The adjusted HFD-B* potential param- also yield comparably good s though all four parameters in
eters, e, R e , and b, are also given in Table III for comparison these potentials are highly correlated. The correlation coeffi-
to the Barrow–Aziz values; e decreases by '2.5% and R e cients can be reduced by fixing a or b, with a consequent
decreases by '0.22%. The weighted standard deviation of increase in the residuals, but the remaining parameters are
the fit was s'2.8 kHz. We also could obtain a comparable, still strongly interdependent. A more reliable estimate of the
but slightly worse fit, by varying D c instead of b, but the fit interdependence between the various potential parameters is
to these limited data would diverge if we tried to fit all four given by the corresponding freedom32 and cofreedom33 val-
parameters simultaneously. ues in Table III.
Ogilvie and Wang15 have also given preliminary esti- For the X-6 potential, the residuals are generally 1 to 3
mates for these transitions by first fitting the Barrow–Aziz kHz for the Ne– 40Ar and Ne– 38Ar isotopomers, whereas
HFD-B potential for Ne–Ar to a Dunham-type expansion of they are 1 to 6 kHz for 20Ne– 36Ar and 8 to 22 kHz for
22
the form, Ne– 36Ar. This may be signaling a breakdown in the Born–
Oppenheimer approximation, but the data are not extensive
V ~ R ! 5a 0 z 2 @ 11S n a n z n # , n51,2,...,10, ~10!
enough to be definitive. The other potentials show similar
where z52(R2R e )/(R1R e ). These a n coefficients can be mass-related deviations.
related analytically29 to the Dunham Y kl vibration–rotation It is apparent from these results in Table III that rather
parameters from which the transition frequencies can be unsophisticated potentials can be used for interpolation and
evaluated. They obtained a J52–1 transition frequency extrapolation of isotopic microwave data, almost to within
about 2 MHz higher than obtained by direct numerical analy- experimental accuracy, without recourse to rotational con-

J. Chem. Phys., Vol. 102, No. 3, 15 January 1995


Downloaded¬22¬Sep¬2009¬to¬129.6.168.96.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright;¬see¬http://jcp.aip.org/jcp/copyright.jsp
1186 Grabow et al.: Rotational spectra of Ne–Ar

distinguish the Maitland–Smith from the Barrow–Aziz


HFD-B. Our adjusted HFD-B* and Hajigeorgiou–Leroy are
also close to each other and about 2% to 3% shallower in
depth and area than HFD-B; our Exponential-6 is about 3.5%
deeper. The Lennard-Jones is much deeper and the Morse
much shallower; so they would not be expected to agree with
the bulk thermodynamic24 or transport25,26 data. All the po-
tentials cross the zero energy axis for R'311~1! pm. For the
repulsive wall shown logarithmically in Fig. 3~b!, the
HFD-B is slightly lower than the other potentials for R,300
pm, but all are within a factor of two at R5200 pm except
for the LJ.
Another key criterion for the potentials is the long-range
R 26 dispersion energy, expected to be in accord with the
calculations of Kumar and Meath.19 Barrow and Aziz10 use
this value of c 6'1.098, which we do not vary in our adjusted
HFD-B*. The Morse potential has an inappropriate exponen-
tial attraction. The Lennard-Jones has c 652; the Maitland–
Smith has c 651. The Exponential-6 potential has an effec-
tive c 65a/~a26! ~'1.7 in this case!; and the Hajigeorgiou–
LeRoy has an effective c 652 exp~22a24b! ~'0.0005 here,
which is anomalously small!. These model potentials may
have difficulty in fitting phenomena sensitive to the long-
range forces.
In Table III, we also give the zero-point dissociation en-
ergy, D 0 , the first vibrational interval, v 51–0, and the
J51–0 pure rotational transition in the v 51 vibration pre-
dicted for each of the model potentials. These predictions
vary widely for the different models, even those which agree
well for the present ground-state microwave data. More
spectroscopic data would be invaluable for better specifying
the potential. Direct submillimeter-wave vibrational excita-
tion of Ne–Ar is possible, though vibrational transition mo-
ments are generally smaller than permanent moments where
the latter do not vanish identically due to symmetry. In this
case, however, ^c1um(R)uc0& may be relatively large since
m(R) varies so rapidly with R.18 Pure rotational transitions in
the excited vibrational states would be expected to have
smaller moments than the ground state since the vibra-
tionally averaged separation would be larger. Also, in typical
molecular beams or jets, any ‘‘hot’’ vibrational population is
effectively relaxed. Probably, laser-induced-fluorescence
methods in the vacuum ultraviolet, as accomplished with
FIG. 3. Comparison of model potentials for Ne–Ar: ~a! well region; ~b! high resolution in Ar2 ,12 could provide useful data, as might
repulsive wall.
Raman scattering.

stants. The R e for all the fitted potentials fall in a narrow V. CONCLUSIONS
range of 348.0~2! pm, even in the highly correlated cases. If
we leave out the unrealistic Lennard-Jones and Morse poten- Direct fitting of simple model potentials to spectroscopic
tials, then the range is even smaller, R e 5348.05~7! pm. measurements on diatomics can provide a very reliable
Therefore, we believe that this equilibrium bond length is means of predicting nearby transitions and isotopic varia-
more reliable than that obtained from the substitution struc- tions. The extreme accuracy of microwave data at low J can
tures given in Table I, which is systematically lower by determine the equilibrium bond length at the potential mini-
'0.5%. In this very anharmonic molecule, R e is '5.5% mum with more consistency than using rotational constants
lower than R 0 for the various isotopomers. and isotopic substitution methods. Potentials constructed to
We have not incorporated any nonspectroscopic fit the best available thermodynamic, transport and scattering
data21–26 into these potentials, as Barrow and Aziz10 did for data require adjustments of a few percent to fit the micro-
the HFD-B; so we compare the potentials graphically in Fig. wave measurements. A wider range of vibrational and rota-
3. In the well region shown in Fig. 3~a!, it is difficult to tional data is required to better specify the well depths, re-

J. Chem. Phys., Vol. 102, No. 3, 15 January 1995


Downloaded¬22¬Sep¬2009¬to¬129.6.168.96.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright;¬see¬http://jcp.aip.org/jcp/copyright.jsp
Grabow et al.: Rotational spectra of Ne–Ar 1187

9
pulsive walls and long-range forces, and nonspectroscopic R. A. Aziz, in Inert Gases, edited by M. L. Klein, Series in Chemical
data need to be incorporated with appropriate weighting. Physics, Vol. 34 ~Springer, Berlin, 1984!, Chap. 2.
10
D. A. Barrow and R. A. Aziz, J. Chem. Phys. 89, 6189 ~1988!.
The high sensitivity of FTMW spectroscopy, demon- 11
E. A. Colburn and A. E. Douglas, J. Chem. Phys. 65, 1741 ~1976!.
strated here for the weakly polar Ne–Ar dimer, suggests that 12
P. R. Herman, P. E. LaRocque, and B. P. Stoicheff, J. Chem. Phys. 89,
the even more weakly bound He–rare gas heterodimers, 4535 ~1988!.
13
which are expected to have rotational transitions in the range R. D. Suenram, F. J. Lovas, G. T. Fraser, J. Z. Gillies, and C. W. Gillies,
of these instruments,15 may be observable. Also, pure rota- J. Mol Spectrosc. 137, 127 ~1989!.
14
C. Chuang, C. J. Hawley, T. Emilsson, and H. S. Gutowsky, Rev. Sci.
tional transitions for highly symmetric molecules with cen- Instrum. 61, 1629 ~1990!.
trifugally induced moments @e.g., CF4 or ~HF!3# or isotopi- 15
J. F. Ogilvie and F. Y. H. Wang, J. Mol. Struct. 291, 313 ~1993!.
cally induced moments beyond H/D substitution may be 16
A. H. Wapstra and K. Bos, At. Data Nucl. Data Tables 19, 175 ~1977!.
within reach.
17
C. H. Townes and A. L. Schawlow, Microwave Spectroscopy ~McGraw-
Hill, New York, 1955!.
18
G. Birnbaum, M. Krauss, and L. Frommhold, J. Chem. Phys. 80, 2669
ACKNOWLEDGMENTS ~1984!.
19
A. Kumar and W. J. Meath, Mol. Phys. 54, 823 ~1985!.
J.-U. G. was supported by a Post-Doctoral Fellowship 20
J. M. Standard and P. R. Certain, J. Chem. Phys. 83, 3002 ~1985!.
from the Deutsche Forschungsgemeinschaft. The work at Il- 21
J. J. H. van der Biesen, R. M. Hermans, and C. J. N. van der Meijdenberg,
linois was supported in part by the National Science Foun- Physica A 115, 396 ~1982!.
22
dation under Grant No. CHE 91-17199 and by the donors of L. Beneventi, P. Casavecchia, and G. G. Volpi, J. Chem. Phys. 84, 4828
~1986!.
the Petroleum Research Fund, administered by the American 23
P. K. Pol, cited in Ref. 9.
Chemical Society. 24
O. Shamma and M. Rigby, J. Chem. Soc. Faraday Trans. 278 689 ~1982!.
25
P. S. Arora, H. L. Robjohns, and P. J. Dunlop, Physica A 95, 561 ~1979!.
1
W. Jäger, X. Xu, and M. C. L. Gerry, J. Chem. Phys. 99, 919 ~1993!.
26
B. Najafi, E. A. Mason, and J. Kestin, Physica A 119, 387 ~1983!.
2
T. J. Balle and W. H. Flygare, Rev. Sci. Instrum. 52, 33 ~1981!.
27
J. W. Cooley, Math. Comput. 15, 363 ~1961!.
28
3
J.-U. Grabow and W. Stahl, Z. Naturforsch. Teil A 45, 1043 ~1990!; J.-U. R. J. LeRoy, Chem. Phys. Res. Rept. CP-230R2, University of Waterloo,
Grabow, PhD. thesis, Kiel University, Germany, 1992. 1985.
4
C. C. Costain, J. Chem. Phys. 29, 864 ~1958!.
29
J. F. Ogilvie, J. Chem. Phys. 88, 2804 ~1988!.
5
J. M. Steed, T. A. Dixon, and W. Klemperer, J. Chem. Phys. 70, 4095
30
G. C. Maitland and E. B. Smith, Chem. Phys. Lett. 22, 443 ~1973!.
31
~1979!. P. G. Hajigeorgiou and R. J. Leroy, 49th International Symposium on
6
W. Jäger, Y. Xu, J. Djauhari and M. C. L. Gerry, 48th International Sym- Molecular Spectroscopy, Columbus OH, WE804 ~1994!.
posium on Molecular Spectroscopy, Columbus, OH, RG804 ~1993!. 32
J.-L. Femenias, J. Mol. Spectrosc. 144, 212 ~1990!.
7
Y. Xu, W. Jäger, and M. C. L. Gerry, J. Chem. Phys. 100, 4171 ~1994!. 33
J.-U. Grabow, N. Heineking, and W. Stahl, J. Mol. Spectrosc. 152, 168
8
R. Ahlrichs, R. Penco, and G. Scoles, Chem. Phys. 19, 119 ~1977!. ~1992!.

J. Chem. Phys., Vol. 102, No. 3, 15 January 1995


Downloaded¬22¬Sep¬2009¬to¬129.6.168.96.¬Redistribution¬subject¬to¬AIP¬license¬or¬copyright;¬see¬http://jcp.aip.org/jcp/copyright.jsp
View publication stats

You might also like