Sunteți pe pagina 1din 48

6 Infrared emission spectroscopy

P. F. Bernath
Department of Chemistry, University of Waterloo, Waterloo, ON, Canada
N2L 3G1
1 Introduction
Infrared spectroscopy has been traditionally carried out mainly in absorption. The
virtues of infrared emission spectroscopy have been largely overlooked, but there
has been a recent surge of interest. The modern infrared Fourier transform
spectrometer has made emission measurements much easier. This article is an
attempt to provide a comprehensive view of the technique and builds on a shorter
Chemical Society review.
1
Previous reviews of infrared emission spectroscopy have concentrated on the basic
principles and on applications in analytical chemistry.
2^6
We shall cover some of this
ground as well, but the main focus will be in applications in high resolution molecular
spectroscopy. We shall not cover the infrared spectra of atoms although excellent
infrared emission spectra have been recorded at Kitt Peak, Orsay and Lund. This
review has some overlap with reviews on the infrared spectra of transient molecules,
including ions, free radicals and high temperature molecules,
7^10
as well as the
spectroscopic reviews of Barrow and Crozet.
11
For the purposes of this review, ``infrared'' is arbitrarily dened as
10^10 000 cm
1
(1^1000 mm), to include the far-infrared, mid-infrared and
near-infrared region. The microwave and submillimeter regions are excluded,
although laboratory microwave emission spectroscopy has become very popular
with the proliferation of Flygare^Balle spectrometers.
12
These instruments detect
the coherent emission of microwave radiation. At slightly higher frequencies radio
astronomers detect molecules in molecular clouds by microwave and millimeter wave
emission spectroscopy. The eld of astrochemistry
13
is built almost entirely on
centimeter wave and millimeter wave emission spectroscopy of species in various
astronomical objects ranging from stellar envelopes to comets.
We shall focus our discussion mainly on emission from gases, although heated
solids
14
and liquids
15
give useful spectra. The secret to recording useful spectra from
condensed phases is to work with thin lms or dispersed particles. For thick samples,
multiple scattering of infrared photons results in a nearly featureless blackbody
spectrum that depends only on the temperature of the emitter.
DOI: 10.1039/b001200i Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
177
This review will cover mainly high resolution spectroscopy. In this context, ``high
resolution'' means that the rotational structure is at least partly resolved for a
gas phase sample. The emission technique also works equally well for large molecules
such as C
60
and C
70
,
16
and for species ranging from DNA bases
17
to polycyclic aro-
matic hydrocarbons,
18
in which the rotational structure is not resolved. For these
molecules the solids are heated to about 200^300

C and emission from the vapour
provides excellent spectra even in the far-infrared region.
18
The attraction of emission spectroscopy is the possibility of an improved
signal-to-noise ratio compared to absorption spectroscopy. Ideally only photons
emitted by the sample are detected (``zero background''), free from the noise pro-
duced by the continuum lamp in an absorption experiment. This improvement in
sensitivity is particularly useful for the spectroscopy of transient molecules because
of their intrinsically low concentrations. This potential advantage of emission
spectroscopy for the detection of ions and free radicals is well known is the visible
and near-UV regions. For example, the violet emission from the CH radical (A
2
DX
2
P) is readily seen by eye in a ame
19
but the measurement of the A
2
DX
2
P absorption
20
is much more difcult. This emission advantage persists into
the infrared region.
The infrared region is unique because all molecules, with the exception of
homonuclear diatomics, have at least one allowed vibration^rotation transition.
While it is true that there are infrared electronic transitions and a few light molecules
have far-infrared rotational transitions, infrared spectroscopy is nearly synonymous
with vibrational spectroscopy. Moreover, vibrational spectroscopy through the con-
cept of group frequencies also provides chemical information in a way that rotational
and electronic transitions do not.
Even the weak electric quadrupole emission transitions of H
2
have been seen by
astronomers from molecular clouds experiencing shock waves.
21
Infrared emission
spectroscopy has the potential to be a sensitive, universal, molecule-specic monitor
of chemical composition.
This review will be organized by sources, ranging from stars to microwave dis-
charges, in which emission spectroscopy has been carried out. The coverage of
the more recent high resolution spectroscopic work aims to be relatively complete.
Older work and various applications in surface science, astronomy, chemical
dynamics or analytical chemistry are more illustrative than complete. But rst a
few basic principles and instrumental considerations will be discussed.
2 Basic principles
Emission spectroscopy is based on a few basic equations and principles. The rst is
the interaction of monochromatic radiation with a sample (Fig. 1) (ignoring
scattering and uorescence). The beam can be reected, absorbed or transmitted, so
a r t = 1 (1)
in which a is the absorptance (absorption factor), r is the reectance (reection
factor) and t is the transmittance (transmission factor) of a body.
5,22
Note that
178
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
the difference between a reectance (``-ance'') and a reectivity (``-ivity'') is that the
latter applies under some set of standard conditions (e.g., smooth surface, thick
sample) while the former applies to a specic sample. Because we are generally dis-
cussing specic objects, we will use the ``-ance'' terms. Factors a, r and t are wave-
length dependent numbers between 0 and 1 and there are three simple limits:
a = 1Y r = t = 0 loi a LlacLLooy (2)
r = 1Y a = t = 0 loi a eilecl miiioi (3)
t = 1Y a = r = 0 loi a eilecl winoow (4)
Kirchhoff's law states that the emittance e of a sample is equal to the absorptance a
(see above). The monochromatic emittance of a sample is dened as:
L = eL
BB
(5)
in which L is the radiance (or ``sterance'', in units of watts per steradian per square
meter of source per hertz of spectral bandwidth)
22
of a sample, and L
BB
is the
radiance of a blackbody. Thus L
BB
is related to the Planck function and e is a
proportionality constant between 0 and 1 that converts it into the observed radiance,
L. The expression for the radiance of a blackbody is:
L
BB
=
2hn
3
c
2
(e
hnakT
1)
(6)
in units of W sr
1
m
2
s. Because the emittance and absorptance are equal,
e = a (7)
all of the selection rules and optical properties of any material (gas, liquid or solid)
that are generally dened in terms of absorption transfer directly to emission.
All of the above equations depend on both frequency (or wavenumber) and tem-
Fig. 1 Light striking an object will be reected, absorbed or transmitted (ignoring uorescence
and scattering).
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
179
perature so Kirchhoff's law can be written as
L( nY T) = a( nY T)L
BB
( nY T) (8)
in which n is the wavenumber in customary non-SI units of cm
1
.
Kirchhoff's law has been tested experimentally using a hot CO
2
gas sample.
23
The
absorptance was measured from the transmittance [e.g., eqn. (1) with r = 0],
a = 1 t (9)
and the emittance measured by comparison with a blackbody source at the same
temperature,
e =
L
L
BB
(10)
Direct comparison for different temperatures and concentrations of CO
2
in N
2
showed that a = e.
The direct application of Kirchhoff's law allows the determination of sample
temperature
24,25
as illustrated in Fig. 2. This gure also illustrates the difference
between gas phase emission and absorption measurements. In the upper panel,
the absorption of a mixture of acetylene, butane and carbon dioxide at 555 K is
displayed. In the middle panel, the measured radiance of the sample is plotted.
At the bottom, the ``normalized radiance'', dened as the radiance divided by
absorptance, is plotted. By Kirchhoff's law this ratio is the blackbody radiance, i.e.,
L
a
=
L
(1 t)
= L
BB
(T) (11)
The radiance of a blackbody at a certain frequency is a function of only the
temperature, which can be adjusted until the calculated Planck function matches
the observed ``normalized radiance''. The shaded regions mark regions that contain
no molecular emission (i.e., no information) and are ignored. The message of this
gure is clear: a thermal emission spectrum is nothing more than the ``inverted''
absorption spectrum modulated by the Planck function.
The quantitative interpretation of sample emission is, in principle, as simple as
that of absorption. In practice, however, samples rarely have uniform temperatures
and ``self-absorption'' is a problem. Fortunately sample emission increases strongly
with temperature. For example, the total radiance (integrated over all frequencies)
of a blackbody is proportional to the fourth power of temperature:
L
lolal
=

o
0
L
BB
on = sT
4
apY (12)
in which s is the Stefan^Boltzmann constant. (The factor of p appears because the
emission of the blackbody, eqn. (6), is per steradian and the normal version of
the equation is simply the total emission from a hole.) This means that the highest
temperature object in the eld of view of the spectrometer dominates the appearance
of the spectrum. The general equation for the observed radiance at a given frequency
180
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
(Fig. 3) for the typical case of a gas sample in front of a wall
26
is
L(0) = L(z
0
)t(z
0
)

t(z
0
)
0
L
BB
(T
z
) ot (13)
in which L(z
0
) is the radiance of an object at z
0
away from the observer located at 0.
Note that Kirchhoff's law gives de = dt. The transmittance t(z
0
) attenuates this
Fig. 2 Spectra for a mixture of acetylene, carbon dioxide and butane in helium at 555 K: (a)
absorption in %; (b) radiance; (c) normalized radiance, with a blackbody t for L
BB
at 565 K,
ignoring the shaded regions.
25
Reproduced from ref. 25, with permission.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
181
radiance from z
0
and the integral accounts for the radiance of the intervening
radiating elements. The transmittance is given by Beer's law,
t(z) = ex

z
0
k(z
/
) oz
/

(14)
in which k is the absorption coefcient. If the sample is uniform and the temperature
is constant then eqn. (14) becomes,
t(z) = e
kz
(15)
and the general equation, eqn. (13), reduces to:
L(0) = L(z
0
) e
kz
0
L
BB
(1 e
kz
0
) = L(z
0
)t(z
0
) L
BB
[1 t(z
0
)] (16)
This equation has two simple limits: the optically thin case when t & 1 and L &
L(z
0
), and the optically thick case when t & 0 and L=L
BB
. In the optically thin
case the radiance is that of the wall, while in the optically thick case only the front
of the gas sample is seen.
The discussion presented so far is on the macroscopic level. The connection to the
microscopic world of atoms and molecules is through the Einstein equations; in par-
ticular, the Einstein A coefcient for emission from level 2 to 1 is:
A
21
=
16p
3
n
3
[m
21
[
2
3he
0
c
3
(17)
in which m
21
is the transition dipole moment, and e
0
is the permittivity of free space.
The Einstein A coefcient measures the rate of photon emission from the excited
level of an atom or molecule,
oN
2
ot
= A
21
N
2
(18)
and has units of s
1
. The above expression has been integrated over the lineshape of
Fig. 3 Radiance, L, seen by an observer at 0 looking through a gas sample to a back wall
[radiance L(z
0
)] at z
0
.
182
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
the transition and the more detailed expression
27
is
(A
21
)
n
=
16p
3
n
3
[m
21
[
2
3he
0
c
3
g(n n
0
) (19)
in which g(n n
0
) is the normalized lineshape function. The strong cubic frequency
dependence of the emission rate means that emission work in the infrared and
far-infrared is more difcult than in the visible and ultraviolet. Moreover, transition
dipole moments for vibration^rotation transitions tend to be smaller than typical
values for allowed electronic transitions or typical values of permanent dipole
moments for pure rotational transitions.
3 Methodology
The basic requirement for emission spectroscopy is a source of radiation and a
detector. For some experiments, a simple infrared lter between the source and
detector can provide useful results. For example, Chang and Klemperer
28
excited
the (HF)
2
molecule in a jet expansion with a tuneable near-infrared laser and moni-
tored the HF infrared emission. This type of action spectroscopy obtained by
scanning a laser and detecting the total emission is common at shorter wavelengths
but also works in the infrared. The simple but effective technique of non-dispersive
infrared absorption spectroscopy (NDIR) has an emission analog called ame
infrared emission (FIRE) spectrometry.
29
FIRE is a simple lter^detector combi-
nation that can be used as a detector in a gas chromatograph.
30
In this analytical
application the hydrocarbon analyte from a gas chromatograph is burnt in a
hydrogen^oxygen ame and hot CO
2
emission is detected.
29,30
The basic requirement for any emission experiment is that the source and detector
have different temperatures. The detector is also an emitter of radiation, and if the
source and detector have the same temperature then there is no net ux of radiation
falling on the detector. Obviously, a room temperature detector such as triglycine
sulfate (TGS) cannot be used to monitor a room temperature sample. But a
liquid-nitrogen-cooled InSb detector can certainly see a room temperature sample.
Generally emission from a detector at 4 K (liquid He) or even 77 K (liquid N
2
)
can be ignored.
A slightly more sophisticated system for emission spectroscopy is based on the use
of a circular or linear variable lter between the source and detector. By rotating the
circular lter the peak transmission of the lter can be changed. Such lters can be
purchased from, for example, Optical Coating Laboratory, Inc. (OCLI) of Santa
Rosa, CA. The use of a circular variable lter results in a simple, compact
spectrometer but with a low resolving power,
R =
l
Dl
=
n
D n
(20)
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
183
of typically less than 100. Circular variable lters are still used in astronomy and for
simple analytical applications in which their high optical throughput and relatively
low cost are an advantage.
Early high resolution infrared emission measurements were made using classical
grating spectrographs or spectrometers to disperse the emission. For example,
the Ballik^Ramsay bands of the C
2
molecule were discovered
31
in 1963 using an
infrared spectrometer constructed by Douglas and Sharma
32
at the National
Research Council of Canada. The C
2
molecule was made by the evaporation of
graphite from the walls of a carbon tube furnace (King furnace) operated at about
2900

C. The emission was detected using a dry-ice-cooled PbS detector. The
vibronic bands of the A
3
S

g
^X
3
P
u
electronic transition were seen at
3800^7100 cm
1
.
The performance of any spectrometer can be improved in the thermal infrared
region (n d3000 cm
1
) by cooling the entire instrument. Heroic early experiments
were carried out by McDonald and co-workers
33
to study the nascent products
of reactions of F atoms with hydrocarbons such as ethylene. The chemiluminescence
emitted by these free radical reactions at low pressures was very weak.
A modern version of a cooled spectrometer was developed originally by Pimentel
and then used by Saykally and co-workers
34
to study the weak emission from
laser-excited polycyclic aromatic hydrocarbons (PAHs) in a supersonic free jet
expansion. In this case the entire spectrometer was cooled to 4 K by liquid helium
in order to take advantage of a new ultrasensitive blocked impurity band detector.
This detector from Rockwell is based on Si:As and has an internal avalanche process
that results in gain. In other words, it is a solid state photomultiplier that operates in
the infrared.
An even more sophisticated instrument can be made using modern infrared array
detectors. The use of a large format array (typically InSb or HgCdTe) with a
spectrograph is attractive because of the multiplex advantage. The most sensitive
infrared instruments are the cryogenic echelle spectrographs that are in use or under
construction at all major observatories. One example is Phoenix (Fig. 4) at Kitt Peak
National Observatory in Tucson, Arizona.
35
Phoenix is cooled to 50 K to eliminate the thermal emission from the spectrograph
and uses an echelle grating in high order to obtain high resolution in a compact
instrument. Order sorting is carried out with cooled infrared lters. Phoenix cur-
rently has a resolving power of about 70 000 or a resolution of 0.03 cm
1
at
2000 cm
1
. The use of a 1024 1024 InSb array allows coverage of the
1800^10 000 cm
1
region.
Phoenix is calculated to have a sensitivity advantage of nearly 100 over a con-
ventional Fourier transform spectrometer. This high sensitivity originates from sev-
eral factors including a very restricted spectral bandpass, cryogenic cooling and
improved detector performance. The most important factor is that infrared array
detectors are fundamentally different from conventional single element detectors.
Because infrared arrays are integrating detectors, very low light levels can be
handled. If the main noise source is read-out noise then the signal-to-noise ratio
grows linearly with time. This is in contrast to a conventional single element detector
184
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
where the signal-to-noise ratio grows with the square root of time. The performance
of Phoenix was tested by recording emission spectra of the NH free radical using a
microwave discharge source.
36
Most infrared emission measurements are made with Fourier transform
spectrometers (FTSs). A typical experimental arrangement
37
is illustrated in Fig.
5. Light is admitted through the ``emission port'', which is located near the internal
sources used for absorption spectroscopy. Fourier transform spectroscopy has
become ``conventional'' so no detailed descriptions are needed.
Infrared emission spectroscopy requires simply that the internal glowbar source be
replaced by the source of interest. The Fourier transform interferometer is best
viewed as a modulator that shifts infrared frequencies into audio frequencies at
Fig. 4 Phoenix cryogenic echelle spectrograph of Hinkle et al.
35
at Kitt Peak National
Observatory, Tucson, AZ, USA.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
185
the detector. For example, the 0^10 000 cm
1
spectral region is mapped into the
0^10 000 Hz frequency range on the detector if the moving mirror changes the
optical path difference at 1 cm s
1
. Unusual infrared emission measurements are
sometimes made by using a ``second'' modulation of the source.
The most common of the double modulation experiments is time-resolved Fourier
transform spectroscopy (TRFTS).
38,39
A conceptually simple approach to TRFTS is
the step-scan method using a periodic infrared source. If the ``moving'' mirror of an
FTS is stopped then the infrared signal can be recorded as a function of time.
The moving mirror is then sent to the next sampling position and another infrared
emission decay is recorded. By accumulating a set of emission decays, each recorded
at the usual sampling position of the interferogram, a set of time-resolved spectra are
obtained.
Fig. 5 Typical long wavelength infrared chemiluminescence emission experiment using a
Fourier transform spectrometer equipped with a copper-doped Ge detector.
37
Reproduced
from ref. 37, with permission.
186
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
Continuously scanning FTSs can also record time-resolved infrared spectra. In
this case, the fringes of the internal He^Ne laser are used (with appropriate time
delays) to trigger the infrared emission source and the sampling of the emission
signal. In this way, an interferogram at a specic time delay (or series of delays)
from the excitation pulse is recorded.
Time-resolved infrared emission spectra are very useful in the study of reaction
dynamics and collisional relaxation. In Fig. 6 the time-resolved OH emission from
the O HCl reaction is displayed.
38
As time progresses the highly excited OH mol-
ecules produced by the chemical reaction are relaxed by collisions. The development
of schemes for various time-resolved FTS work continues with, for example, the
recent development of the ``event-locked'' method by Weidner and Peale.
40
Double modulation techniques are attractive in the emission spectroscopy of tran-
sient molecules because they discriminate against the more abundant precursor
molecules. The selectivity of velocity,
41,42
concentration
43
and Zeeman
modulation
44,45
have all been demonstrated with FTSs. Unfortunately they generally
do not also offer an increase in sensitivity and work most easily with step-scan instru-
ments so they have not been widely adopted for high resolution work.
An important consideration in emission work is the ``contrast'' between the sample
and background. This is particularly important in remote sensing of gases and in the
far-infrared region. If the eld-of-view of the spectrometer includes a gaseous sample
against a background at a certain temperature, then the lines of the sample will
appear in absorption or emission depending on the sample temperature. If the sample
is warmer than the background then the lines will appear in emission, while if it is
cooler then the lines will appear in absorption. If the sample and background have
the same temperature then the sample lines will disappear, i.e., there is no contrast.
This spectral line reversal technique has been used to determine ame
temperature.
19
A ame is viewed against the continuum of a lamp and a metal atom
such as Na is added to give a bright emission line. As the lamp lament is increased
in temperature from below the ame temperature, the atomic emission line will dis-
appear when the ame and lament temperatures are equal. Filament temperatures
are then easily determined with an optical pyrometer.
Fig. 6 The time-resolved Fourier transform emission spectrum of the reaction of O with HCl.
The highly excited OH molecules are quenched by collisions with HCl and this causes the
OH vibration^rotation line intensities to decrease with time.
38
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
187
The problem of contrast is particularly severe in the far-infrared region and this
causes problems for both absorption and emission measurements. In the far-infrared
sources operate usually in the Rayleigh^Jeans limit with hn `` kT, so the blackbody
radiance reduces to
46
L
BB
=
2n
2
kT
c
2
=
2kT
l
2
X (21)
Because the radiance is only linearly proportional to temperature, the contrast
between a heated sample and the surroundings is much less than in the visible.
Indeed, usually the spectrometer itself contributes a major portion of the ``signal''
in a far-infrared emission experiment.
4 Chemiluminescence
To a chemist and spectroscopist, the light emitted from a chemical reaction has an
undeniable fascination. Unfortunately chemiluminescence tends to be relatively
weak and not very common. Chemiluminescence is closely related to the
spectroscopy of ames and to the excitation of molecules by energy transfer from
metastable species. These sources of infrared emission will be discussed separately
below.
A Nobel prize has been awarded to J. Polanyi (shared with Y. T. Lee and D.
Herschbach) for the study of reaction dynamics by infrared chemiluminescence.
47
This work started in 1958 with the observation of HCl emission from the
H Cl
2
HCl Cl (22)
reaction by Cashion and Polanyi.
48
A commercial infrared spectrometer with an
NaCl prism and a thermocouple detector was used. A similar chemiluminescent
reaction (H
2
Cl
2
ame) was later used by Clayton et al.
49
at Penn State with
a high resolution spectrometer. Infrared chemiluminescence in the style of Polanyi
is still carried in a number of groups such as those of Setser and Leone. Butkovskaya
and Setser
50
studied the dynamics of the
OH HBi H
2
O Bi (23)
reaction, for example, while Klaassen et al.
51
recorded the rst high resolution
spectrum of HOI from the
C
2
H
5
! O HO! C
2
H
4
(24)
chemical reaction.
Much of the chemical dynamics work is now carried out by time-resolved Fourier
transform spectroscopy. Recent examples include the work of the Sloan laboratory
at 1 ms time resolution for reactions of H atoms with uorochlorocarbons.
52
With
a step-scan FTS system even 10 ns time resolution is possible.
53^55
188
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
Chemiluminescence has also proved to be very useful in studies devoted to new
spectroscopy (rather than dynamics) of molecules. At long wavelengths, a remark-
able high resolution spectrum of FO was recorded by Hammer et al.
37
from the
energetic
! O
3
!O O
2
(25)
reaction. The F atoms were made in a microwave discharge of F
2
(Fig. 7). FO emi-
ssion up to the 87 vibration^rotation band could be assigned near 1000 cm
1
(Fig.
8). The analogous OH (or OD) reaction
56^58
H O
3
HO O
2
(26)
gives an exceptionally ne infrared emission spectrum with OH populated up to u =
9. Interestingly this same reaction (26) is responsible for atmospheric nightglow (see
below) and astronomers have detected emission from u = 10.
58
Infrared electronic emissions are also possible and the work of E. Fink and
co-workers is particularly noteworthy. Most ``normal'' stable main group molecules
do not have low-lying electronic states, with the exception of O
2
and NO: O
2
has a p
2
conguration that leads to a X
3
S

g
ground state with a
1
D
g
and b
1
S

g
states at 7882
and 13121 cm
1
, respectively; NO has a regular X
2
P ground state with two spin
components,
2
P
1/2
and
2
P
3/2
, split by 123 cm
1
. Closed-shell main group molecules
Fig. 7 Chemiluminescence reactor used in the F O
3
infrared emission experiment to make
FO.
37
The F atoms are created in a microwave discharge of F
2
in He. Reproduced from ref.
37, with permission.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
189
generally have excited electronic states that give rise to UV spectra. By virtue of their
unpaired electrons, free radicals often have low-lying electronic states. Most free
radicals are very reactive molecules but there are a few ``stable'' free radicals such
as O
2
and NO.
Fink and co-workers have recorded the infrared electronic spectra of a number of
forbidden transitions of free radicals with the same p
1
or p
2
congurations as
NO or O
2
. The NH, PH, AsH, SbH and BiH family have p
2
congurations and
Beutel et al. have measured the a
1
D^X
3
S

infrared transitions of PH,


59
AsH
60
and SbH.
61
These molecules were made by the reaction of H atoms with heated
elemental solids. [BiH was also made in a similar fashion but emission was excited
by energy transfer from O
2
(
1
D), see below.] Although the a
1
D^X
3
S

transition
is forbidden by normal electric dipole selection rules, it is allowed by the magnetic
dipole transition moment. Excellent high resolution spectra were obtained with a
high purity Ge detector.
The TeF, TeCl, TeBr and TeI molecules are isovalent with OHand have inverted X
2
Pground states from a p
3
conguration. The spin^orbit coupling constants for TeF
and TeCl, however, are about 4000 cm
1
, as compared to 139 cm
1
for OH.
Ziebarth et al.
62
measured the magnetic dipole emission X
2
P
1/2
X
2
P
3/2
between
the two spin components for TeF and TeCl at high resolution. The excited TeF and
TeCl molecules were made by the reaction of a TeH/TeH
2
mixture with F
2
or Cl
2
.
Low resolution infrared spectra for TeF, TeCl, TeBr and TeI are also known.
63
Another interesting example from the Fink group is the infrared electronic
transitions of BiP, BiAs and BiSb.
64
These diatomics are isovalent with N
2
and have
X
1
S

ground states. Unlike N


2
, however, the
3
S

and
5
S

states that also arise from


Fig. 8 Chemiluminescent vibration^rotation emission from FO created in the F O
3
reaction.
37
The marks at the top indicate the vibrational band origins. Reproduced from ref.
37, with permission.
190
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
the lowest energy
4
S
4
S atomic asymptote are low-lying. The a
3
S

1
X
1
S

electronic transitions of BiP, BiAs and BiSb were detected in the 7000^10 000 cm
1
region through chemiluminescence from the reaction of P, As and Sb atoms with
Bi
x
vapour.
Very recently we have adapted the Broida oven ow reactor to study the infrared
chemiluminescence of the classic metal plus oxidizer reactions.
65
In particular,
the reactions
Ca
2
O CaO
2
(27)
Si
2
O SiO
2
(28)
yield excellent near-infrared electronic spectra. The A
1
S

X
1
S

and A
/
1
P X
1
S

transitions of SrO are displayed in (Fig. 9).


Fig. 9 Chemiluminescence from SrO from the Sr N
2
O reaction. The bands are mainly due to
the A
/ 1
P X
1
S

electronic transition and some isolated Sr atomic lines can also be seen.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
191
5 Excitation by energy transfer from metastables
Excitation can be delivered to a molecule by a chemical reaction or by energy transfer
from another excited molecule. Both of these processes occur in ames so that the
distinction that we have made between the three emission sources ^
chemiluminescence, excitation by energy transfer and ames ^ is somewhat articial.
Popular metastables include the rare gases, O
2
(
1
D), NF(
1
D) and active nitrogen. Of
the rare gases, He is particularly popular because the metastable 2
3
S state has
19.8 eV of available energy. Active nitrogen is made by passing N
2
gas through dis-
charge and is a complex mixture of N and N
2
in both ground and excited states.
O
2
(
1
D) and NF(
1
D) are unique in that both species can be made in relatively high
concentrations by purely chemical means. This is attractive for the pumping of
chemical lasers such as the COIL (chemical oxygen^iodine laser) system.
66
O
2
(
1
D)
atoms are produced by the reaction of Cl
2
gas with basic hydrogen peroxide,
Cl
2
(g) H
2
O
2
(l) 2KOH(l) O
2
(
1
D)(g) 2KCl(l) 2H
2
O (29)
Energy transfer to I
2
causes dissociation and lasing on the I
2
P
1/2

2
P
3/2
transition.
NF(
1
D) can be formed by the thermolysis of the explosive gas uorine azide
67
at
1000 K,
!
3
!(
1
D)
2
(30)
The chemical production of O
2
and NF metastables is clearly not for the faint
hearted! For laboratory purposes an electrical or microwave discharge is generally
used to make metastable atoms and molecules.
The COIL laser, in fact, was associated with a mysterious red and infrared emi-
ssion that turned out to be CuCl
2
. Traces of chlorine reacted with heated copper
tubing to produce CuCl
2
, which was excited by energy transfer from the O
2
(
1
D)
molecule.
68,69
Ultimately this spectroscopic mystery led to the high resolution analy-
sis of the linear CuCl
2
molecule.
70
Both visible and infrared electronic transitions of
CuCl
2
were analyzed using a variety of sources.
The group of Fink has made extensive use of the O
2
(
1
D) metastable to record
spectra of an amazing number of main group molecules. The main types of main
group free radicals that give rise to infrared electronic transitions have already been
mentioned in the section on chemiluminescence. For diatomic molecules they gen-
erally have p
1
(
2
P
r
), p
2
(
3
S

,
1
D,
1
S

) or p
3
(
2
P
i
) congurations as exemplied by
CH, NH and OH free radicals, respectively. In the CH family, the forbidden X
2
P
3/2
X
2
P
1/2
transitions
71,72
were detected for PbF, PbCl, PbBr and PbI. These mol-
ecules were all made by the reaction of Pb vapour with halogens and then excited
by O
2
(
1
D) metastables. The high quality of the spectra is illustrated with PbF (Fig.
10). For the NH family, the a
1
D X
3
S

transitions were analyzed for AsI,


73
SbF,
74
SbCl,
74
SbBr,
74
SbI,
74
BiCl,
75
BiBr
75
and BiI.
75
In addition the X
2
3
S

1
X
1
3
S

0
ne structure transition of BiH
76
(as well as for BiF,
77
BiCl,
78
BiBr
78
and BiI
78
) was also measured. The X
2
P
1/2
X
2
P
3/2
transition
79
of TeH and
TeD was seen near 4000 cm
1
.
192
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
The O
2
molecule has a similar energy level pattern to the NH family and has been
studied extensively by magnetic dipole and collision-induced
80
emission. The O
2
a
1
D
g
^X
3
S

g
transition occurs near 8000 cm
1
, so the
1
D metastable carries about
1 eV of energy. The O
2
b
1
S

g
^a
1
D
g
electric quadrupole transition
81
is seen near
5000 cm
1
(Noxon band
82
). The a
1
D^X
3
S

or b
1
S

^ X
3
S

infrared electronic
transitions were detected for the isovalent SO,
83
S
2
,
84,85
SeO,
86,87
SeS,
88
Se
2
,
89
TeO,
90
TeS,
90
TeSe
91,92
and Te
2
93
molecules, all excited by O
2
(
1
D).
The BiO molecule has a similar electronic structure to the isovalent NO. Extensive
infrared electronic transitions of BiO have been studied
94
including the X
2
P
3/2
X
2
P
1/2
transition
95,96
near 7000 cm
1
. This transition displays a remarkable hyperne
structure (Fig. 11). In the N
2
family, the bands analogous to the Vergard^Kaplan
system of N
2
are shifted from the near-UV region into the near-infrared for the
heavier members. The a
3
S

u
^X
1
S

g
transitions of Sb
2
97
and Bi
2
98
are found near
9000 and 5000 cm
1
, respectively. Both ground and excited states correlate to
ground state N(
4
S) atoms, and there is a large change in bond length and vibrational
frequency. Analogous transitions have been seen for BiN,
99
BiP
100
and BiAs.
100
All
of the above emission spectra were induced by energy transfer from O
2
metastables.
The best known main group polyatomic that has an infrared electronic spectrum is
HO
2
. The bent hydroperoxyl free radical has the A
2
A
/
^

X
2
A
//
electronic transition
near 7000 cm
1
. Early high resolution emission measurements were made by Tuckett
et al.
101
using a SISAM spectrometer. The denitive analysis, however, is the recent
work of Fink and Ramsay.
102
Analogous transitions of HS
2
,
103
HSe
2
and HTe
2
79
have also been seen by Fink and co-workers.
Another new free radical, BiOH, with an infrared electronic spectrum near
6000 cm
1
was discovered by Fink et al.
104
BiOH forms when H
2
O is added to
the reaction of Bi vapour and metastable O
2
. BiOH is similar to BiF (rather than
Fig. 10 Fourier transform emission of the X
2
P
3/2
X
2
P
1/2
transition of PbF.
72
Reproduced
from ref. 72, with permission.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
193
the isovalent HNO molecule) in electronic structure. BiOH, however, is bent so the X
3
S

state of BiF splits into



X
1
1
A
/
,

X
2
1
A
//
and

X
3
1
A
/
electronic states. The observed
infrared electronic transitions thus correspond to the ``ne'' structure transitions

X
2


X
1
and

X
3


X
1
.
Fig. 11 High resolution emission spectrum of the X
2
2
P
3/2
X
1
2
P
1/2
transition of BiO. The
lower panel displays the impressive Bi hyperne structure in the rotational lines.
94
Reproduced
with permission, from ref. 94.
194
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
Active nitrogen also gives interesting new spectra. For example, when Vilesov et
al.
105
added Xe to a nitrogen discharge, a new band appeared very near to the
forbidden
2
P
o

2
D
o
emission of the atom near 9600 cm
1
(the
2
P
o
,
2
D
o
and
4
S
o
states arise from the 2p
3
conguration of N). They assigned the structure to
a bound^bound transition of the Xe:N excimer molecule.
105
Active nitrogen can
also be used to excite the vibration^rotation emission of stable molecules
106
such
as CO
2
.
He metastables are commonly used to create ions by Penning ionization. For
example, the A
2
P X
2
S

near-infrared electronic transition of CS

was measured
by Horani and Vervloet.
107
One of the main problems with using He metastables in a
owing afterglow is that their concentration is low and thus the emission of the
product ions or molecules tends to be weak. A solution to this difculty was found
by Vervloet, who adopted the Engelking supersonic corona discharge
108
(see below)
as the source of He metastables rather than the customary DC discharge. The
Engelking source is a prolic generator of metastable He atoms and provided a much
brighter source of CS

than usual in a owing afterglow. This same technique was


applied by Huber and Vervloet
109
to record the emission spectrum of the b
3
P
a
3
S

transition of NO

near 6000 cm
1
.
Another powerful source of rare gas metastables is the novel apparatus originally
built by Cossart
110
(Fig. 12). This source is often called the ``Cossart source'',
but this name is not unambiguous because Cossart has created a number of novel
Fig. 12 Cossart source
110
for the Penning excitation of N
2
(in this case) with metastable rare
gases. Reproduced from ref. 110b, with permission.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
195
and useful sources of molecules. For example, the Cossart source was one of the
sources used by Dabrowski et al. to record infrared emission experiments of
ArH,
111^114
ArD, KrH
115^117
and KrD, XeH
118
and XeD. These rare gas hydrides
are examples of what Herzberg has called ``Rydberg molecules'' as distinct from
Rydberg states in molecules.
119
Rydberg molecules have no stable ground states
and are based on putting an electron in a Rydberg orbital of a bound ion core.
For example, ArH has no chemically bound ground state but ArH

is a deeply bound
ion. By placing an electron in a Rydberg orbital built on the ArH

ion core, a series


of bound excited states are obtained. The excimer molecule ArF (of laser fame)
is another example of a Rydberg molecule. Rydberg^Rydberg transitions of
ArH are readily detected in the infrared region. Indeed, ArH is so easy to make
even as an impurity that we generally use Ne gas for infrared emission work with
a hollow cathode lamp. The rare gas hydride molecules were made in the Cossart
source by reacting H
2
(or D
2
) with a ow of metastable rare gases.
110^118
6 Engelking corona-excited supersonic jet source
Droege and Engelking
108
developed a simple but remarkably effective source for the
emission spectroscopy of jet-cooled free radicals (Fig. 13). The source is particularly
effective for the production of small non-metal free radicals. The corona excitation
occurs on the high pressure side of the nozzle and the resulting plasma is rapidly
expanded into vacuum. The high voltage, low current corona discharge is an efcient
source of electronically excited free radicals, which are rotationally cooled by jet
expansion. The main difculties with the source are that large molecules are not
always cooled and that the source is prone to various oscillatory electrical modes.
108
Fig. 13 Schematic of the tip of the Engelking corona-excited supersonic jet expansion source.
The vacuum pump, typically a Roots blower, is grounded to complete the circuit.
108
Reproduced from ref. 108b, with permission.
196
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
Vervloet has adapted this source for infrared emission spectroscopy of
electronically excited N
2
and NO. Vervloet and co-workers
120,121
studied the a
1
P
g
^a
/
1
S

u
, w
1
D
u
^a
1
P
g
and C
//
5
P
u
^A
/
5
S

g
(Herman) infrared bands of N
2
.
The b
4
S

a
4
P system of NO was analyzed by Huber and Vervloet.
122
High-l
Rydberg transitions (e.g., 5g^4f) of NO were also reported.
123,124
A completely unrelated method of recording emission spectra of cold molecules is
to irradiate liquid and gaseous He with a proton beam. Tokaryk et al.
125
saw
near-infrared emission from the d
3
S

u
^c
3
S

g
electronic transition of He
2
. He
2
is another example of a Rydberg molecule.
7 Magnetically conned plasma (Penning) sources
The application of a magnetic eld to a DC electrical discharge forces the ions and
electrons into circular orbits. The plasma is thus conned to a smaller volume
and a brighter source results. The use of a magnetic eld in this way was rst
advocated by Penning and Penning-type pressure gauges based on this principle
are used still. The magnetic eld allows the Penning source to operate over a very
wide pressure range. Cossart has used a number of Penning-type sources
126^128
for emission spectroscopy including the application of a Penning-type discharge
to a jet expansion source.
126
Penning-type sources are good sources of molecular
ions.
Bernard et al.
129
used a Penning source to study the A
2
PX
2
S

transition of
N

2
in the near-infrared. Highly excited infrared electronic transitions of CO (E
1
PB
1
S

and C
1
S

B
1
S

) were also investigated in this way.


130,131
8 Flames
The infrared spectroscopy of ames has a long history, starting in 1890 with Julius, as
discussed by Gaydon.
19
Hydrocarbon ames provide hot infrared emission spectra
of H
2
O, CO
2
, CN, C
2
, CO and OH.
For the purpose of this review the term ``ame'' refers mainly to hydrocarbon plus
oxygen (or air) ames. Infrared spectra of many other types of ames such as acety-
lene plus nitrous oxide
132
or C
2
F
4
plus oxygen
133
have been recorded mainly to study
the combustion process (rather than spectroscopy). Worden et al.
134
recorded the
infrared emission spectrum of a forest re from an airplane at high resolution. They
detected emission features due to several molecules, including hot H
2
O. Most of
the molecules detected (NH
3
, CO, CH
3
OH, etc.), however, appeared in absorption
against the continuum emitted by the re.
Early use of ames for high resolution infrared spectroscopy includes the work on
the A
2
PX
2
S

transition of CN by Bacis et al.,


135
and the Phillips (A
1
P
u
^X
1
S

g
)
136
and Ballik^Ramsay (b
3
S

g
^a
3
P
u
)
137
systems of C
2
. C
2
was produced
in an oxyacetylene torch and CN in a similar nitrous oxide^acetylene ame.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
197
The detection of vibration^rotation spectra is also possible. Indeed, the infrared
emission spectra of hydrocarbon ames are dominated by the main combustion
products CO
2
and H
2
O. A low pressure (20 mbar) methane plus oxygen ame
was used to study CO
2
and CO emission in the 1800^5000 cm
1
region.
138,139
Numerous highly excited energy levels could be assigned. An oxyacetylene torch
was also used to record the emission spectrum of hot water
140,141
and the OH
142
free radical from about 3000 to 10 000 cm
1
. The corresponding OD
vibration^rotation bands were made in a D
2
O
2
ame.
143
The Meinel system
of OH covers nearly the entire near-infrared region with bands in the Du = 1, 2
and 3 vibrational sequences.
The major advantage of ames is that they are a bright source of hot molecules
with temperatures ranging up to 3000 K. However, they have a number of draw-
backs including large pressure-broadened linewidths (typically *0.1 cm
1
) if they
are operated at atmospheric pressure. Because CO
2
and H
2
O so dominate the emi-
ssion spectra, recording data for other species can be difcult. More recent
spectroscopic work with ames (e.g., for H
2
O) has been at reduced total pressures
to reduce the pressure-broadened linewidths.
144
Finally, ames have been used to measure the rotational emission spectrum of
OH
145
in the far-infrared between 50 and 375 cm
1
. The rotational emission
spectrum of hot H
2
O has also been measured at somewhat higher wavenumbers.
144
9 Pyrotechnic and propellant combustion
The visible spectra of reworks are well known but infrared emission spectra of
various solid propellants, ares and other pyrotechnic materials have also been
recorded. The goal of this work is an understanding of the combustion process
and the measurement of the temperature. Infrared emission spectra of the combus-
tion of boron-containing,
146
silicon-containing
147
and magnesium-containing
148
pyrotechnics have been recorded by FTS emission spectroscopy at low resolution.
Various combustion products ranging from HCl to BO could be identied depending
on the composition of the are.
10 Shock tubes
Another interesting source for infrared emission is the shock tube. The shock com-
pression of an inert gas seeded with 1^10% or so of a precursor molecule results
in temperatures of 1000^4000 K, depending on the experimental conditions. The
heating is of short duration (*5 ms) so a rapid-scan infrared spectrometer is
required to make measurements. The collision-induced quadrupole emission from
H
2
was measured near 4000 cm
1
using a shock tube.
149,150
Emission spectra of
shock-heated small hydrocarbons were recorded by Stephens and Bauer.
151
A small
HgCdTe infrared array detector has been used with a spectrograph to measure
198
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
the emission from shock-heated NO.
152
All of these measurements were at low res-
olution, but a modern time-resolved FTS or a spectrograph like Phoenix
35
would
give much improved spectra.
11 Positive column of a DC discharge
In a typical direct current (DC) discharge, most of the cell is lled by the positive
column. The positive column is infrequently used for infrared emission spectroscopy
because the molecular glow is much less bright than from the plasma inside a hollow
cathode. The positive column is characterized by a nearly constant electric eld and a
modest voltage drop, leading to a relatively extended diffuse glow.
153
By moving the
electrodes to the side of the cell (Fig. 14), only the positive column emission is sent
into the spectrometer.
154
In addition, by injecting the carrier gas near the electrodes
and the precursor molecule directly into the positive column, a steady discharge
can be obtained. Although a hollow cathode discharge is brighter than the positive
column, the cathode surface is not very tolerant of hydrocarbon ``impurities''.
The larger discharge volume and the more stable operation in the presence of CH
4
,
etc., make the positive column an attractive emission source for some
applications.
154
Tokaryk and Civis
155,156
built a cooled hollow cathode similar to those used by
Oka and co-workers for the laser spectroscopy of ions. They were able to record
clean emission spectra of the

L
3
P
g
a
3
P
u
transition of
12
C
3
and
13
C
3
. A
methane^helium discharge was used to create a remarkably extensive series of bands
in the Du
2
= 0 sequence. A similar source provided ArH

in emission, although in
this case a hollow cathode gives a better signal.
157
Fig. 14 Schematic of the positive column discharge cell used to measure the vibration^rotation
emission of CD.
154
Reproduced with permission, from ref. 154.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
199
In Orsay, France, a DC discharge through pure NO gave the spectra of several
infrared electronic transitions of NO (M
2
S

^E
2
S

, D
2
S

^A
2
S

, E
2
S

^D
2
S

and E
2
S

^A
2
S

).
158,159
In addition, the vibration^rotation spectra of NO (Fig.
15) could be recorded up to u = 22 in the Du = 3 sequence.
160,161
The highly excited
vibration^rotation bands of CO and CO
2
were also recorded from the positive
column of a DC discharge.
162^167
In this case the CO
2
was diluted with He and/or
N
2
to mimic the conditions found in a CO
2
laser plasma. Time-resolved FTS spectra
of CO emission have also been carried out to study energy transfer.
168
12 Hollow cathode discharge
There are a large number of different types of hollow cathode discharges. For
example, Herzberg et al.
169
built a hollow cathode that is cooled by liquid nitrogen
in order to study the H
2
and H
3
molecules (Fig. 16). In this source it is possible
to view either the hollow cathode or the positive column.
169,170
This was very useful
because H
3
was found only in the cathode but H
2
appeared in both the cathode
and positive column regions. Bacis
171
has another design for a liquid-nitrogen-cooled
cathode but more typically water cooling is used.
An uncooled hollow cathode can run very hot and Davis et al. exploited this to
record new infrared spectra of the Phillips (A
1
P
u
^X
1
S

g
)
172
and Ballik^Ramsay
(b
3
S

g
^a
3
P
u
)
173
systems of C
2
. In this case, the C
2
was made largely by evaporation
of the red hot cathode rather than by sputtering from the surface. Ferguson et al.
174
Fig. 15 Vibration^rotation emission of NO excited in the positive column of a DC dis-
charge.
161
The 12^9 band head is visible near 4830 cm
1
. Reproduced from ref. 161, with
permission.
200
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
used a water-cooled steel cathode to record new spectra of the A
2
P
u
^X
2
S

g
(Meinel) system of N
2

. The Meinel system of N


2

is commonly seen in the spectra


of aurora (see below). A composite wall (SiC/Cu) hollow cathode
171
was used
to record the rst spectra of SiC
175,176
(d
1
S

^b
1
P, A
3
P

^X
3
P), SiC is isovalent
with C
2
but was produced by sputtering rather than evaporation.
Infrared electronic transitions of Rydberg molecules such as He
2
,
177
XeH
178
and
H
3
169
can also be seen in hollow cathodes. Rydberg transitions (up to 6 h5 g)
of H
2
and D
2
are also prominent
179^182
and the ionization potential of D
2
was deter-
mined by extrapolation
180
to be 124 745.353 cm
1
.
Transition-metal-containing molecules also commonly possess infrared electronic
transitions. The presence of an open d shell leads to a large number of low-lying
electronic states that are conveniently studied by infrared emission. Metal
monoxides are made by adding a trace of O
2
to the carrier gas (Ne or Ar, commonly)
and the metal is sputtered from the cathode surface, e.g., for HfO,
183
NiO,
184
PtO,
185
CoO,
186
CuO,
187
and AgO.
188
For metal mononitrides (ScN,
189
YN,
190
HfN,
191
Fig. 16 Liquid-nitrogen-cooled hollow cathode that was used by Herzberg et al.
169
to study an
infrared electronic transition of the Rydberg molecule, H
3
. Notice that the positive column
(anode glow) could also be observed for comparison purposes. Ions and Rydberg molecules
are more abundant in the cathode region.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
201
RuN
192
and OsN
193
) or monohydrides (ScH,
194
YH,
195
LaH
196
and PtH
197
) a trace
of N
2
or H
2
is effective, but monohalides are better made in other ways (furnaces
and microwave discharges) because the hollow cathode discharge is usually unstable.
All of the metal mononitrides listed above are new molecules and a typical spectrum
is illustrated for YN
190
in Fig. 17.
Vibration^rotation emission spectra are more difcult to measure than electronic
transitions because they are generally weaker and are at lower wavenumbers. Indeed,
no emission spectrum of a transient molecule has been detected using a hollow
cathode lamp below the InSb cut-off of 1800 cm
1
. The protonated rare gases,
NeH

,
198
ArH

,
199,200
KrH

,
200
and XeH

,
201
however, give excellent spectra from
a discharge of a rare gas plus a trace of H
2
. HeH

is too weak and is best detected by


laser spectroscopy. The vibration^rotation emission spectrum for CuH
202
was
inadvertently detected during the course of experiments on NeH

and remains
the only metal hydride detected in this way. Finally, a special high current hollow
cathode discharge was made to detect the H
3

and D
3

ions in emission.
203^206
This
work used the clever trick of pressure labelling to distinguish H
3

emission from
that of H
2
. Experimentally it was found that H
3

lines were prominent at 50 Torr


of total pressure but nearly absent at 8 Torr. The interfering H
2
lines, on the other
hand, tended to decrease in intensity as the pressure increased. H
3

emission can
be seen in the spectra of gas giant planets as discussed below.
13 Radio-frequency and microwave discharges
Radio-frequency (RF) and microwave discharges are similar (but not identical)
sources of excited molecules. RF discharges generally operate at 27 MHz and
microwave discharges at 2450 MHz to avoid interference with communications
Fig. 17 The 1^1 band of the A
1
S

^X
1
S

electronic transition of YN.


190
202
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
users. The main advantage of these sources is that they can operate with electrodes
external to the discharge tube. Although they can therefore work with very ``dirty''
precursor molecules such as hydrocarbons or metal halides, the build-up of deposits
inside the tube does change the power coupling into the discharge cell. By switching
the microwave source on and off, time-resolved spectra can be recorded.
207
The groups of Guelachvili and Vervloet are particularly fond of RF discharges. A
schematic of the reactor of Chollet et al.
208
is provided in Fig. 18 and some spectra
from this source in Fig. 19. An RF discharge was used to record the vibronic bands
of the A
2
P^

X
2
S

transition
209
of C
2
H and the c
3
P^b
3
S

transition
210
of
CO. Vibration^rotation emission spectra of HNO(n
1
),
211
NH
2
(2n
2
),
212
HNSi(n
1
),
213,214
HCN(n
1
),
215
HNC(n
1
),
215
NH,
216
SiH
217,218
and SH
219
were all
recorded with an RF plasma reactor (Fig. 18). The HNSi is a new transient molecule
related to HCN, but with the H bonded to the N atom, and was recorded along with
the A
2
P^X
2
S

spectrum of SiN
220
from a SiH
4
plus N
2
mixture.
A microwave discharge tube has also provided spectra of SH
221
and NH
36
/ND
222
as well as the BH,
223
PH
222,224
and CH
225,226
molecules. Very recently the spectrum
of SeH was recorded
227
and this molecule was the last main group monohydride
to be detected by vibration^rotation spectroscopy. The NH work includes a direct
comparison of the performance of the cryogenic echelle spectrograph, Phoenix, with
a Fourier transform spectrometer.
36
There is another common RF discharge called the inductively coupled plasma
(ICP) source that is used largely for analytical chemistry. The ICP is generally
operated at atmospheric pressure with a ow of argon gas. The discharge operates
at 27 MHz with 1^2 kW of power to achieve plasma temperatures of about
Fig. 18 The 3 m long radio-frequency plasma reactor described by Chollet et al.
208
with
White-type mirrors to collect more emission. Note that a simple loop of wire around a quartz
tube, as in the work of Vervloet or in the ICP source, also makes a satisfactory RF emission
cell. Reproduced from ref. 208, with permission.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
203
5000^6000 K.
228
The goal is to dissociate all molecules into atoms, which are then
monitored by emission spectroscopy or mass spectrometry. The ICP plasma, unlike
``normal'' RF and microwave discharges that are operated at low pressure and
*100 W of power, is a prolic source of ions. The ICP has, however, been occasion-
ally used for molecular spectroscopy. The infrared emission spectra of OH
229
and
OD
230
have been recorded with an ICP.
The microwave discharge is a particularly useful source for the infrared emission
of non-metal diatomics and a few triatomics. The precursor molecules are usually
introduced as gases. Molecules studied in this way include BN,
231
BF,
232
C
2
,
233^236
C
3
,
237
CN,
238
CP,
239,240
SiF,
241,242
NO,
243^245
N
2
,
246^250
P
2
251,252
and O
2
.
253
Of par-
ticular note are the Rydberg^Rydberg transitions of NO and SiF, and the discovery
of two new infrared electronic transitions
236
of C
2
(B
1
D
g
^A
1
P
u
, B
/
1
S

g
^A
1
P
u
)
and a triplet^triplet transition
237
of C
3
.
The electronic emission spectra of metal oxides, suldes and halides are readily
produced. The secret is to nd a relatively volatile precursor molecule (e.g.,
organometallics or halides) to aid in the introduction of the metal into the plasma.
For example, FeO can be made by reacting a mixture of ferrocene, argon and oxygen
in the discharge region.
254
Early work in this area was carried out, for example, by
Merer and co-workers. Molecules studied in this way are TiO,
255
VO,
256
CrO,
257
NbO,
258
AlO,
259
ZrS,
260
HfS,
261
TaS,
262
TiCl,
263
ZrCl
264,265
and YI.
266
An
interesting example is the discovery of the b
/
3
P^a
3
D system of ZrS by Jonsson
and Lindgren.
260
The ZrS emission was detected in the 7400^9700 cm
1
region from
Fig. 19 A time series of spectra recorded by Chollet et al.
208
with various gas mixtures.
Carbon-containing impurities in the cell are responsible for CO and HCN. Reproduced from
ref. 208, with permission.
204
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
the microwave discharge of a mixture of ZrCl
4
and sulfur powders without a carrier
gas. These new bands of ZrS matched absorption spectra of an unknown molecule
in the S-type star, R And.
14 Laser-excited infrared emission
With the exception of some vibration^rotation emission work on acetylene
267
and
the HF dimer,
28
almost all work has been on laser-excited electronic transitions.
The technique was pioneered by Verge s and Amiot, who, along with the Lyon group
in France, continue to dominate the eld. The typical experimental apparatus used
by Amiot
268
for Rb
2
is displayed in Fig. 20. Metal dimers are made in a heat pipe
oven and a laser beam is focussed through a hole in a mirror. The mirror is used
to send the laser-induced uorescence into a high resolution Fourier transform
spectrometer. To access highly excited states two lasers can be used. Verge s et al.
269
have written a recent review of the technique with examples.
Fig. 20 The apparatus used by Amiot
268
to record the infrared laser-induced uorescence of
K
2
. The l meter is a wavemeter and F.P. is a Fabry^Perot etalon that are used to monitor
the laser wavelength. Reproduced from ref. 268, with permission.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
205
The heat pipe oven has allowed numerous electronic states of all of the alkali
dimers, Li
2
,
270^274
Na
2
,
275^277
K
2
,
278^281
Rb
2
,
282^287
and Cs
2
,
288^292
(except the radio-
active Fr
2
), to be studied. The ground states have been followed nearly to dis-
sociation and highly reliable dissociation energies have been extracted. Fig. 21
shows a typical long vibrational progression in the A
1
S

u
state of
6
Li
2
. In this exper-
iment
272
the F
1
S

g
and E
1
S

g
states are populated by optical^optical double
resonance spectroscopy. The uorescence could be observed to the last few bound
levels of the A
1
S

g
state
272
and the dissociation energy D
e
was found to be
8517.03 cm
1
for the ground state of
6
Li
2
. The technique has also been applied
to the heteronuclear dimers, LiNa,
293
NaK
294,295
and RbCs.
296,297
The problem with
using a heat pipe oven for these molecules is the difference in vapour pressures of the
alkali metals. A heat pipe oven can thus make LiNa but LiCs is very difcult.
The original work on FTS detection of laser-induced uorescence was carried out
on I
2
. The B
3
P
0
u
^X
1
S

g
transition was excited with an argon ion laser and a long
ground state progression in the visible and near-infrared was seen.
298
Similar exper-
iments were carried out for Te
2
,
299
CsH
300
and Bi
2
.
301
A classic experiment was car-
ried out by Vervloet
302
who made NH
2
in an RF discharge and excited various
levels with a dye laser. The bending levels of NH
2
could be assigned up to u
2
=
10 and clearly showed the re-ordering of the K levels at the barrier to linearity.
The heat pipe oven was also used in a series of experiments on the alkaline earth
monohalides, CaF,
303,304
BaF,
305,306
BaCl
307^310
and BaI.
311
The monohalides were
made by heating a mixture of the metal and the metal dihalide to about 1000

C,
Fig. 21 The laser-induced uorescence of the F
1
S

g
A
1
S

u
and E
1
S

g
A
1
S

u
transitions of
6
Li
2
.
272
Fluorescence to the last few bound vibrational levels of the A state
is seen on the left. Reproduced from ref. 272, with permission.
206
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
e.g.,
Ba Ba!
2
2Ba! (31)
For CaF, BaF and BaCl, the B
/
2
D states were located by laser excitation of the C
2
P
X
2
S

transition and Fourier transform emission spectroscopy of the C


2
P B
/
2
D infrared transitions. For BaF and BaCl extensive measurements were made
on all of the low-lying electronic states.
Laser-induced uorescence can also be measured with the time-resolved FTS
technique. Typically a pulsed ultraviolet laser dissociates a molecule such as HCCH
and the infrared emission from a product such as C
2
H is detected,
312
C
2
H
2

hn
C
2
H H (32)
These experiments are carried out mainly to investigate photodissociation dynamics
although they could be adapted for high resolution spectroscopy.
15 Furnaces
Like the hollow cathode and arc discharges, the furnace is a traditional source of
molecular spectra. When Fourier transform spectrometers became available it
was natural to extend emission spectroscopy of electronic transitions into the
infrared. (The arc discharge is difcult to use in the infrared because source
uctuations are a problem for FTSs.) The early work on the Ballik^Ramsay system
of C
2
in a King furnace has already been cited
31
and similar spectra of Si
2
(d
1
S

g
^b
1
P
u
) have been measured.
313
Thermal emission spectra of the new molecule
BaLi
314,315
were measured through two near-infrared electronic transitions (2)
2
S

^X
2
S

and (2)
2
P^X
2
S

. A similar heat pipe oven source was used for Bi


2
316
and the forbidden B
/
2
D^X
2
S

transition of BaH.
317
Other molecules studied in
this way are MnH,
318
MnCl,
319
CrF,
320
CrCl,
321
CoF,
322
ScF,
323,324
ScCl,
325,326
ScI,
327
LaF,
328
LaCl,
329
LaS,
330
TiS,
331
CuS
332
and AlS.
333
Most notable here is
the analysis of the complex B
6
P^X
6
S

transition of CrF by Wallin et al.


320
CrF was made by mixing Cr powder with CrF
3
and heating the mixture to 1500 K
in a ceramic furnace.
The most interesting development, however, is the adaptation of thermal emission
for rotational and vibration^rotation spectroscopy. The method works for stable
molecules such as HF,
334,335
HCl,
335,336
HBr,
337
DCN
338
and H
2
O.
339^347
For
HF and HCl, pure rotational emission was also measured.
334,335
The H
2
O emission
spectra allowed the identication of hot H
2
O absorption in sunspots (``water on
the Sun'')
339,340
(Fig. 22). Water emission spectra have now been assigned from
400 to 6000 cm
1
for pure rotational as well as vibration^rotation emission. The
number of known energy levels of water was more than doubled in this work.
339^347
The very complex spectra of hot water led to the use of a direct variational prediction
of energy levels from a state-of-the-art ab initio surface in order to assign quantum
numbers.
340
A King furnace was used for the vibration^rotation emission spectra
of CN,
348
AlH,
349
SH,
350
CS,
351,352
CuH,
353
AgH,
353
and AuH.
353
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
207
Early work (1989) in vibration^rotation emission spectroscopy below the
1800 cm
1
cut-off of the InSb detector was carried out on GeS by Uehara et al.,
354
in addition to the work on FO.
37
This work was followed by the SiS emission
measurements by Frum et al.
355
in 1990. This spectacular emission measurement
(Fig. 23) was recorded inadvertently and convinced the author of this review of
the power and utility of the technique. Similar work on SiO
356
and GeO
357
followed.
In the case of SiO,
356
the laboratory emission measurements at 1400

C were com-
bined with SiO absorption measurements in a sunspot at 3200 K. The laboratory
SiO data were found by accident because of the reaction of molten gallium with
a mullite (2Al
2
O
3
SiO
2
) ceramic tube.
Vibration^rotation emission spectra of the metal hydrides LiH,
358
CaH,
359
SrH,
360
BaH,
361
AlH,
362
GaH,
363
InH
364
and BiH
365
were recorded using ceramic
(alumina or mullite) furnaces. All of the above molecules were formed by the
reaction of molten metals with H
2
gas (or D
2
, for deuterides) at high temperature.
Interestingly, the spontaneous reactions of Ca and Sr with H
2
did not occur,
360,361
presumably because of a large energy barrier to reaction. An electrical discharge
was then used to promote the reaction (Fig. 24) in the furnace. The pure rotational
emission spectrum of LiH and LiD (Fig. 25) was also recorded in the far-infrared
region.
358
Clearly infrared emission spectroscopy has the sensitivity to detect tran-
sient molecules even at very long wavelengths. Metal halides also give excellent
Fig. 22 The absorption spectrumof a sunspot (*3000 K) is on top and the laboratory emission
spectrum of hot water (*1800 K) is on the bottom.
347
All of the strong lines have been assigned
and the asterisks mark lines seen in the laboratory but not in the sunspot.
208
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
spectra, as illustrated by LiF,
366
NaF,
367
KF,
368
MgF,
369
CaF,
370
SrF,
371
BaF,
372
BF,
373
AlF,
373,374
GaF,
375
InF,
376,377
TlF,
378
NaCl,
379
KCl,
379,380
AlCl,
381
NaBr,
382
AlBr,
383
LiI,
384
CsI
385
and BeF
2
.
386
The CsI molecule gave a spectrumnear 100 cm
1
but only the vibration^rotation band heads could be measured.
385
The BeF
2
antisymmetric stretching mode and associated hot bands near 1500 cm
1
allowed
an equilibrium Be^F bond length of 1.373 to be determined.
386
16 Atmospheric science and remote sensing
Emission spectroscopy is relatively common in atmospheric science but the term has
a peculiar meaning in this area. For example, the infrared radiance of the Earth can
be detected from an aircraft or a satellite viewing in the nadir direction. This is often
Fig. 23 Vibration^rotation thermal emission spectrum of SiS
355
observed at 13 mm obtained
from a mixture of Si and SiS
2
powders at 1300 K.
Fig. 24 Schematic of the furnace used to record the vibration^rotation emission spectrum of
CaH
359
by discharging a mixture of Ca vapour and H
2
.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
209
called an ``emission'' spectrum although the atmospheric molecules of interest
appear in absorption against the blackbody radiance of the Earth. This review,
however, will adopt the usual denition of emission spectroscopy.
The most spectacular infrared emission spectra of our atmosphere are recorded in
the far-infrared region. Typically, a high resolution Fourier transform spectrometer
in a balloon records stratospheric spectra by looking out at the limb of the Earth.
An example of such a spectrum is presented as Fig. 26. The groups of Carli
387^389
in Florence and Traub
390
at the Smithsonian Astrophysics Observatory record
far-infrared spectra of this type. Specic molecules such as OH can be monitored
with a simpler system based on Fabry^Perot etalons, but nothing beats a Fourier
transform spectrometer for wide spectral coverage.
389
Mid-infrared emission exper-
iments of the limb of the Earth will be made with two satellite instruments,
MIPAS
391
(Michelson Interferometer for Passive Atmospheric Sounding) and
TES
392
(Tropospheric Emission Sounder). The difculty with these measurements
is the weak radiance emitted by the atmospheric molecules at 200^300 K. The
geometry is as drawn in Fig. 3, but the radiance of the end wall is very small because
the view is to deep space in the limb geometry. The goal in these experiments is to
understand the chemistry, particularly ozone chemistry, in our upper atmosphere.
Emission measurements of our atmosphere can also be made from the ground with
a Fourier transform spectrometer. A typical spectrum (Fig. 27) recorded by Evans
and Puckrin shows both emission and absorption lines.
393
Highly excited infrared emission from our atmosphere can be recorded from
airglow (nightglow and dayglow) as well as from aurora. Prominent infrared airglow
emission
394,395
is from the a
1
D
g
^X
3
S

g
system of O
2
and the Meinel system of OH.
Fig. 25 Pure rotational lines of a mixture of
6
LiH,
6
LiD,
7
LiH and
7
LiD obtained by detection
of far-infrared thermal emission with a Fourier transform spectrometer.
358
Reproduced from
ref. 358, with permission.
210
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
Fig. 26 Far-infrared emission spectrum of the stratosphere recorded by viewing the limb of the
Earth from a balloon.
26
Notice the excellent signal-to-noise ratio for the OH lines obtained with
a Fourier transform spectrometer. Reproduced from ref. 26, with permission.
Fig. 27 Low resolution infrared emission spectrum of the zenith sky in winter.
393
Most of the
lines appear in absorption but some are in emission. Reproduced from ref. 393, with the per-
mission of the American Meteorological Society.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
211
The OH emission is prominent at night from the mesopause region at 85^90 km and
this emission can be used to monitor atmospheric gravity waves.
395
Infrared auroral
emissions include the Meinel system
396
of N
2

.
The remote sensing of forest res by infrared emission spectroscopy has already
been discussed in the section on ames.
134
Other types of infrared emission remote
sensing are also possible. For example, the emission of stack gas has been monitored
at low resolution with a portable Fourier transform spectrometer.
397
A number of
recent emission spectra of volcanic plumes have also been recorded.
398,399
Love
et al.
399
noted that the ratio of SiF
4
/SO
2
seems to decrease just before a volcanic
eruption. There are also military applications for the remote sensing of rocket plumes
and other engine exhaust in the infrared.
17 Matrices
Free radicals and other transient molecules can be trapped in rare gas matrices.
Upon excitation, generally with a laser, the molecule will sometimes emit radiation
in the infrared. Two reviews by Bondybey et al.
400,401
have covered both the
spectroscopy and photophysics of recent matrix work. Electronic emission of the
Phillips and Ballik^Ramsay systems of C
2
402
as well as the BN,
403
BC
404
and
CuCl
2
405
molecules were seen in the Bondybey group. Vibrational emission from
WO
406
and ND
407
was measured, and ND displayed nearly unperturbed rotational
structure.
18 Liquids
Infrared emission spectroscopy of liquids has been carried out for thin lms of
molten salts.
15
For example, Bates and Boyd studied nitrate and chlorate melts.
408
Strong bands show some distortions but the use of a thick sample as a reference
alleviates this problem.
15
Infrared emission of molecules in solution is also possible
and Ogilby and co-workers, for example, have studied O
2
(
1
D
g
) emission in various
solvents.
409
19 Surfaces and solids
The presence of thin lm, even monolayer, surfaces can be readily monitored by
emission spectroscopy. As already mentioned, the best spectra are obtained with
thin samples of a few micrometers in thickness. Some band distortions are often
noticeable due to self-absorption
410
etc., but the use of a thick sample as a reference
eliminates the problem.
15
Sullivan et al.
411
have reviewed the use of emission
spectroscopy for surface analysis. A wide variety of lms have been studied, includ-
ing CO, metal oxides and organic molecules on a variety of substrates including
metals, non-metals such as Si, metal oxides and zeolites. A typical application would
be the study of a chemical reaction
412
(e.g., NO NH
3
on a catalyst surface, such as
212
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
V
2
O
5
). Emission spectroscopy has been used to monitor thin lms formed on silicon
wafers as part of the fabrication of semiconductor devices.
413
The emission of solid
particles in combustion has proved to be a useful diagnostic.
24,25
The spectra of bulk solids are most easily obtained by grinding the sample to a
powder and sprinkling it on a heated metal substrate.
414,415
Photoluminescence
of direct band gap semiconductors like GaAs is an important tool in solid state
physics.
416
Impurities in solids, e.g., ZnO:Cu
2
(some of these materials can lase),
can also be studied through their emission.
417
Time-resolved FTS spectroscopy
has also been applied to study the emission of solids, particularly for solid state
laser media.
418
An interesting variation on the technique is the use of a laser to heat
a thin surface layer of a solid in order to record emission.
419
20 Astronomy
Most astronomical infrared spectra appear in absorption against a bright continuum
due to a star or dust. Some sources, however, such as comets, provide spectacular
infrared emission spectra. Solar radiation evaporates the ``icy snowball'' as it nears
the Sun. The hot evaporated molecules and their photolysis products appear in emi-
ssion against the dark sky. Infrared electronic spectra of C
2
and CN are readily
seen
420
and more recently vibration^rotation lines of many molecules were found
in the bright comets Hale-Bopp and Hyakutake.
421
In a recent review Crovisier
421
has summarized the species detected through
vibration^rotation transitions so far and they include H
2
O, CO, CO
2
, CH
4
, C
2
H
2
,
C
2
H, CH
3
OH, HCN and OCS. To this list can be added CH.
422
Fig. 28 The spectrum of Jupiter's north tropical zone showing stratospheric emission from the
n
9
band of ethane on the left.
427
Reprinted, with permission, from ref. 427, #1984 by Annual
Reviews www.AnnualReviews.org
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
213
Most of these new molecules were detected with cryogenic echelle spectrographs
such as Phoenix
422
and a few observations were made from space with the Infrared
Space Observatory, ISO.
423
A particular surprise was the unexpectedly high abun-
dance of ethane, C
2
H
6
, in Hyakutake.
424
In our solar system, atomic and molecular features in the Sun and sunspots appear
in absorption although a few high-l Rydberg transitions of Mg and other atoms
appear in emission near 1000 cm
1
. The lines have very large Zeeman effects
and can be used to study solar magnetic elds.
425
Spectra of most of the planets,
satellites and asteroids are also absorption spectra, because the features are viewed
against the planetary blackbody emission. Typical spectra of rocks and gaseous
Fig. 29 The emission of the n
2
fundamental band of H
3

in the aurora of Uranus and Jupiter.


428
Reproduced from ref. 428, with permission.
214
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
CO
2
were recorded by the Thermal Emission Spectrometer (TES) on Mars.
426
As in
atmospheric science, the word ``emission'' is not always used in the same sense
as in this review. Some of the giant planets, however, display some interesting emi-
ssion spectra.
427
For example, the n
9
band of C
2
H
6
appears in emission in the strato-
sphere of Jupiter
427
(Fig. 28). The most beautiful spectra, however, are those of H
3

in the aurora of Jupiter


428,429
(Fig. 29). Saturn
430
and Uranus
431
also display H
3

emission. On Jupiter (and to a lesser extent on Saturn and Uranus)


431
the H
3

is formed by the interaction of the solar wind with the planetary magnetic eld
and emission is, therefore, localized near the magnetic poles. H
3

has become an
important tool for the study of Jupiter's atmosphere.
428
The three most important molecules in astronomy are H
2
, CO and H
2
O. The
forbidden H
2
vibration^rotation and pure rotational lines can be detected in shocked
regions.
21,432
H
2
, OH, H
2
O and CO emission can be observed in supernova
remnants.
432,433
H
2
emission is even detectable in starburst and Seyfert galaxies.
434
Star-forming regions in the Orion nebula also show H
2
,
21
H
2
O
435
and CO
436
in
infrared emission. Circumstellar envelopes can also show infrared emission as seen
for CO in the carbon star IRC10216
437
and H
2
O in the Mira variable (red giant),
R. Cas.
438
Planetary nebulae also show very strong emission spectra of mainly atoms and a
set of broad features that have been assigned to gaseous PAHs.
18,439^441
The idea
is that PAHs absorb ultraviolet radiation and after internal conversion emit in
the infrared.
439,440
This PAH assignment, however, is not universally accepted.
An alternative assignment is a carbonaceous solid
441
and even a giant
molecule/small particle is quite possible.
References
1 P. F. Bernath, Chem. Soc. Rev., 1996, 25, 111.
2 P. V. Huong, in Advances in Infrared and Raman Spectroscopy, Heyden, London, 1978, vol. 4, p. 85.
3 F. J. DeBlase and S. Compton, Appl. Spectrosc., 1991, 45, 611.
4 J. Mink and G. Keresztury, Appl. Spectrosc., 1993, 47, 1446.
5 G. Keresztury, in Encyclopaedia of Analytical Chemistry: Instrumentation and Applications, ed. R. A.
Meyers, J. Wiley & Sons, New York, 2000.
6 J. B. Bates, in Fourier Transform Infrared Spectroscopy, ed. J. R. Ferraro and L. J. Basile, Academic
Press, New York, 1978.
7 P. Bernath, Annu. Rev. Phys. Chem., 1990, 41, 91.
8 E. Hirota, Annu. Rep. Prog. Chem., Sect. C, 1994, 91, 3; E. Hirota, this volume.
9 E. Hirota, Chem. Rev., 1992, 92, 141.
10 M. Jacox, J. Phys. Chem. Ref. Data, Monograph No. 3, 1994.
11 R. F. Barrow and P. Crozet, Annu. Rep. Prog. Chem., Sect. C, 1992, 89, 353; R. F. Barrow and P.
Crozet and, Annu. Rep. Prog. Chem., Sect. C, 1997, 93, 147.
12 A. C. Legon, Annu. Rev. Phys. Chem., 1983, 34, 275.
13 E. Herbst, Annu. Rev. Phys. Chem., 1995, 46, 27.
14 For example,P. F. Bernath, S. A. Sinqueeld, L. L. Baxter, G. Sclippa, C. M. Rohlng and M. Bareld,
Vibr. Spectrosc., 1998, 16, 95.
15 J. Hvistendahl, E. Rytter and H. A. Oye, Appl. Spectrosc., 1983, 37, 182.
16 L. Nemes, R. S. Ram, P. F. Bernath, F. A. Tinker, M. C. Zumwalt, L. D. Lamb and D. R. Huffman,
Chem. Phys. Lett., 1994, 218, 295.
17 P. Colarusso, K.-Q. Zhang, B. Guo and P. F. Bernath, Chem. Phys. Lett., 1996, 269, 39.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
215
18 K.-Q. Zhang, B. Guo, P. Colarusso and P. F. Bernath, Science, 1996, 274, 582.
19 A. G. Gaydon, The Spectroscopy of Flames, Chapman and Hall, London, 2nd edn., 1974.
20 L. Lynds and B. A. Woody, Appl. Opt., 1988, 27, 1225.
21 N. Z. Scoville, D. N. B. Hall, S. G. Kleinman and S. T. Ridgway, Astrophys. J., 1982, 253, 136.
22 W. L. Wolfe, in The Infrared and Electro-Optical System Handbook, ed. G. J. Zissis, SPIE, Bellingham,
WA, 1993.
23 J. Bak and S. Clausen, J. Quant. Spectrosc. Radiat. Transfer, 1999, 61, 687.
24 J. R. Markham, K. Kinsella, R. M. Carangelo, C. R. Brouillette, M. D. Carangelo, P. E. Best and P. R.
Solomon, Rev. Sci. Instrum., 1993, 64, 2515.
25 P. R. Solomon and P. E. Best, in Combustion Measurements, ed. N. Chigier, Hemisphere, New York,
1991, p. 385.
26 B. Carli and M. Carlotti, in Spectroscopy of the Earth's Atmosphere and Interstellar Medium, ed. K. N.
Rao and A. Weber, Academic Press, San Diego, 1992.
27 P. F. Bernath, Spectra of Atoms and Molecules, Oxford University Press, New York, 1995, ch. 1.
28 H.-C. Chang and W. Klemperer, J.Chem. Phys., 1994, 100, 1.
29 M. A. Busch and K. W. Busch, Appl. Spectrosc., 1993, 47, 912.
30 Y. Zhang, K. W. Busch and M. A. Busch, Appl. Spectrosc., 1992, 46, 930.
31 E. A. Ballik and D. A. Ramsay, Astrophys. J., 1963, 137, 61.
32 A. E. Douglas and D. Sharma, J. Chem. Phys., 1953, 21, 448.
33 J. G. Moehlmann, J. T. Gleaves, J. W. Hudgens and J. D. McDonald, J. Chem. Phys., 1974, 60, 4790.
34 D. J. Cook, S. Schlemmer, N. Balucani, D. R. Wagner, J. A. Harrison, B. Steiner and R. J. Saykally, J.
Phys. Chem. A, 1998, 102, 1465.
35 K. H. Hinkle, R. Cuberly, N. Gaughan, J. Heynssens, R. R. Joyce, S. Ridgway, P. Schmitt and J. E.
Simmons, Proc. SPIE, 1998, 3354, 810; see also website:
www.noao.edu/kpno/phoenix/phoenix.htm1.
36 R. S. Ram, P. F. Bernath and K. H. Hinkle, J. Chem. Phys., 1999, 110, 5557.
37 P. D. Hammer, A. Sinha, J. B. Burkholder and C. J. Howard, J. Mol. Spectrosc., 1988, 129, 99.
38 J. J. Sloan and E. J. Kruus, in Time-Resolved Spectroscopy, ed. R. J. H. Clark and R. E. Hester, Wiley,
New York, 1989.
39 R. A. Palmer, G. D. Smith and P. Chen, Vibr. Spectrosc., 1999, 19, 131.
40 H. Weidner and R. E. Peale, Appl. Spectrosc., 1997, 51, 1106.
41 P. A. Martin and G. Guelachvili, Phys. Rev. Lett., 1990, 65, 2535.
42 N. Picque and G. Guelachvili, Vibr. Spectrosc., 1999, 19, 295; N. Picque and G. Guelachvili, Appl.
Opt., 1999 38, 1224.
43 A. Benidar, G. Guelachvili and P. A. Martin, Chem. Phys. Lett., 1991, 177, 563.
44 G. Guelachvili, J. Opt. Soc. Am. B, 1986, 3, 1718.
45 M. Elhanine, R. Farrenq and G. Guelachvili, Appl. Opt., 1989, 28, 4024.
46 M. A. Janssen, in Atmospheric Remote Sensing by Microwave Radiometry, ed. M. A. Janssen, Wiley,
New York, 1993.
47 J. C. Polanyi, Science, 1987, 236, 680.
48 J. K. Cashion and J. C. Polanyi, J. Chem. Phys., 1958, 28, 455.
49 C. M. Clayton, D. W. Merdes, J. Pliva, T. K. McCubbin and R. H. Tipping, J. Mol. Spectrosc., 1983,
98, 168.
50 N. I. Butkovskaya and D. W. Setser, J. Chem. Phys., 1997, 106, 5028.
51 J. J. Klaassen, J. Lindner and S. R. Leone, J. Chem. Phys., 1996, 104, 7403.
52 C. A. Carere, W. S. Neil and J. J. Sloan, Appl. Opt., 1996, 35, 2857.
53 D. E. Heard, R. A. Brownsword, D. G. Weston and G. Hancock, Appl. Spectrosc., 1993, 47, 1438.
54 T. J. Johnson, A. Simon, J. M. Weil and G. W. Harris, Appl. Spectrosc., 1993, 47, 1376.
55 C. D. Pibel, E. Sivota, J. Brenner and H.-L. Dai, J. Chem. Phys., 1998, 108, 1297.
56 D. D. Nelson, A. Schiffman, D. J. Nesbitt, J. J. Orlando and J. B. Burkholder, J. Chem. Phys., 1990, 93,
7003.
57 M. C. Abrams, S. P. Davis, M. L. P. Rao and R. Engleman, J. Mol. Spectrosc., 1994, 165, 57; M. C.
Abrams, S. P. Davis, M. L. P. Rao, R. Engleman and J. W. Brault, Astrophys. J. Suppl., 1994,
93, 351.
58 D. E. Osterbrock, J. P. Fulbright, P. C. Crosley and T. A. Barlow, Publ. Astron. Soc. Pacic, 1998, 110,
1499.
59 M. Beutel, K. D. Setzer, O. Shestakov and E. H. Fink, Chem. Phys. Lett., 1996, 249, 183.
60 M. Beutel, K. D. Setzer, O. Shestakov and E. H. Fink, J. Mol. Spectrosc., 1996, 178, 165.
61 M. Beutel, K. D. Setzer, O. Shestakov and E. H. Fink, J. Mol. Spectrosc., 1996, 179, 79.
62 K. Ziebarth, K. D. Setzer and E. H. Fink, J. Mol. Spectrosc., 1995, 173, 488.
63 E. H. Fink, K. D. Setzer, D. A. Ramsay and M. Vervloet, Chem. Phys. Lett., 1991, 177, 265.
64 R. Breidohr, O. Shestakov and E. H. Fink, J. Mol. Spectrosc., 1994, 168, 126.
65 H. Li, C. Focsa, R. Skelton and P. F. Bernath, unpublished.
66 W. E. McDermott, N. R. Pchelkin, D. J. Bernard and R. R. Bousek, Appl. Phys. Lett., 1978, 32, 469.
216
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
67 D. J. Bernard, B. K. Winker, T. A. Seder and R. H. Cohen, J. Phys. Chem., 1989, 93, 4790.
68 R. Huang, R. Zhang and R. N. Zare, Chem. Phys. Lett., 1990, 170, 437.
69 H. P. Yang, Y. Qin, T. J. Cui, Q. N. Yuan, X. B. Xie, Q. Zhuang and C. H. Zhang, Chem. Phys. Lett.,
1992, 191, 130.
70 A. J. Ross, P. Crozet, R. Bacis, S. Churassy, B. Erba, S. H. Ashworth, N. M. Lakin, M. R. Wickham, I.
R. Beattie and J. M. Brown, J. Mol. Spectrosc., 1996, 177, 134.
71 K. Ziebarth, R. Breidohr, O. Shestakov and E. H. Fink, Chem. Phys. Lett., 1992, 190, 271.
72 K. Ziebarth, K. D. Setzer, O. Shestakov and E. H. Fink, J. Mol. Spectrosc., 1998, 191, 108.
73 M. Beutel, K. D. Setzer and E. H. Fink, J. Mol. Spectrosc., 1999, 194, 250.
74 M. Beutel, K. D. Setzer and E. H. Fink, J. Mol. Spectrosc., 1999, 195, 147.
75 M. Beutel, K. D. Setzer and E. H. Fink, J. Mol. Spectrosc., 1996, 175, 48.
76 E. H. Fink, K. D. Setzer, D. A. Ramsay, M. Vervloet and J. M. Brown, J. Mol. Spectrosc., 1990, 142,
108.
77 E. H. Fink, K. D. Setzer, D. A. Ramsay, J. P. Towle and J. M. Brown, J. Mol. Spectrosc., 1996, 178,
143.
78 E. H. Fink, K. D. Setzer, D. A. Ramsay and M. Vervloet, Chem. Phys. Lett., 1991, 179, 95.
79 E. H. Fink, K. D. Setzer, D. A. Ramsay and M. Vervloet, J. Mol. Spectrosc., 1989, 138, 19.
80 J. Wildt, E. H. Fink, P. Briggs, R. P. Wayne and A. F. Vilesov, Chem. Phys., 1992, 159, 127.
81 E. H. Fink, H. Kruse, D. A. Ramsay and M. Vervloet, Can. J. Phys., 1986, 64, 242.
82 J. F. Noxon, Can. J. Phys., 1961, 39, 1110.
83 K. D. Setzer, E. H. Fink and D. A. Ramsay, J. Mol. Spectrosc., 1999, 198, 163.
84 I. Barnes, K. H. Becker and E. H. Fink, Chem. Phys. Lett., 1979, 67, 314.
85 E. H. Fink, H. Kruse and D. A. Ramsay, J. Mol. Spectrosc., 1986, 119, 377.
86 E. H. Fink, K. D. Setzer, U. Kottsieper, D. A. Ramsay and M. Vervloet, J. Mol. Spectrosc., 1988, 131,
127.
87 E. H. Fink, K. D. Setzer, D. A. Ramsay and M. Vervloet, J. Mol. Spectrosc., 1987, 125, 66.
88 E. H. Fink, H. Kruse, D. A. Ramsay and D.-C. Wang, Mol. Phys., 1987, 60, 277.
89 E. H. Fink, K. D. Setzer, D. A. Ramsay and Q.-S. Zhu, Can. J. Phys., 1994, 72, 919.
90 R. Winter, I. Barnes, E. H. Fink, J. Wildt and F. Zabel, J. Mol. Struct., 1982, 80, 75.
91 R. Winter, E. H. Fink, J. Wildt and F. Zabel, Chem. Phys. Lett., 1983, 94, 335.
92 E. H. Fink, K. D. Setzer, D. A. Ramsay and M. Vervloet and G. Z. Xu, J. Mol. Spectrosc., 1989, 136,
218.
93 R. Winter, I. Barnes, E. H. Fink, J. Wildt and F. Zabel, Chem. Phys. Lett., 1982, 86, 118.
94 O. Shestakov, R. Breidohr, H. Demes, K. D. Setzer and E. H. Fink, J. Mol. Spectrosc., 1998, 190, 28.
95 C. Linton, R. Bacis, F. Martin, S. Rosenwaks and J. Verge s, J. Chem. Phys., 1992, 96, 3422.
96 E. H. Fink, K. D. Setzer, D. A. Ramsay and M. Vervloet, Chem. Phys. Lett., 1991, 179, 103.
97 R. Breidohr, O. Shestakov and E. H. Fink, Chem. Phys. Lett., 1993, 218, 13.
98 R. Breidohr, K. D. Setzer, O. Shestakov, E. H. Fink and W. Zyrnicki, J. Mol. Spectrosc., 1994, 166,
251.
99 R. Breidohr, K. D. Setzer, O. Shestakov, E. H. Fink and W. Zyrnicki, J. Mol. Spectrosc., 1994, 166,
471.
100 R. Breidohr, O. Shestakov, K. D. Setzer and E. H. Fink, J. Mol. Spectrosc., 1995, 172, 369.
101 R. P. Tuckett, P. A. Freedman and W. J. Jones, Mol. Phys., 1979, 37, 379; R. P. Tuckett, P. A.
Freedman and W. J. Jones, Mol. Phys. 1979, 37, 403.
102 E. H. Fink and D. A. Ramsay, J. Mol. Spectrosc., 1997, 185, 304.
103 K. J. Holstein, E. H. Fink, J. Wildt and F. Zabel, Chem. Phys. Lett., 1985, 113, 1.
104 E. H. Fink, O. Shestakov and K. D. Setzer, J. Mol. Spectrosc., 1997, 183, 163.
105 A. F. Vilesov, J. Wildt and E. H. Fink, Chem. Phys., 1991, 153, 531.
106 D. Bailly, S. A. Tashkun, V. I. Perevalov, J. L. Teffo and Ph. Arcas, J. Mol. Spectrosc., 1998, 190, 1.
107 M. Horani and M. Vervloet, Astron. Astrophys., 1992, 256, 683.
108 (a) A. T. Droege and P. C. Engelking, Chem. Phys. Lett., 1983, 96, 316; (b) P. C. Engelking, Rev. Sci.
Instrum., 1986, 57, 2274.
109 K. P. Huber and M. Vervloet, J. Mol. Spectrosc., 1991, 146, 188.
110 (a) D. Cossart, J. Mol. Spectrosc., 1994, 167, 11; (b) D. Cossart, C. Cossart-Magos, B. Gandara and J.
M. Robbe, J. Mol. Spectrosc., 1985, 109, 166.
111 I. Dabrowski, D. W. Tokaryk and J. K. G. Watson, J. Mol. Spectrosc., 1998, 189, 95.
112 I. Dabrowski, D. W. Tokaryk, R. H. Lipson and J. K. G. Watson, J. Mol. Spectrosc., 1998, 189, 110.
113 I. Dabrowski, G. Di Lonardo, G. Herzberg, J. W. C. Johns, D. A. Sadovskii and M. Vervloet, J. Chem.
Phys., 1992, 97, 7093.
114 I. Dabrowski, D. W. Tokaryk, M. Vervloet and J. K. G. Watson, J. Chem. Phys., 1996, 104, 8245.
115 I. Dabrowski, G. Herzberg, B. P. Hurley, R. H. Lipson, M. Vervloet and D. C. Wang, Mol. Phys., 1988,
63, 269.
116 I. Dabrowski and D. A. Sadovskii, Mol. Phys., 1994, 81, 291.
117 I. Dabrowski and D. A. Sadovskii, J. Chem. Phys., 1997, 107, 8874.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
217
118 I. Dabrowski, G. Herzberg and R. H. Lipson, Mol. Phys., 1988, 63, 289.
119 G. Herzberg, Annu. Rev. Phys. Chem., 1987, 38, 27.
120 F. Roux, F. Michaud and M. Vervloet, J. Mol. Spectrosc., 1994, 164, 510.
121 K. P. Huber and M. Vervloet, J. Mol. Spectrosc., 1992, 153, 17.
122 K. P. Huber and M. Vervloet, J. Mol. Spectrosc., 1988, 129, 1.
123 M. Vervloet, A. L. Roche and Ch. Jungen, Phys. Rev. A, 1988, 38, 5489.
124 K. P. Huber, M. Vervloet, Ch. Jungen and A. L. Roche, Mol. Phys., 1987, 61, 501.
125 D. W. Tokaryk, G. R. Wagner, R. L. Brooks and J. L. Hunt, J. Chem. Phys., 1995, 103, 10439.
126 D. Cossart and C. Cossart-Magos, Chem. Phys. Lett., 1996, 250, 128.
127 D. Cossart, M. Bonneau and J. M. Robbe, J. Mol. Spectrosc., 1987, 125, 413.
128 D. Cossart and M. Elhanine, Chem. Phys. Lett., 1998, 285, 83; D. Cossart, J. Chim. Phys., 1979, 76,
1045.
129 A. Bernard, M. Larzillie re, C. Effantin and A. J. Ross, Astrophys. J., 1993, 413, 829.
130 C. Amiot, J.-Y. Roncin and J. Verge s, J. Phys. B, 1986, 19, L19.
131 J.-Y. Roncin, A. Ross and E. Boursey, J. Mol. Spectrosc., 1993, 162, 353.
132 S. T. Darian and M. Vanpee, Combust. Flame, 1987, 70, 65.
133 C. H. Douglass, H. D. Ladouceur, V. A. Shamanian and J. R. McDonald, Combust. Flame, 1995, 100,
529.
134 H. Worden, R. Beer and C. P. Rinsland, J. Geophys. Res., 1997, 102, 1287.
135 R. Bacis, D. Cerny, J. d'Incan, G. Guelachvili and F. Roux, Astrophys. J., 1977, 214, 946.
136 J. Chauville, J.-P. Maillard and A. W. Mantz, J. Mol. Spectrosc., 1977, 68, 399.
137 C. Amiot, J. Chauville and J.-P Maillard, J. Mol. Spectrosc., 1979, 75, 19.
138 D. Bailly, C. Camy-Peyret and R. Lanquetin, J. Mol. Spectrosc., 1997, 182, 10.
139 D. Bailly, S. A. Tashkun, V. I. Perevalov, J. L. Teffo and Ph. Arcas, J. Mol. Spectrosc., 1999, 197, 114.
140 J.-M. Flaud, C. Camy-Peyret and J.-P. Maillard, Mol. Phys., 1976, 32, 499.
141 C. Camy-Peyret, J.-M. Flaud, J.-P. Maillard and G. Guelachvili, Mol. Phys., 1977, 33, 1641.
142 J.-P. Maillard, J. Chauville and A. W. Mantz, J. Mol. Spectrosc., 1976, 63, 120.
143 C. Amiot, J.-P. Maillard and J. Chauville, J. Mol. Spectrosc., 1981, 87, 196.
144 R. Lanquetin, L. H. Coudert and C. Camy-Peyret, J. Mol. Spectrosc., 1999, 195, 54.
145 J. L. Hardwick and G. C. Whipple, J. Mol. Spectrosc., 1991, 147, 267.
146 K. J. Smit, R. J. Hancox, D. J. Hatt, S. P. Murphy and L. V. De Yong, Appl. Spectrosc., 1997, 51, 1400.
147 K. J. Smit, L. V. De Yong and R. Gray, Chem. Phys. Lett., 1996, 254, 197.
148 J. Wang, Z. Chen, Y. Luo, C. Zhu, X. Jin and T. Wang, Spectrosc. Lett., 1992, 25, 1355.
149 R. Krech, G. Caledonia, S. Schertzer, K. Ritter, T. Wilkerson, L. Cotnoir, R. Taylor and G. Birnbaum,
Phys. Rev. Lett., 1982, 49, 1913.
150 G. E. Caledonia, R. H. Krech, T. D. Wilkerson, R. L. Taylor and G. Birnbaum, Phys. Rev. A, 1991, 43,
6010.
151 K. M. Stephens and S. H. Bauer, Spectrochim. Acta, 1994, 50A, 741.
152 W. T. Rawlins, R. R. Foutter and T. E. Parker, J. Quant. Spectrosc. Radiat. Transfer, 1993, 49, 423.
153 A. von Engel, Electric Plasmas: Their Nature and Uses, Taylor and Francis, London, 1983.
154 I. Morino, K. Matsumura and K. Kawaguchi, J. Mol. Spectrosc., 1995, 174, 123.
155 D. W. Tokaryk and S. Civis, J. Chem. Phys., 1995, 103, 3928.
156 S. Civis and D. W. Tokaryk, J. Mol. Spectrosc., 1995, 172, 543.
157 P. A. Martin and G. Guelachvili, Chem. Phys. Lett., 1991, 180, 344.
158 C. Amiot and J. Verge s, Phys. Scr., 1982, 26, 422.
159 C. Amiot, Chem. Phys. Lett., 1981, 83, 40.
160 C. Amiot and J. Verge s, J. Mol. Spectrosc., 1980, 81, 424.
161 C. Amiot, J. Mol. Spectrosc., 1982, 94, 150.
162 D. Bailly, J. Mol. Spectrosc., 1998, 192, 257.
163 D. Bailly, C. Rossetti and G. Guelachvili, Chem. Phys., 1985, 100, 101.
164 D. Bailly and N. Legay, J. Mol. Spectrosc., 1993, 157, 1.
165 A. Campargue, D. Bailly, J.-L. Teffo, S. A. Tashkun and V. I. Perevalov, J. Mol. Spectrosc., 1999, 193,
204.
166 R. Farrenq, C. Rossetti, G. Guelachvili and W. Urban, Chem. Phys., 1985, 92, 389.
167 D. Bailly and C. Rossetti, J. Mol. Spectrosc., 1984, 105, 215; D. Bailly and C. Rossetti, J. Mol.
Spectrosc., 1984, 105,229; D. Bailly and C. Rossetti, J. Mol. Spectrosc., 1984, 105, 331.
168 T. Nakanaga, F. Ito and H. Takeo, Chem. Phys. Lett., 1993, 206, 73.
169 G. Herzberg, H. Lew, J. J. Sloan and J. K. G. Watson, Can. J. Phys., 1981, 59, 428.
170 S. P. Reddy and C. V. V. Prasad, J. Phys. E: Sci. Instrum., 1989, 22, 308.
171 R. Bacis, J. Phys. E: Sci. Instrum., 1976, 9, 1081.
172 S. P. Davis, M. C. Abrams, J. G. Phillips and M. L. P. Rao, J. Opt. Soc. Am. B, 1988, 5, 2280.
173 S. P. Davis, M. C. Abrams, Sandalphon, J. W. Brault and M. L. P. Rao, J. Opt. Soc. Am. B, 1988, 5,
1838.
174 D. W. Ferguson, K. N. Rao, P. A. Martin and G. Guelachvili, J. Mol. Spectrosc., 1992, 153, 599.
218
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
175 P. F. Bernath, S. A. Rogers, L. C. O'Brien, C. R. Brazier and A. D. McLean, Phys. Rev. Lett., 1988, 60,
197.
176 C. R. Brazier, L. C. O'Brien and P. F. Bernath, J. Chem. Phys., 1989, 91, 7384.
177 S. A. Rogers, C. R. Brazier, P. F. Bernath and J. W. Brault, Mol. Phys., 1988, 63, 901.
178 M. Douay, S. A. Rogers and P. F. Bernath, Mol. Phys., 1988, 64, 425.
179 I. Dabrowski and G. Herzberg, Acta Phys. Hung., 1984, 55, 219.
180 Ch. Jungen, I. Dabrowski, G. Herzberg and M. Vervloet, J. Mol. Spectrosc., 1992, 153, 11.
181 Ch. Jungen, I. Dabrowski, G. Herzberg and D. Kendall, J. Chem. Phys., 1989, 91, 3926.
182 Ch. Jungen, I. Dabrowski, G. Herzberg and M. Vervloet, J. Chem. Phys., 1990, 93, 2289.
183 R. S. Ram and P. F. Bernath, J. Mol. Spectrosc., 1995, 169, 268.
184 R. S. Ram and P. F. Bernath, J. Mol. Spectrosc., 1992, 155, 315.
185 C. I. Frum, R. Engleman and P. F. Bernath, J. Mol. Spectrosc., 1991, 150, 566.
186 R. S. Ram, C. N. Jarman and P. F. Bernath, J. Mol. Spectrosc., 1993, 160, 574.
187 L. C. O'Brien, R. L. Kubicek, S. J. Wall, D. E. Koch, R. J. Friend and C. R. Brazier, J. Mol. Spectrosc.,
1996, 180, 365.
188 L. C. O'Brien, S. J. Wall and M. K. Sieber and J. Mol. Spectrosc., 1997, 183, 57.
189 R. S. Ram and P. F. Bernath, J. Chem. Phys., 1992, 96, 6344.
190 R. S. Ram and P. F. Bernath, J. Mol. Spectrosc., 1994, 165, 97.
191 R. S. Ram and P. F. Bernath, J. Mol. Spectrosc., 1997, 184, 401.
192 R. S. Ram, J. Lie vin and P. F. Bernath, J. Chem. Phys., 1998, 109, 6329.
193 R. S. Ram, J. Lie vin and P. F. Bernath, J. Chem. Phys., 1999, 111, 3449.
194 R. S. Ram and P. F. Bernath, J. Chem. Phys., 1996, 105, 2668; R. S. Ram and P. F. Bernath, J. Mol.
Spectrosc., 1997, 183, 263.
195 R. S. Ram and P. F. Bernath, J. Chem. Phys., 1994, 101, 9283; R. S. Ram and P. F. Bernath, J. Mol.
Spectrosc., 1995, 171, 169.
196 R. S. Ram and P. F. Bernath, J. Chem. Phys., 1996, 104, 6444.
197 M. C. McCarthy, R. W. Field, R. Engleman and P. F. Bernath, J. Mol. Spectrosc., 1993, 158, 208.
198 R. S. Ram, P. F. Bernath and J. W. Brault, J. Mol. Spectrosc., 1985, 113, 451.
199 J. W. Brault and S. P. Davis, Phys. Scr., 1982, 25, 268.
200 J. W. C. Johns, J. Mol. Spectrosc., 1984, 106, 124.
201 S. A. Rogers, C. R. Brazier and P. F. Bernath, J. Chem. Phys., 1987, 87, 159.
202 R. S. Ram, P. F. Bernath and J. W. Brault, J. Mol. Spectrosc., 1985, 113, 269.
203 W. A. Majewski, M. D. Marshall, A. R. W. McKellar, J. W. C. Johns and J. K. G. Watson, J. Mol.
Spectrosc., 1987, 122, 341.
204 W. A. Majewski, P. A. Feldman, J. K. G. Watson, S. Miller and J. Tennyson, Astrophys. J. Lett., 1989,
347, L51.
205 W. A. Majewski, A. R. W. McKellar, D. Sadovskii and J. K. G. Watson, Can. J. Phys., 1994, 72, 1016.
206 T. Amano, M.-C. Chan, S. Civis, A. R. W. McKellar, W. A. Majewski, D. Sadovskii and J. K. G.
Watson and Can. J. Phys., 1994, 72, 1007.
207 G. Durry and G. Guelachvili, J. Mol. Spectrosc., 1994, 168, 82.
208 P. Chollet, G. Guelachvili, M. Morillon-Chapey, P. Gressier and J. P. M. Schmitt, J. Opt. Soc. Am. B,
1986, 3, 687.
209 M. Vervloet and M. Herman, Chem. Phys. Lett., 1988, 144, 48.
210 I. Dabrowski, M. Vervloet and D.-C. Wang, Can. J. Phys., 1987, 65, 1171.
211 J. C. Petersen and M. Vervloet, Chem. Phys. Lett., 1987, 141, 499.
212 A. R. W. McKellar, M. Vervloet, J. B. Burkholder and C. J. Howard, J. Mol. Spectrosc., 1990, 142,
319.
213 M. Elhanine, R. Farrenq and G. Guelachvili, J. Chem. Phys., 1991, 94, 2529.
214 M. Elhanine, B. Hanoune and G. Guelachvili, J. Chem. Phys., 1993, 99, 4970.
215 N. Picque and G. Guelachvili, Spectrochim. Acta., 2000, 56A, 681.
216 D. Boudjaadar, J. Brion, P. Chollet, G. Guelachvili and M. Vervloet, J. Mol. Spectrosc., 1986, 119, 352.
217 J. C. Knights, J. P. M. Schmitt, J. Perrin and G. Guelachvili, J. Chem. Phys., 1982, 76, 3414.
218 M. Betrencourt, D. Boudjaadar, P. Chollet, G. Guelachvili and M. Morillon-Chapey, J. Chem. Phys.,
1986, 84, 4121.
219 A. Benidar, R. Farrenq, G. Guelachvili and C. Chackerian, J. Mol. Spectrosc., 1991, 147, 383.
220 M. Elhanine, B. Hanoune, G. Guelachvili and C. Amiot, J. Phys. II (France), 1992, 2, 931.
221 R. S. Ram, P. F. Bernath, R. Engleman and J. W. Brault, J. Mol. Spectrosc., 1995, 172, 34.
222 R. S. Ram and P. F. Bernath, J. Mol. Spectrosc., 1996, 176, 329.
223 F. S. Pianalto, L. C. O'Brien, P. C. Keller and P. F. Bernath, J. Mol. Spectrosc., 1988, 129, 348.
224 R. S. Ram and P. F. Bernath, J. Mol. Spectrosc., 1987, 122, 275.
225 P. F. Bernath, J. Chem. Phys., 1987, 86, 4838.
226 P. F. Bernath, C. R. Brazier, T. Olsen, R. Hailey, W. T. M. L. Fernando, C. Woods and J. L. Hardwick,
J. Mol. Spectrosc., 1991, 147, 16.
227 R. S. Ram and P. F. Bernath, unpublished.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
219
228 L. M. Faires, B. A. Palmer, R. Engleman and T. M. Niemczyk, Spectrochim. Acta, 1984, 39B, 819.
229 M. C. Abrams, S. P. Davis, M. L. P. Rao and R. Engleman, Astrophys. J., 1990, 363, 326.
230 M. C. Abrams, S. P. Davis, M. L. P. Rao and R. Engleman, Pramana, 1992, 39, 163.
231 R. S. Ram and P. F. Bernath, J. Mol. Spectrosc., 1996, 180, 414.
232 H. Brehdohl, I. Dubois, F. Me len and M. Vervloet, J. Mol. Spectrosc., 1988, 129, 145.
233 C. Amiot and J. Verge s, Astrophys. J., 1982, 263, 993.
234 K. Islami and C. Amiot, J. Mol. Spectrosc., 1986, 118, 132.
235 M. Douay, R. Nietmann and P. F. Bernath, J. Mol. Spectrosc., 1988, 131, 250.
236 M. Douay, R. Nietmann and P. F. Bernath, J. Mol. Spectrosc., 1988, 131, 261.
237 H. Sasada, T. Amano, C. Jarman and P. F. Bernath, J. Chem. Phys., 1991, 94, 2401.
238 C. Amiot and J. Verge s, Chem. Phys. Lett., 1983, 95, 189.
239 R. S. Ram and P. F. Bernath, J. Mol. Spectrosc., 1987, 122, 282.
240 R. S. Ram, S. Tam and P. F. Bernath, J. Mol. Spectrosc., 1992, 152, 89.
241 H. Bredohl, I. Dubois, F. Me len and M. Vervloet, J. Mol. Spectrosc., 1988, 129, 232.
242 M. Vervloet, H. Bredohl, I. Dubois and F. Me len, J. Mol. Spectrosc., 1988, 131, 53.
243 C. Amiot and J. Verge s, Chem. Phys. Lett., 1979, 66, 570.
244 C. Amiot and J. Verge s, Phys. Scr., 1982, 25, 302.
245 A. Bernard, C. Effantin, J. d'Incan, C. Amiot and J. Verge s, Mol. Phys., 1991, 73, 221.
246 F. Roux, C. Effantin, J. d'Incan and J. Verge s, J. Mol. Spectrosc., 1982, 91, 238.
247 F. Roux and F. Michaud, J. Mol. Spectrosc., 1991, 149, 441.
248 F. Roux, D. Cerny and J. Verge s, J. Mol. Spectrosc., 1982, 94, 302.
249 F. Roux, F. Michaud and J. Verge s, J. Mol. Spectrosc., 1983, 97, 253.
250 A. Faye, Q. Kou, R. Farrenq and G. Guelachvili, J. Mol. Spectrosc., 1999, 197, 147.
251 C. Effantin, R. Bacis, C. Amiot and J. Verge s, J. Mol. Spectrosc., 1978, 69, 79.
252 C. Amiot, C. Effantin, J. d'Incan and J. Verge s, J. Mol. Spectrosc., 1978, 72, 189.
253 C. Amiot and J. Verge s, Can. J. Phys., 1981, 59, 1391.
254 A. W. Taylor, A. S.-C. Cheung and A. J. Merer, J. Mol. Spectrosc., 1985, 113, 487.
255 D. C. Galehouse, J. W. Brault and S. P. Davis, Astrophys. J. Suppl., 1980, 42, 241.
256 A. J. Merer, G. Huang, A. S.-C. Cheung and A. W. Taylor, J. Mol. Spectrosc., 1987, 125, 465.
257 A. S.-C. Cheung, W. Zyrnicki and A. J. Merer, J. Mol. Spectrosc., 1984, 104, 315.
258 O. Launila, B. Schimmelpfenning, H. Fagerli, O. Gropen, A. G. Taklif and U. Wahlgren, J. Mol.
Spectrosc., 1997, 186, 131.
259 O. Launila and J. Jonsson, J. Mol. Spectrosc., 1994, 168, 1.
260 J. Jonsson and B. Lindgren, J. Mol. Spectrosc., 1995, 169, 30.
261 O. Launila, J. Jonsson, G. Edvinsson and A. G. Taklif, J. Mol. Spectrosc., 1996, 177, 221.
262 S. Wallin, G. Edvinsson and A. G. Taklif, J. Mol. Spectrosc., 1998, 192, 368; S. Wallin, G. Edvinsson
and A. G. Taklif, J. Mol. Spectrosc., 1997, 184, 466.
263 R. S. Ram and P. F. Bernath, J. Mol. Spectrosc., 1997, 186, 113.
264 J. G. Phillips, S. P. Davis and D. C. Galehouse, Astrophys. J. Suppl., 1980, 43, 417.
265 R. S. Ram and P. F. Bernath, J. Mol. Spectrosc., 1997, 186, 335.
266 A. Bernard, F. Roux and J. Verge s, J. Mol. Spectrosc., 1980, 80, 374.
267 M. Saarinen, D. Permogorov and L. Halonen, J. Chem. Phys., 1999, 110, 1424.
268 C. Amiot, J. Chem. Phys., 1990, 93, 8591.
269 J. Verge s, C. Amiot, R. Bacis and A. J. Ross, Spectrochim. Acta, 1995, 51A, 1191.
270 J. Verge s, R. Bacis, B. Barakat, P. Carrot, S. Churassy and P. Crozet, Chem. Phys. Lett., 1983, 98, 203.
271 I. Russier, F. Martin, C. Linton, P. Crozet, A. J. Ross, R. Bacis and S. Churassy, J. Mol. Spectrosc.,
1994, 168, 39.
272 C. Linton, F. Martin, I. Russier, A. J. Ross, P. Crozet, S. Churassy and R. Bacis, J. Mol. Spectrosc.,
1996, 175, 340.
273 I. Russier, A. Yiannopoulou, P. Crozet, A. J. Ross, F. Martin and C. Linton, J. Mol. Spectrosc., 1997,
184, 129.
274 A. J. Ross, P. Crozet, C. Linton, F. Martin, I. Russier and A. Yiannopoulou, J. Mol. Spectrosc., 1998,
191, 28.
275 J. Verge s, C. Effantin, J. d'Incan, A. Topouzkhanian and R. F. Barrow, Chem. Phys. Lett., 1983, 94, 1.
276 J. Verge s, C. Effantin, J. d'Incan, D. L. Cooper and R. F. Barrow, Phys. Rev. Lett., 1984, 53, 46.
277 R. F. Barrow, C. Amiot, J. Verge s, J. d'Incan, C. Effantin and A. Bernard, Chem. Phys. Lett., 1991,
183, 94.
278 C. Amiot, J. Mol. Spectrosc., 1991, 146, 370.
279 C. Amiot, J. Verge s and C. E. Fellows, J. Chem. Phys., 1995, 103, 3350.
280 I. Russier, M. Aubert-Fre con, A. J. Ross, F. Martin, A. Yiannopoulou and P. Crozet, J. Chem. Phys.,
1998, 109, 2717.
281 I. Russier-Antoine, A. J. Ross, M. Aubert-Fre con, F. Martin, P. Crozet and S. Magnier, J. Phys. B,
1999, 32, 4039.
282 C. Amiot, P. Crozet and J. Verge s, Chem. Phys. Lett., 1985, 121, 390.
220
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
283 C. Amiot and J. Verge s, Mol. Phys., 1987, 61, 51.
284 C. Amiot, Chem. Phys. Lett., 1995, 241, 133.
285 C. Amiot and J. Verge s, Chem. Phys. Lett., 1997, 274, 91.
286 C. Amiot, Mol. Phys., 1986, 58, 667.
287 C. Amiot, O. Dulieu and J. Verge s, Phys. Rev. Lett., 1999, 83, 2316.
288 C. Amiot, C. Crepin and J. Verge s, Chem. Phys. Lett., 1984, 106, 162; C. Amiot, C. Crepin and J.
Verge s, Chem. Phys. Lett., 1983, 98, 608.
289 C. Amiot and J. Verge s, Chem. Phys. Lett., 1985, 116, 273.
290 C. Amiot, J. Chem. Phys., 1988, 89, 3993.
291 J. Verge s and C. Amiot, J. Mol. Spectrosc., 1987, 126, 393.
292 C. Amiot, W. Demtrder and C. R. Vidal, J. Chem. Phys., 1988, 88, 5265.
293 C. E. Fellows, J. Verge s and C. Amiot, Mol. Phys., 1988, 63, 1115; C. E. Fellows, J. Verge s and C.
Amiot, J. Chem. Phys., 1990, 93, 6281.
294 A. J. Ross, R. M. Clements and R. F. Barrow, J. Mol. Spectrosc., 1988, 127, 546.
295 R. F. Barrow, R. M. Clements, G. Delacretaz, J. d'Incan, A. J. Ross, J. Verge s and L. Wste, J. Phys. B,
1987, 20, 3047.
296 T. Gustavsson, C. Amiot and J. Verge s, Mol. Phys., 1988, 64, 279; T. Gustavsson, C. Amiot and J.
Verge s, Chem. Phys. Lett., 1988, 143, 101.
297 C. E. Fellows, R. F. Gutterres, A. P. C. Campos, J. Verge s and C. Amiot, J. Mol. Spectrosc., 1999, 197,
19.
298 D. Cerny, R. Bacis and J. Verge s, J. Mol. Spectrosc., 1986, 116, 458.
299 J. Verge s, C. Effantin, O. Babaky, J. d'Incan, S. J. Prosser and R. F. Barrow, Phys. Scr., 1982, 25, 338.
300 C. Crepin, J. Verge s and C. Amiot, Chem. Phys. Lett., 1984, 112, 10.
301 R. F. Barrow, F. Taher, J. d'Incan, C. Effantin, A. J. Ross, A. Topouzkhanian, G. Wannous and J.
Verge s, Mol. Phys., 1996, 87, 725.
302 M. Vervloet, Mol. Phys., 1988, 63, 433.
303 J. Verge s, C. Effantin, A. Bernard, A. Topouzkhanian, A. R. Allouche, J. d'Incan and R. F. Barrow, J.
Phys. B, 1993, 26, 279.
304 C. M. Gittens, N. A. Harris, R. W. Field, J. Verge s, C. Effantin, A. Bernard, J. d'Incan, W. E. Ernst, P.
Bndgen and B. Engels, J. Mol. Spectrosc., 1993, 161, 303.
305 R. F. Barrow, A. Bernard, C. Effantin, J. d'Incan, G. Fabre, A. El Hachimi, R. Stringat and J. Verge s,
Chem. Phys. Lett., 1988, 147, 535.
306 C. Effantin, A. Bernard, J. d'Incan, G. Wannous, J. Verge s and R. F. Barrow, Mol. Phys., 1990, 70,
735.
307 C. Amiot and J. Verge s, Chem. Phys. Lett., 1991, 185, 310.
308 M. Had, C. Amiot and J. Verge s, Chem. Phys. Lett., 1993, 210, 45.
309 C. Amiot, M. Had and J. Verge s, J. Phys. B, 1993, 26, L407.
310 C. Amiot, M. Had and J. Verge s, J. Mol. Spectrosc., 1996, 180, 121.
311 R. F. Gutterres, J. Verge s and C. Amiot, J. Mol. Spectrosc., 1999, 196, 29.
312 T. R. Fletcher and S. R. Leone, J. Chem. Phys., 1989, 90, 871.
313 S. P. Davis and J. W. Brault, J. Opt. Soc. Am. B, 1987, 4, 20.
314 R. Stringat, G. Fabre, A. Boulezhar, J. d'Incan, C. Effantin, J. Verge s and A. Bernard, J. Mol.
Spectrosc., 1994, 168, 514.
315 J. d'Incan, C. Effantin, A. Bernard, G. Fabre, R. Stringat, A. Boulezhar and J. Verge s, J. Chem. Phys.,
1994, 100, 945.
316 R. F. Barrow, J. Chevaleyre, C. Effantin, M. A. Lebeault-Dorget, A. J. Ross, G. Wannous and J.
Verge s, Chem. Phys. Lett., 1993, 214, 293.
317 R. F. Barrow, B. J. Howard, A. Bernard and C. Effantin, Mol. Phys., 1991, 72, 971.
318 W. J. Balfour, O. Launila and L. Klynning, Mol. Phys., 1990, 69, 443.
319 O. Launila, Mol. Phys., 1992, 76, 319.
320 S. Wallin, R. Koivisto and O. Launila, 1996, 105, 388.
321 M. Bencheikh, R. Koivisto, O. Launila and J. P. Flamand, J. Chem. Phys., 1997, 106, 6231.
322 R. S. Ram, P. F. Bernath and S. P. Davis, J. Chem. Phys., 1996, 104, 6949.
323 E. A. Shenyavskaya, J. Verge s, A. Topouzkhanian, M.-A. Lebeault-Dorget, J. d'Incan, C. Effantin and
A. Bernard, J. Mol. Spectrosc., 1994, 164, 129.
324 E. A. Shenyavskaya, M.-A. Lebeault-Dorget, C. Effantin, J. d'Incan, A. Bernard and J. Verge s, J. Mol.
Spectrosc., 1995, 171, 309.
325 F. Taher, C. Effantin, A. Bernard, E. A. Shenyavskaya and J. Verge s, J. Mol. Spectrosc., 1996, 199,
223.
326 F. Taher, A. Bernard, C. Effantin, J. d'Incan, E. A. Shenyavskaya and J. Verge s, J. Mol. Spectrosc.,
1996, 179, 229.
327 F. Taher, C. Effantin, A. Bernard, J. d'Incan and J. Verge s, J. Mol. Spectrosc., 1998, 189, 220.
328 J. Verge s, C. Effantin, J. d'Incan, A. Bernard and E. A. Shenyavskaya, J. Mol. Spectrosc., 1999, 198,
196.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
221
329 J. Xin and L. Klynning, Phys. Scr., 1994, 49, 209.
330 R. J. Winkel, S. P. Davis and M. C. Abrams, Appl. Opt., 1996, 35, 2874.
331 J. Jonsson and O. Launila, Mol. Phys., 1993, 79, 95.
332 L. C. O'Brien, M. Dulick and S. P. Davis, J. Mol. Spectrosc., 1999, 195, 328.
333 O. Launila and J. Jonsson, J. Mol. Spectrosc., 1994, 168, 483.
334 R. S. Ram, Z. Morbi, B. Guo, K.-Q. Zhang, P. F. Bernath, J. Vander Auwera, J. W. C. Johns and S. P.
Davis, Astrophys. J. Suppl., 1996, 103, 247.
335 R. B. LeBlanc, J. B. White and P. F. Bernath, J. Mol. Spectrosc., 1994, 164, 574.
336 T. Parekunnel, T. Hirao, R. J. Le Roy and P. F. Bernath, J. Mol. Spectrosc., 1999, 195, 185.
337 V. Braun and P. F. Bernath, J. Mol. Spectrosc., 1994, 167, 282.
338 W. Quapp, M. Hirsch, G. C. Mellau, S. Klee, M. Winnewisser and A. Maki, J. Mol. Spectrosc., 1999,
195, 284.
339 L. Wallace, P. Bernath, W. Livingston, K. Hinkle, J. Busler, B. Guo and K. Zhang, Science, 1995, 268,
1155.
340 O. L. Polyansky, N. F. Zobov, S. Viti, J. Tennyson, P. F. Bernath and L. Wallace, Science, 1997, 277,
346.
341 O. L. Polyansky, J. R. Busler, B. Guo, K. Zhang and P. F. Bernath, J. Mol. Spectrosc., 1996, 176, 305.
342 O. L. Polyansky, J. Tennyson and P. F. Bernath, J. Mol. Spectrosc., 1997, 186, 213.
343 O. L. Polyansky, N. F. Zobov, J. Tennyson, J. A. Lotoski and P. F. Bernath, J. Mol. Spectrosc., 1997,
184, 35.
344 N. F. Zobov, O. L. Polyansky, J. Tennyson, J. A. Lotoski, P. Colarusso, K.-Q. Zhang and P. F.
Bernath, J. Mol. Spectrosc., 1999, 193, 118.
345 O. L. Polyansky, N. F. Zobov, S. Viti, J. Tennyson, P. F. Bernath and L. Wallace, J. Mol. Spectrosc.,
1997, 186, 422.
346 O. L. Polyansky, N. F. Zobov, S. Viti, J. Tennyson, P. F. Bernath and L. Wallace, Astrophys. J. Lett.,
1997, 489, L205.
347 N. F. Zobov, O. L. Polyansky, J. Tennyson, S. V. Shirin, R. Nassar, T. Hirao, T. Imajo, P. F. Bernath
and L. Wallace, Astrophys. J. Suppl., 2000, 530, 994.
348 S. P. Davis, M. C. Abrams, M. L. P. Rao and J. W. Brault, J. Opt. Soc. Am. B, 1991, 8, 198.
349 J. L. Deutch, W. S. Neil and D. A. Ramsay, J. Mol. Spectrosc., 1987, 125, 115.
350 R. J. Winkel and S. P. Davis, Can. J. Phys., 1984, 62, 1420.
351 R. Winkel, S. P. Davis, R. Pecyner and J. W. Brault, Can. J. Phys., 1984, 62, 1414.
352 R. S. Ram, P. F. Bernath and S. P. Davis, J. Mol. Spectrosc., 1995, 173, 146.
353 J. Y. Seto, Z. Morbi, F. Charron, S. K. Lee, P. F. Bernath and R. J. Le Roy, J. Chem. Phys., 1999, 110,
11756.
354 H. Uehara, K. Horiai, K. Sueoka and K. Nakagawa, Chem. Phys. Lett., 1989, 160, 149.
355 C. I. Frum, R. Engleman and P. F. Bernath, J. Chem. Phys., 1990, 93, 5457.
356 J. M. Campbell, D. Klapstein, M. Dulick, P. F. Bernath and L. Wallace, Astrophys. J. Suppl., 1995,
101, 237.
357 E. G. Lee, J. Y. Seto, T. Hirao, P. F. Bernath and R. J. Le Roy, J. Mol. Spectrosc., 1999, 194, 197.
358 M. Dulick, K.-Q. Zhang, B. Guo and P. F. Bernath, J. Mol. Spectrosc., 1998, 188, 14.
359 C. I. Frum and H. M. Pickett, J. Mol. Spectrosc., 1993, 159, 329.
360 C. I. Frum, J. J. Oh, E. A. Cohen and H. M. Pickett, J. Mol. Spectrosc., 1994, 163, 339.
361 K. A. Walker, H. G. Hedderich and P. F. Bernath, Mol. Phys., 1993, 78, 577.
362 J. B. White, M. Dulick and P. F. Bernath, J. Chem. Phys., 1993, 99, 8371.
363 J. M. Campbell, M. Dulick, D. Klapstein, J. B. White and P. F. Bernath, J. Chem. Phys., 1993, 99, 8379.
364 J. B. White, M. Dulick and P. F. Bernath, J. Mol. Spectrosc., 1995, 169, 410.
365 H. G. Hedderich and P. F. Bernath, J. Mol. Spectrosc., 1993, 158, 170.
366 H. G. Hedderich, C. I. Frum, R. Engleman and P. F. Bernath, Can. J. Phys., 1991, 69, 1659.
367 A. Muntianu, B. Guo and P. F. Bernath, J. Mol. Spectrosc., 1996, 176, 274.
368 M.-C. Liu, A. Muntianu, K.-Q. Zhang, P. Colarusso and P. F. Bernath, J. Mol. Spectrosc., 1996, 180,
188.
369 B. E. Barber, K.-Q. Zhang, B. Guo and P. F. Bernath, J. Mol. Spectrosc., 1995, 169, 583.
370 F. Charron, B. Guo, K.-Q. Zhang and Z. Morbi and P. F. Bernath, J. Mol. Spectrosc., 1995, 171, 160.
371 P. Colarusso, B. Guo, K.-Q. Zhang and P. F. Bernath, J. Mol. Spectrosc., 1996, 175, 158.
372 B. Guo, K.-Q. Zhang and P. F. Bernath, J. Mol. Spectrosc., 1995, 170, 59.
373 K.-Q. Zhang, B. Guo, V. Braun, M. Dulick and P. F. Bernath, J. Mol. Spectrosc., 1995, 170, 82.
374 H. G. Hedderich and P. F. Bernath, J. Mol. Spectrosc., 1992, 153, 73.
375 H. Uehara, K. Horiai, K. Nakagawa and H. Suguro, Chem. Phys. Lett., 1991, 178, 553.
376 T. Karkanis, M. Dulick, Z. Morbi, J. B. White and P. F. Bernath, Can. J. Phys., 1994, 72, 1213.
377 H. Uehara, K. Horiai, T. Mitani and H. Suguro, Chem. Phys. Lett., 1989, 162, 137.
378 H. Uehara, K. Horiai, A. Kerim and N. Aota, Chem. Phys. Lett., 1992, 189, 217.
379 R. S. Ram, M. Dulick, B. Guo, K.-Q. Zhang and P. F. Bernath, J. Mol. Spectrosc., 1997, 183, 360.
380 H. Uehara, K. Horiai, T. Konno and K. Miura, Chem. Phys. Lett., 1990, 169, 599.
222
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
381 H. G. Hedderich, M. Dulick and P. F. Bernath, J. Chem. Phys., 1993, 99, 8363.
382 H. Uehara, K. Horiai, A. Kerim, Y. Ozaki and T. Konno, Chem. Phys. Lett., 1993, 213, 101.
383 H. Uehara, K. Horiai, Y. Osaki and T. Konno, Chem. Phys. Lett., 1993, 214, 527.
384 B. Guo, M. Dulick, S. Yost and P. F. Bernath, Mol. Phys., 1997, 91, 459.
385 V. Braun, B. Guo, K.-Q. Zhang, M. Dulick, P. F. Bernath and G. A. McRae, Chem. Phys. Lett., 1994,
228, 633.
386 C. I. Frum, R. Engleman and P. F. Bernath, J. Chem. Phys., 1991, 95, 1435.
387 B. Carli, F. Mencaraglia and A. Bonetti, Int. J. Infrared Millimeter Waves, 1980, 1, 263; B. Carli, F.
Mencaraglia and A. Bonetti, Int. J. Infrared Millimeter Waves, 1982, 3, 385.
388 B. Carli, F. Mencaraglia, A. Bonetti, B. M. Dinelli and F. Forni, Int. J. Infrared Millimeter Waves,
1983, 4, 475.
389 B. Carli, M. Carlotti, B. M. Dinelli, F. Mencaraglia and J. H. Park, J. Geophys. Res., 1989, 94, 11049.
390 D. G. Johnson, K. W. Jucks, W. A. Traub and K. V. Chance, J. Geophys. Res., 1995, 100, 3091.
391 H. Fischer and H. Oelhaf, Appl. Opt., 1996, 35, 2787.
392 R. Beer, Jet Propulsion Laboratory, see website: tes.jpl.nasa.gov
393 W. F. J. Evans and E. Puckrin, J. Climate, 1995, 8, 3091.
394 W. F. J. Evans, D. M. Hunten, E. J. Llewellyn and A. Vallance Jones, J. Geophys. Res., 1968, 73, 2885.
395 R. P. Lowe and D. N. Turnbull, Geophys. Res. Lett., 1995, 22, 2813.
396 A. Vallance Jones, in Auroral Physics., ed. C.-I. Meng, M. J. Rycroft and L. A. Frank, Cambridge
University Press, Cambridge, 1991.
397 R. C. Carlson, A. F. Hayden and W. V. Telfair, Appl. Opt., 1988, 27, 4952.
398 C. Oppenheimer, P. Francis, M. Burton, A. J. H. Maciejewski and L. Boardman, Appl. Phys. B, 1998,
67, 505.
399 S. P. Love, F. Goff, D. Counce, C. Siebe and H. Delgado, Nature, 1998, 396, 563.
400 V. E. Bondybey, A. M. Smith and J. Agreiter, Chem. Rev., 1996, 96, 2113.
401 V. E. Bondybey, M. Rsnen and A. Lammers, Annu. Rep. Prog. Chem., Sect. C, 1999, 95, 331.
402 G. M. Lask, J. Agreiter, R. Schlachta and V. E. Bondybey, Chem. Phys. Lett., 1993, 205, 31.
403 M. Lorenz, J. Agreiter, A. M. Smith and V. E. Bondybey, J. Chem. Phys., 1996, 104, 3143.
404 A. M. Smith, M. Lorenz, J. Agreiter and V. E. Bondybey, Mol. Phys., 1996, 88, 247.
405 M. Lorentz, N. Caspary, W. Foeller, J. Agreiter, A. M. Smith and V. E. Bondybey, Mol. Phys., 1997,
91, 483.
406 M. Lorentz, J. Agreiter, N. Caspary and V. E. Bondybey, Chem. Phys. Lett., 1998, 291, 291.
407 N. Caspary, B. E. Wurfel, A. M. Smith and V. E. Bondybey, Chem. Phys. Lett., 1997, 220, 241.
408 J. B. Bates and G. E. Boyd, Appl. Spectrosc., 1973, 27, 204.
409 R. D. Scurlock and P. R. Ogilby, J. Phys. Chem., 1987, 91, 4599.
410 P. R. Grifths, Appl. Spectrosc., 1972, 26, 73.
411 D. H. Sullivan, W. C. Conner and M. P. Harold, Appl. Spectrosc., 1992, 46, 811.
412 D. H. Sullivan, M. P. Harold and W. C. Conner, J. Catal., 1998, 178, 108.
413 J. E. Franke, T. M. Niemczyk and D. M. Haaland, Spectrochim. Acta, 1994, 50A, 1687.
414 A. M. Vassallo, P. A. Cole-Clarke, L. S. K. Pang and A. J. Palmisano, Appl. Spectrosc., 1992, 46, 73.
415 A. M. Vassallo and K. S. Finnie, Appl. Spectrosc., 1992, 46, 1477.
416 A. Bignazzi, E. Grilli, M. Radice, M. Guzzi and E. Castiglioni, Rev. Sci. Instrum., 1996, 67, 666.
417 B. M. Kimpel and H.-J. Schulz, Phys. Rev. B, 1991, 43, 9938.
418 H. Weidner and R. E. Peale, in OSA TOPS on Advanced Solid-State Lasers, ed. S. A. Payne and C.
Pollock, Optical Society of America, Washington, D. C., 1996, vol. 1.
419 L. T. Lin, D. D. Archibald and D. E. Honigs, Appl. Spectrosc., 1988, 42, 477.
420 J. R. Johnson, U. Fink and H. P. Larson, Astrophys. J., 1983, 270, 769.
421 J. Crovisier, Faraday Discuss., 1998, 108, 437.
422 D. Reuter, D. Jennings, K. Hinkle, S. Ridgway, L. Wallace, P. Bernath and J. Waller, IAU Circular,
6681, 1997.
423 J. Crovisier, K. Leech, D. Bockele e-Morvan, T. Y. Brooke, M. S. Hanner, B. Altieri, H. U. Keller and
E. Lellouch, Science, 1997, 275, 1904.
424 M. J. Mumma, M. A. DiSanti, N. Dello Russo, M. Fomenkova, K. Magee-Sauer, C. D. Kaminski and
D. A. Xie, Science, 1996, 272, 1310.
425 D. Deming, R. J. Boyle, D. E. Jennings and G. Wiedemann, Astrophys. J., 1988, 333, 978.
426 P. R. Christensen, D. L. Anderson, S. C. Chase, R. T. Clancy, R. N. Clark, B. J. Conrath, H. H. Kieffer,
R. O. Kuzmin, M. C. Malin, J. C. Pearl, T. L. Roush and M. D. Smith, Science, 1998, 279, 1692.
427 S. T. Ridgway and J. W. Brault, Annu. Rev. Astron. Astrophys., 1984, 22, 291.
428 S. Miller, H. A. Lam and J. Tennyson, Can. J. Phys., 1994, 72, 760.
429 T. Oka and T. R. Geballe, Astrophys. J. Lett., 1990, 351, L53.
430 T. R. Geballe, M.-F. Jagod and T. Oka, Astrophys. J. Lett., 1993, 408, L109.
431 H. A. Lam, S. Miller, R. D. Joseph, T. R. Geballe, L. A. Trafton, J. Tennyson and C. H. Ballester,
Astrophys. J. Lett., 1997, 474, L73.
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224
223
432 M. A. Richter, J. R. Graham, G. S. Wright, D. M. Kelly and J. H. Lacy, Astrophys. J. Lett., 1995, 449,
L83.
433 W. T. Reach and J. Rho, Astrophys. J. Lett., 1998, 507, L93.
434 H. Mouri, Y. Tamiguchi, K. Kawara and M. Nishida, Astrophys. J. Lett., 1989, 346, L73.
435 M. Harwit, D. A. Neufeld, G. J. Melnick and M. J. Kaufman, Astrophys. J. Lett., 1998, 497, L105.
436 T. Geballe and R. Garden, Astrophys. J., 1987, 317, L107.
437 R. Sahai and P. G. Wannier, Astrophys. J., 1985, 299, 424.
438 Truong-Bach, R. J. Sylvester, M. J. Barlow, Nguyen-Q-Rieu, T. Lim, X. W. Liu, J. P. Baluteau, S.
Deguchi, K. Justtanont and A. G. G. M. Tielens, Astron. Astrophys., 1999, 345, 925.
439 J. L. Puget and A. Le ger, Annu. Rev. Astron. Astrophys., 1989, 27, 161.
440 L. J. Allamandola, A. G. G. M. Tielens and J. R. Barker, Astrophys. J. Suppl., 1989, 71, 733.
441 C. Joblin, Faraday Discuss., 1998, 109, 349.
224
Annu. Rep. Prog. Chem., Sect. C, 2000, 96, 177^224

S-ar putea să vă placă și