Вы находитесь на странице: 1из 11

Fourier Transform Infrared Studies of Ammonia Photochemistry in

Solid Parahydrogen
Mahmut Ruzi and David T. Anderson*
Department of Chemistry, University of Wyoming, Laramie, Wyoming 82071, United States
*S Supporting Information
ABSTRACT: We present 193 nm in situ photochemical
studies of NH
3
isolated in solid parahydrogen (pH
2
) at 1.8 K
using Fourier Transform Infrared (FTIR) spectroscopy. By
recording FTIR spectra during and after irradiation we are able
to identify and assign a number of rovibrational transitions to
ortho-NH
2
(X
2
B
1
) and NH(X
3

). Spectroscopic analysis
shows that these two radical species rotate freely in solid
pH
2
and that eects of the unpaired electron spin remain
essentially unchanged from the gas phase. We provide detailed
mechanistic studies that show the nascent ortho-NH
2
photoproduct is rapidly cooled within the pH
2
matrix to the
ground vibrational and rotational state before (1) subsequent
photodissociation or (2) tunneling-driven reaction (k
tun
= 1.88(17) min
1
) with the pH
2
host to produce ortho-NH
3
in a defect
site. Once the ortho-NH
3
is produced in this defect site it slowly converts (k
conv
= 7.72(51) 10
3
min
1
) back to a single
substitution site even at 1.8 K. We demonstrate the in situ photolysis of NH
3
can be utilized to generate NH doped pH
2
solids
that are relatively stable at low temperature. However, the ortho-NH
2
+ pH
2
ortho-NH
3
+ H back reaction substantially limits
the sequential two-photon conversion of NH
3
to NH. These studies also reveal that extended photolysis of the NH
3
/pH
2
system
results in the generation of high concentrations of orthohydrogen that must result from repeated cycles of photodissociation and
NH
2
back reaction within the pH
2
host.

INTRODUCTION
In the present work we investigate the feasibility of using in situ
photochemistry of NH
3
trapped in parahydrogen (pH
2
)
matrixes as an eective means to prepare high concentrations
of the imidogen radical (NH). We also hope to utilize the NH
3
photochemical system as a benchmark to develop a microscopic
understanding of the in situ photochemistry to better engineer
sample conditions for future cryochemistry experiments.
15
The NH
3
photochemical precursor interests us for three
reasons: (i) the NH
3
gas phase photochemistry has been
extensively studied,
69
(ii) the NH
3
/pH
2
chemical system only
involves nitrogen and hydrogen, and (iii) there is the possibility
that the pH
2
matrix can participate in the photochemistry. We
started these experiments with the photochemical strategy of
producing NH through two-photon sequential photodissocia-
tion of NH
3
,
= + + h NH (X, v 0) (193 nm) NH (X B ) H( S)
3 2 2
2
1
2
(1)
+

+ h NH (X B ) (193 nm) NH(A ) H( S)
2
2
1
3 2
(2)
This two-photon formation of NH has been studied in the gas
phase;
6,7
absorption of a 193 nm photon excites NH
3
to the
NH
3
(A,v
2
= 6) excited state that then dissociates with near unit
quantum yield ( = 0.965) to produce NH
2
(X
2
B
1
) + H. The
NH
2
(X
2
B
1
) product is produced highly vibrationally and
rotationally excited with 70% of the available energy going
into internal excitation of this fragment.
8
The ground electronic
state NH
2
(X
2
B
1
) radical can then absorb another 193 nm
photon to produce NH(A

3
). In our studies we utilize a low
per pulse photolysis uence to minimize multiphoton
processes, but high repetition rates to rapidly drive the
sequential two-photon chemistry. Possible reactions of the
two photoproducts NH
2
and NH in their ground electronic
state with the matrix host are as follows.
10,11
+ + =

H NH (X B ) H NH H 1.74 kcal mol
2
2
1 2 3
1
(3)
+ + = +

H NH(X ) H NH H 11.8 kcal mol


3
2 2
1
(4)
Accordingly, we expect reaction 3 should readily occur in solid
pH
2
because the reaction partners are held in constant contact
and therefore this reaction must be compensated by fast
photoproduction of cold NH
2
(X
2
B
1
) to successfully produce
NH(A

3
). Reaction 3 has been extensively studied (both
forward and reverse) as a prototypical polyatomic H-atom
Special Issue: Terry A. Miller Festschrift
Received: August 20, 2013
Revised: September 24, 2013
Published: September 24, 2013
Article
pubs.acs.org/JPCA
2013 American Chemical Society 13832 dx.doi.org/10.1021/jp408336n | J. Phys. Chem. A 2013, 117, 1383213842
abstraction reaction.
10,12
In the forward direction reaction 3 is
slightly exothermic and is predicted
10
to have a barrier of
approximately +11.34 kcal mol
1
. This means that reaction 3
will be a tunneling-driven chemical reaction between the
NH
2
(X
2
B
1
) photoproduct and the pH
2
host. On the other hand
reaction 4 is endothermic and thus should be completely shut
down at cryogenic temperatures.
11
The insertion reaction of
NH(X
3

) with H
2
(not indicated) to produce triplet NH
3
is a
high-barrier endothermic process that is not expected to be
important.
13
Given this reaction scheme, NH(X
3

) in solid
pH
2
should be kinetically trapped with respect to reactions with
the pH
2
host even though thermodynamically the reaction NH
+ H
2
NH
3
is highly exothermic (H 96 kcal mol
1
).
14
What we did not know before this study was to what extent the
nascent NH radical in the A
3
excited electronic state will react
with the pH
2
host and would we be able to suciently
overcome reaction 3 using a high repetition rate broadband ArF
laser.
The other unknown was the importance of H-atom reactions
in the measured photochemistry. If H-atoms are produced
either directly or as byproducts of the in situ photochemistry,
and given reactions 14 they will be, then the photochemical
mechanisms will get more complicated. Even at 1.8 K, the
temperature at which most of the photochemistry was
performed in this study the H-atom is still mobile via the H
+ H
2
H
2
+H chemical tunneling mechanism.
15
Given that the
reaction NH(X
3

) + H NH
2
is highly exothermic and
barrierless,
16
we might expect this reaction to proceed at the
diusion limit
17
under high H-atom concentrations and thus
may limit the achievable NH concentration. We are interested
in the importance of H-atom reactions for the present NH
3
photochemical studies, but also for other precursor molecules
because H-atoms are likely byproducts for almost all in situ
photochemical studies in solid pH
2
.
Another reason to study the in situ photochemistry of NH
3
in solid pH
2
is the possible photoproducts are all expected to
freely rotate. This is not true in general; typically only small
molecules and hydrides that occupy single substitution sites
rotate freely in solid pH
2
. For example, the molecules CH
4
,
H
2
O, CO, and HCl have all been shown to freely rotate,
1822
but the larger species HCOOH, CH
3
OH, CH
3
F, and N
2
O do
not.
2326
Previous studies from our group have shown that
NH
3
freely rotates in solid pH
2
.
27
Therefore, we expect that
NH
2
and NH should freely rotate. Since both of these species
are radicals with unpaired electron spin, we are also interested
to see how the pH
2
matrix perturbs the spin angular
momentum of these radicals. Conservation of this rotational
quantum label for a molecule trapped in solid pH
2
permits
more detailed assignments of the species involved. For species
with indistinguishable H-atoms (NH
2
and NH
3
), we can also
follow whether nuclear spin is conserved in the various
photochemical steps.
28
All of these reasons combine to make
the NH
3
/pH
2
chemical system an excellent test case for the
study of photochemistry in a pH
2
quantum solid.

EXPERIMENTAL METHODS
The ammonia (NH
3
) doped pH
2
crystals are prepared using
the rapid vapor deposition method of Fajardo and Tam.
29,30
Here we outline the general procedure and emphasize the
specic details for these studies. The crystal is grown by
codeposition of independent gas ows of NH
3
and pH
2
onto a
precooled BaF
2
optical substrate held at approximately 2.5 K
during deposition using a sample-in-vacuum liquid-He bath
cryostat. The NH
3
(Sigma-Aldrich, 99.9%) gas is introduced to
the cryostat through a stainless steel tube equipped with a
needle valve. The NH
3
concentration in the pH
2
samples is
determined using the measured ow rates of dopant and pH
2
gas. For these studies the NH
3
concentration ranged from 10 to
100 ppm. The pH
2
solids are prepared by enriching normal-H
2
gas to greater than 99.97% pH
2
enrichment levels using a
variable temperature ortho/para converter operated near 14.0
K. The orthohydrogen (oH
2
) concentration in the sample can
be checked using the integrated intensity of the oH
2
-induced
Q
1
(0) feature
29
and the measured crystal thickness.
5
To generate NH
2
and NH in solid pH
2
we use 193 nm in situ
photolysis of the NH
3
precursor. The 193 nm radiation is
generated with a broadband ArF excimer laser (Gam Laser
EX5) with unstable resonator optics (2 1 mrad far eld
divergence), 8 ns pulse duration, and a maximum pulse energy
of 12 mJ pulse
1
. The unfocused output of the ArF excimer
laser is steered into the cryostat at an angle of 45 with respect
to the surface normal of the BaF
2
optical substrate. This
permits Fourier Transform Infrared (FTIR) spectra to be
recorded within the photolysis region either during or
immediately after 193 nm irradiation. The laser uence,
measured with a power meter after an adjustable iris before
the photolysis window on the cryostat, ranged from 130 to 300
J cm
2
per pulse. We also vary the repetition rate from 20 to
250 Hz. We found the initial NH
3
loss rate scales linearly with
the laser uence per pulse over the range 140 to 190 J cm
2
suggesting one-photon processes are dominant. However, we
did not conduct studies over the full range of pulse energies to
rule out the possibility that two-photon fragmentation is
occurring. Typically, we vary the repetition rate to deliver a
photolysis uence in the range from 30 to 75 mW cm
2
.
High-resolution FTIR spectroscopy (0.05 cm
1
) is per-
formed on the sample using a normal incidence transmission IR
setup where the IR beam is focused through the sample and
recollimated with 2 diameter o-axis parabolic mirrors with an
8 eective focal length. FTIR spectra are recorded in one of
two ways: (1) Normal mode for high quality signal-to-noise
spectra (64 coadded scans) and (2) in rapid scan mode (e.g.,
acquisition times of 98 s for 9 scans at 0.05 cm
1
resolution)
using the minimum spectral resolution necessary. The FTIR
spectrometer (Bruker IFS 120HR) used to record IR spectra is
equipped with a glowbar source and a Ge-coated KBr
beamsplitter. We use a liquid nitrogen cooled HgCdTe
detector to record spectra from 700 to 3900 cm
1
at 0.05
cm
1
resolution and an InSb detector to record spectra from
1800 to 7000 cm
1
at 0.04 cm
1
resolution. The optical path
outside the cryostat and spectrometer is purged with dry air to
minimize atmospheric absorptions. We performed some
experiments with a 3861 cm
1
long pass lter in the FTIR
beam path to block IR light at energies above 3900 cm
1
to test
for vibron mediated processes; no evidence for these eects
were found either during or after 193 nm irradiation.

EXPERIMENTAL RESULTS AND ANALYSIS


Infrared Spectroscopy of ortho-NH
2
(X
2
B
1
). To study the
IR spectroscopy of the radical species NH
2
(X
2
B
1
) in solid pH
2
(hereafter NH
2
), we deposit a NH
3
doped pH
2
sample and wait
until nuclear spin conversion (NSC) of NH
3
to the lower
energy ortho-NH
3
spin isomer is complete. This process takes
approximately 50 min after a 1 h long deposition, and we
reported the kinetics of NH
3
NSC in a previous paper.
27
After
NSC is complete, we measure a 0.05 cm
1
resolution IR
The Journal of Physical Chemistry A Article
dx.doi.org/10.1021/jp408336n | J. Phys. Chem. A 2013, 117, 1383213842 13833
spectrum at 1.81 K with 64 coadded scans that takes
approximately 12 min to record, and this spectrum in the
region of the NH
2

2
1
11
0
00
rovibrational transition is
displayed in the bottom trace in Figure 1. The traces in Figure 1
are oset for comparison and color coded blue if the ArF laser
is o while the IR spectrum is recorded or red if the laser is on.
Further, the midpoint time (in minutes) of each IR spectrum
(time = start time + (scan time)/2) with respect to the start of
the ArF laser is indicated on the left-hand side of each trace. In
this particular 193 nm exposure used to generate the spectra in
Figure 1, the sample is irradiated with a photolysis uence of
59.2 mW cm
2
(237 J cm
2
pulse
1
, 250 Hz) for a total time
of 20 min. The bottom trace therefore is recorded 34.4 min
before the photolysis laser is started, and there is no discernible
peak in this region. The six FTIR spectra recorded during
photolysis, red traces increasing with time from bottom to top,
are recorded in rapid scan mode with 3 min scan times starting
the rst FTIR scan simultaneously with the start of the ArF
laser. With the laser on, a peak appears at around 1527.85 cm
1
with an approximately constant intensity. We assign this peak
to the N
KaKc
= 1
11
0
00
rovibrational transition of the
2
bend
vibration of ortho-NH
2
. We use the conventional asymmetric
top rotational quantum numbers N
KaKc
for a radical with one
unpaired electron (S = 1/2).
31,32
Because of the two
indistguishable H-atoms in NH
2
, each rotational level with
(K
a
,K
c
) = (even,even) or (odd,odd) has a nuclear spin weight
of 3 and is designated ortho-NH
2
while rotational levels with
(K
a
,K
c
) = (even,odd) or (odd,even) have a nuclear spin weight
of 1 and are designated para-NH
2
. The single observed
transition corresponds to ortho-NH
2
, which is a transition out
of the ground rotational state and therefore we do not expect to
observe NH
2
NSC from this state. The top three blue traces in
Figure 1 were recorded with 41 s scan times (4 coadded scans
and 0.05 cm
1
) starting the rst scan (labeled 20.3 min)
immediately after the ArF laser is shut o. As soon as the laser
is shut o, the ortho-NH
2
peak starts to decay and after 2 min is
completely gone. We estimate based on the top three scans in
Figure 1 that the ortho-NH
2
peak has a half-life of t
1/2
= 22 s
after the laser is turned o. It is important to point out this is
the decay rate of cold ortho-NH
2
. The nascent NH
2
photofragment is likely produced highly vibrationally and
rotationally excited,
8
and many of these hot NH
2
photofrag-
ments possibly go on to react with the pH
2
host; however, our
FTIR measurements are not sensitive to these processes. We
only measure the NH
2
photofragments that are successfully
cooled to the N
KaKc
= 0
00
ground rotational state. Therefore, the
ability to detect cold NH
2
during 193 nm irradiation is a
testament to the high dissipation rates
1
of solid pH
2
that very
rapidly cool the nascent NH
2
photofragment making its
detection possible.
The rapid decay of the ortho-NH
2
peak after the ArF
photolysis laser is shut o permits dierence spectra to be
constructed to unambiguously assign additional ortho-NH
2
absorption features. By generating FTIR dierence spectra
with the laser on and o, we have a powerful method to
selectively identify peaks due to ortho-NH
2
. This is
demonstrated using dierence spectra (laser on - o) in Figure
2 for the three fundamental vibrations of ortho-NH
2
. Positive
peaks are due to the production of ortho-NH
2
during
photolysis, and we detect single features for all three
fundamentals. We expect that ortho-NH
2
undergoes nearly
free rotation within the pH
2
solid analogous to NH
3
or
H
2
O.
19,20,27
Accordingly, we assign each feature to the allowed
rovibrational transition for each band from the lowest ortho-
NH
2
rotational state. For the
2
bend and
1
symmetric N-H
stretch vibrations this is the 1
11
0
00
transition (b-type) and for
the
3
asymmetric N-H stretch, it is the 1
01
0
00
transition (a-
type). Please note that the sharp negative dip in the ortho-NH
2

1
peak is due to an accidental overlap with the ortho-NH
3
2
4
a
r
R(0,0) vibration-inversion-rotation (VIR) peak which de-
creases in intensity during photolysis and thus appears as a
negative peak. The frequencies of the observed ortho-NH
2
peaks are compared to the gas phase values in Table 1.
Each of the ortho-NH
2
rovibrational peaks are quite broad
such that we do not resolve the spin doublets for these peaks
arising from the unpaired electron spin, as for example
measured in the gas phase.
31,32
Spin doublets result from the
Figure 1. IR spectra showing the ortho-NH
2

2
1
11
0
00
rovibrational
transition recorded at 1.8 K for a NH
3
doped pH
2
sample ([NH
3
]
0
=
31 ppm, [oH
2
] = 140 ppm, d = 0.23(1) cm) before, during, and after a
20 min 193 nm photolysis exposure (59.2 mW cm
2
, 250 Hz). The
spectra are color coded blue for IR spectra recorded with the ArF laser
o and red for laser on. The spectra are oset for comparison. The
time in minutes (increasing from bottom to top) of each IR scan with
respect to the start of the 193 nm photolysis is indicated on the left-
hand side of each trace. The six IR spectra recorded during photolysis
(red) have a total FTIR scan time of 3 min, and the three after
photolysis have a scan time of 41 s.
Figure 2. IR dierence spectra (laser ono) of the lowest energy
rovibrational transition of ortho-NH
2
embedded in solid pH
2
at 1.8 K
in the regions of the
2
,
1
, and
3
fundamental modes. The red lines at
the bottom of each spectral region show the corresponding gas phase
transition frequencies. See text for details.
The Journal of Physical Chemistry A Article
dx.doi.org/10.1021/jp408336n | J. Phys. Chem. A 2013, 117, 1383213842 13834
F
1
F
1
and F
2
F
2
selection rule for vibrationrotation
transitions out of the J = N + 1/2 (F
1
) and J = N 1/2
(F
2
) sublevels produced by the spin-rotation interaction.
31,32
For comparison, the transition energies of the gas phase spin
doublets are indicated by red lines at the bottom section of
each spectral region in Figure 2. On the wavenumber scale of
Figure 2 (each region covers 42 cm
1
) the gas phase spin
doublets appear as single lines. This indicates the breadth of the
ortho-NH
2
absorptions are much greater than the spin doublet
splitting, and accordingly each transition appears as a single
peak. The observed linewidths are comparable to the analogous
peaks of H
2
O solvated
19
in solid pH
2
and the broadening is
likely caused by rapid rotational relaxation or dephasing in the
excited ortho-NH
2
rovibrational state. Rovibrational transitions
(b-type) that access the 1
11
rotational level can rotationally
relax to the 0
00
level; however, transitions (a-type) that access
the 1
01
state cannot rotationally relax because the 1
01
level is the
lowest rotational level for ortho-NH
2
in the
3
= 1 excited
vibrational state. Accordingly, the a-type rovibrational tran-
sitions are typically narrower than the b-type making them
easier to detect using high resolution FTIR spectroscopy.
Further, in all three cases the peak observed for ortho-NH
2
in
solid pH
2
comes at a slightly lower transition energy than the
analogous gas phase peak indicating small decreases in the
vibrational and/or rotational constants of ortho-NH
2
solvated
in solid pH
2
. However, because we only observe a single
transition for each band, we cannot easily separate the observed
shifts into the respective vibrational and rotational contribu-
tions.
As further proof of the power of the dierence spectrum
approach for the assignment of ortho-NH
2
absorption features,
we present three additional feastures assigned to ortho-NH
2
in
the Supporting Information, Figure S1. These assignments are
conducted in the following manner. We make a vibrational
assignment by comparing the measured transition frequency
with variational calculations of the J = 0 rovibronic energies of
NH
2
on an analytic potential energy surface including the
Renner eect and spinorbit coupling.
33
We then use the
vibronic symmetry of the calculated state to predict the allowed
rovibrational transition from the 0
00
ground rotational state of
ortho-NH
2
. The strongest combination band peak is observed
at 4794.06 cm
1
with a width of 0.60 cm
1
and is assigned to
the 1
01
0
00
rovibrational peak of the
2
+
3
combination band.
The observed transition frequencies and widths of the
combination transitions are included in Table 1.
We next use the dierence spectrum approach to assign
peaks in the so-called H
2
induced region to ortho-NH
2
.
Dopant-induced IR activity and dopant-host cooperative
absorptions in solid pH
2
have been reported previ-
ously.
22,27,3437
The presence of the dopant induces the
Q
1
(0) transition (v,J = 1,0 0,0) in neighboring pH
2
molecules through intermolecular interactions that induce a
transition moment between the dopant-pH
2
pair much the
same way as in collision-induced absorption in the gas phase.
The Q
1
(0) transition is absent in neat pH
2
solids because of the
cancelation eect and thus the intensity of the dopant-
induced Q
1
(0) peak directly reects the dopant concentra-
tion.
38
The dopant-pH
2
intermolecular interaction also shifts
the vibrational frequency of the neighboring pH
2
molecules
typically to lower energy than the Q
1
(0) vibron band of solid
pH
2
. This vibrational shift breaks the coupling to the vibron
band and has the eect of making the dopant-induced Q
1
(0)
peak extremely sharp. Figure 3a shows FTIR spectra in the H
2
induced region from 4140 to 4200 cm
1
for a NH
3
doped pH
2
solid with the ArF laser on (red trace) and o (blue trace). For
this rst 193 nm irradiation of the sample, NH
3
is the most
concentrated dopant species ([NH
3
]
0
= 45 ppm) and
accordingly the H
2
induced region in the blue trace in Figure
3a is dominated by NH
3
induced peaks. When the ArF laser is
on (131 J cm
2
pulse
1
, 250 Hz), however, additional peaks
are produced which are potential ortho-NH
2
induced peaks. By
using the dierence spectrum (laser ono) generated the
same way as the one used to assign ortho-NH
2
rovibrational
peaks, we can assign the ortho-NH
2
induced Q
1
(0) feature.
This procedure allows us to pick out the ortho-NH
2
induced
Q
1
(0) peak from the more intense NH
3
induced Q
1
(0) feature.
The positive peak at 4149.07 cm
1
in Figure 3b is assigned to
the ortho-NH
2
induced Q
1
(0) peak. For comparison, we
previously assigned
27
the ortho-NH
3
induced Q
1
(0) features at
4148.73 cm
1
, and this peak is also observed in Figure 3b as a
negative dierence peak.
The broad asymmetric peak (labeled NH
3
) near 4165.52
cm
1
in Figure 3a is the Q
1
(0) pH
2
+ aR(0,0) NH
3
cooperative
transition which corresponds to simultaneous pure vibrational
Table 1. Observed Wavenumbers and Widths (FWHM) of
Transitions Assigned to ortho-NH
2
Trapped in Solid pH
2
at
1.8 K
band transition gas
a
pH
2
width shift

2
1
11
0
00
1531.35 1527.85 2.9 3.50

1
1
11
0
00
3250.54 3245.34 3.5 5.20

3
1
01
0
00
3322.32 3317.14 0.63 5.18

2
+
3
1
01
0
00
4806.66 4794.06 0.60 12.6

3
+2
2
1
01
0
00
6247.84 6238.47 0.42 9.37

1
+
3
1
01
0
00
6362.03 6385.40 0.75 15.98
a
All gas phase values are averaged over spin doublets;
2
values from
ref 31,
1
, and
3
values from ref 32, and combination bands from ref
33.
Figure 3. (a) IR spectra recorded with the laser on (red) and o
(blue) for a NH
3
doped pH
2
sample ([NH
3
]
0
= 45 ppm, d = 0.24(1)
cm) in the H
2
induced region from 4140 to 4200 cm
1
. (b) The IR
dierence spectrum (laser ono) generated from the spectra in (a)
showing positive peaks for the ortho-NH
2
-induced Q
1
(0) and Q
1
(0)
pH
2
+ 1
11
0
00
NH
2
cooperative transitions.
The Journal of Physical Chemistry A Article
dx.doi.org/10.1021/jp408336n | J. Phys. Chem. A 2013, 117, 1383213842 13835
excitation of a pH
2
molecule and inversion-rotation excitation
of NH
3
with a single IR photon.
27
Similar cooperative (photon
sharing) transitions have been observed for other small
molecules that freely rotate in solid pH
2
and again are induced
by intermolecular interactions between the dopant-pH
2
pair.
22,27,3437
Note the intensity of this peak decreases slightly
in the red trace of Figure 3a when the ArF laser is on because
the NH
3
concentration decreases because of photodissociation.
A new peak (labeled NH
2
) is observed in the red trace of
Figure 3a near 4180 cm
1
which is assigned to the Q
1
(0) pH
2
+
1
11
0
00
ortho-NH
2
cooperative transition. The peak position
of this cooperative transition provides direct information on the
pure rotational transition energy of ortho-NH
2
by subtracting
the frequency of the ortho-NH
2
induced Q
1
(0) feature. Given
the peak maxima for the cooperative transition occurs at
4179.65 cm
1
, we determine the pure 1
11
0
00
rotational
frequency of ortho-NH
2
isolated in solid pH
2
to be 30.58 cm
1
.
Rigorously, this energy dierence corresponds to the pure
rotational transition energy of an ortho-NH
2
molecule next to a
v = 1 vibrationally excited pH
2
molecule, but for our purposes it
will serve as a measure of the pure rotational transition. This
value compares favorably with the 31.8517 cm
1
gas phase
1
11
0
00
transition energy of ortho-NH
2
measured using
terahertz spectroscopy.
39
This suggests that the rotational
energy spacing decreases to 96% of the gas phase value when
ortho-NH
2
is solvated in solid pH
2
. Further, the observation of
the cooperative transition and the extracted pure rotational
energy are strong evidence that the ortho-NH
2
prolate top is
well described by the free rotor limit when isolated in solid
pH
2
.
Infrared Spectroscopy of NH(X
3

). The 193 nm
irradiated NH
3
doped pH
2
solids also show evidence for the
production of NH(X
3

), hereafter referred to as NH. Similar


to ortho-NH
2
, denitive assignment of the NH rovibrational
spectrum is dicult because of the large NH rotational
constant (2B
NH
/k
B
= 47 K) that results at liquid helium
temperatures in a one-peak rovibrational spectrum.
40,41
The
ground electronic state of NH is X
3

and therefore there are


no Q-branch transitions and we only expect to observe the
triplet R(0) transition near 3157 cm
1
. Here we label the
R(0) transition to correspond to the N = 10 rotational
transition that is split into three lines (J = S + N) corresponding
to the J = 01, 21, and 11 ne structure transitions. The
ground state is the nondegenerate N, J = (0, 1) state and the
three upper states are N, J = (1, 0), (1, 2), and (1, 1), in order
of increasing energy. With the approximate J = N selection
rule,
42
this leads to line strength factors for the three peaks of
0.34, 1.67, and 0.99, respectively.
43
We will use this expected
ne structure of the R(0) transition therefore to more
denitively assign the observed peak.
Shown in Figure 4 is an expanded view of the peak observed
after extensive (>4 h) 193 nm photolysis. The peak has a
maximum at 3154.43 cm
1
, is broad (fwhm = 2.29 cm
1
) and
slightly asymmetric. The gas phase peak positions
44
for the ne
structure resolved NH R(0) rovibrational transition are shown
on the scale at the bottom of Figure 4 labeled by J in the upper
state. Unlike ortho-NH
2
where the spin doublets were not
resolved, the ne structure splitting in NH is much larger and
thus can be detected as broadening of the NH R(0) transition.
We can model the expected ne structure in the peak by tting
the line shape. We allow the absolute frequency of all three
transitions to shift in the least-squares tting routine, but keep
the relative spacings xed at their gas phase values.
44
We
compare the relative intensities of the three tted components
with the expected line strength factors to test if such a model/
assignment makes sense. The result of such a t using a pseudo-
Voigt approximation for the line shape of each ne structure
component is shown in Figure 4 as the black line. The relative
intensities of the three components indicated by the blue lines
are, in order of increasing transition energy and scaled to the
strongest component transition, 0.34(2), 1.67(2), and 0.90(2).
We view this as a near quantitative t to the expected line
strengths for the three ne structure transitions and thus assign
the observed line shape to the NH R(0) absorption with
partially resolved ne structure.
Similar to the assignment and analysis of the ortho-NH
2
rovibrational peaks, we can use the H
2
induced region to
identify the NH induced Q
1
(0) feature and the Q
1
(0) pH
2
+
R(0) NH cooperative transition to further support this
assignment. A dierence spectrum generated from the long
time photolysis spectrum shown in Figure 4 and a spectrum
recorded before photolysis (afterbefore) is shown in Figure 5
for the H
2
induced region between 4140 and 4190 cm
1
. In this
dierence spectrum, all positive peaks are due to chemical
species that are produced by photolysis, and the negative peaks
are due to species that are consumed. The broad negative peak
near 4165 cm
1
is due to the Q
1
(0) pH
2
+ aR(0,0) NH
3
cooperative transition
27
and therefore reects the decrease in
NH
3
concentration with photolysis. The strong positive peak
near 4153.12 cm
1
is the well-known oH
2
-induced Q
1
(0)
peak,
29
which indicates that oH
2
is formed during photolysis.
The other strong positive peak near 4181 cm
1
has a triplet
structure reminiscent of the NH R(0) fundamental peak. As we
will show, we assign this peak to the Q
1
(0) pH
2
+ R(0) NH
cooperative transition. Unfortunately, we looked for peaks in
the Q
1
(0) region whose intensity scaled with this Q
1
(0) pH
2
+
R(0) NH cooperative peak, but were unable to denitively
assign a feature to the NH-induced Q
1
(0) peak. The negative
dierence peaks caused by the NH
3
-induced Q
1
(0) features
may mask an underlying NH-induced Q
1
(0) feature, or the
peak is just too weak to detect at these NH concentrations.
Figure 4. IR spectrum of the NH R(0) fundamental rovibrational
transition recorded at 1.8 K for a NH
3
doped pH
2
sample (same as
Figure 3) that was extensively photolyzed. The data are represented by
red circles. The gas phase frequencies of the three ne structure
components of this transition are indicated on the scale at the bottom
of the gure and labeled by J in the upper state. The black line is the
result of a least-squares t of the data to the sum of three pseudo-Voigt
lineshapes, and the blue lines are the individual tted lineshapes. The
peak positions of each tted component are indicated on the scale at
the top of the gure and labeled by J. See text for details.
The Journal of Physical Chemistry A Article
dx.doi.org/10.1021/jp408336n | J. Phys. Chem. A 2013, 117, 1383213842 13836
We can analyze the Q
1
(0) pH
2
+ R(0) NH cooperative peak
further using a line shape analysis similar to the one applied to
the NH R(0) fundamental transition. A blow-up of the Q
1
(0)
pH
2
+ R(0) NH cooperative peak along with the tted pseudo-
Voigt line shape for three independent components are shown
in the Supporting Information, Figure S2, and the tted
parameters are presented in Table 2. Unlike the t to the NH
R(0) fundamental, in this case the frequencies (
i
), linewidths
(b
i
), fractional Lorentzian contribution (c
i
), and intensities (A
i
)
could all be tted separately because of the partial resolution of
all three peaks. We again observe quantitative agreement in the
ne structure compared to the N = 10 pure rotational
transition in the gas phase;
45
the peak splitting (0.93(2) and
0.87(1) cm
1
in pH
2
versus 0.934 and 0.867 cm
1
in the gas
phase) and relative line strengths (0.36(6):1.66(14):1.2(22)) of
the three components match the gas phase values. This
indicates the spin-spin and spin-rotation interactions for NH
isolated in solid pH
2
are quantitatively the same as the gas
phase. We did not observe a NH-induced Q
1
(0) feature, so we
cannot translate the observed cooperative peak frequencies into
pure rotational energies. However, if we approximate the NH-
induced Q
1
(0) transition frequency using the frequency of the
NH
3
-induced Q
1
(0) peak, we estimate the J = 01, 21, and
11 ne structure transitions come at 31.61, 32.53, and 33.41
cm
1
for NH solvated in solid pH
2
which are all within 0.05
cm
1
from the corresponding gas phase frequencies
45
of 31.56,
32.49, and 33.36 cm
1
.
Assignment of NH
3
Satellite Peaks. Now that the NH
2
and NH species have been characterized spectroscopically, we
can study the 193 nm in situ photolysis of NH
3
in more detail.
Shown in Figure 6 are spectra in the NH
3

2
umbrella mode
region recorded before, during, and after a 193 nm exposure
(286 J cm
2
pulse
1
, 100 Hz, 10.8 min) of a NH
3
doped pH
2
sample. The bottom blue trace is recorded before photolysis
and shows the strong aR(0,0) NH
3
VIR peak for the freely
rotating monomer.
27
Weaker features to the red of this intense
feature are likely due to NH
3
complexed with oH
2
molecules in
the sample, but a detailed assignment was not conducted. The
middle red trace in Figure 6 is recorded (16 coadded scans, 178
s scan time) during photolysis and shows slight growth of para-
NH
3
as evidenced by the weak aQ(1,1) peak. The production
of ortho-NH
2
is also observed via the
2
1
11
0
00
peak at
1527.85 cm
1
. In addition, we observe the appearance of two
new peaks in the NH
3

2
region at slightly lower and higher
energies of the strong aR(0,0) ortho-NH
3
peak. These features
have relatively large peak absorption maxima, and we will label
them S1 and S2 for convenience. The top blue trace in Figure 6
is recorded (16 coadded scans, 178 s scan time) starting
immediately after the 193 nm laser is shut o. This trace shows
that the ortho-NH
2
peak almost completely disappears at this
time resolution (3 min), while the aQ(1,1) NH
3
peak decays
more slowly via NSC with a time constant ( = 8.9(7) min at
1.8 K) that was measured previously.
27
However, note that
both the S1 and S2 peaks signicantly increase in intensity right
after the laser is stopped. After this initial fast increase in the
intensities of S1 and S2, these peaks then decay while the
sample is held at 1.8 K.
We assign these new S1 and S2 peaks as satellite peaks of the
strong ortho-NH
3
aR(0,0) transition that are due to ortho-NH
3
produced in a defect site. Specically, some of the ortho-NH
2
that is produced during photolysis reacts via reaction 3 with
one of its nearest neighbor pH
2
molecules creating a vacancy or
H-atom next to the nascent ortho-NH
3
. Normally the ortho-
NH
3
molecule occupies a single substitution site and the
2
aR(0,0) transition is a single relatively sharp peak (see Figure
6).
27
However, for the ortho-NH
3
that is created in situ via
reaction 3 the resulting solvation environment of the defect site
Figure 5. IR dierence spectrum (photolysisbefore) in the region
from 4140 to 4190 cm
1
showing the changes in the spectrum
produced by extensive 193 nm photolysis of a NH
3
doped pH
2
sample
(same sample as Figure 3). Positive peaks for the oH
2
-induced Q
1
(0)
and Q
1
(0) pH
2
+ R(0) NH cooperative transition indicate both oH
2
and NH are being produced by photolysis. The negative Q
1
(0) pH
2
+
aR(0,0) NH
3
peak indicates NH
3
is being consumed by photolysis.
Table 2. Fit Parameters for the Q
1
(0) pH
2
+ R(0) NH
Cooperative Peak Recorded at 1.8 K
a
parameter N,J = 1,0 N,J = 1,2 N,J = 1,1
A
0
0.0065(10) 0.0300(24) 0.021(39)
b
0
0.604(53) 0.698(15) 0.747(20)
c
0
0.94(11) 0.80(11) 1.0(19)
v
0
4180.335(16) 4181.2649(28) 4182.1392(70)
a
Fit equation: f = 2c
0
A
0
/(b
0
)(1/(1 + 4(v v
0
)/b
0
)
2
)+(1 c
0
)A
0
/
b
0
(4ln(2)((v v
0
)/b
0
)
2
).
Figure 6. Three spectra recorded during a 193 nm photolysis
experiment (28.6 mW cm
2
, 10.8 min) on a NH
3
doped pH
2
sample.
Spectra are shown in the region of the para-NH
3

2
aQ(1,1) transition,
ortho-NH
3

2
aR(0,0) transition, and the ortho-NH
2

2
1
11
0
00
transition with the intensity in this region multiplied by four. Spectra
are oset and color coded blue for laser o and red for laser on. The
bottom trace was recorded at 1.8 K before photolysis, the middle red
trace was recorded during photolysis (scan time = 178 s), and the top
blue trace was recorded immediately after photolysis (scan time = 178
s) at 1.8 K. Note the two peaks labeled S1 and S2 that appear during
photolysis actually increase right after the laser is turned o (top blue
trace). See text for details.
The Journal of Physical Chemistry A Article
dx.doi.org/10.1021/jp408336n | J. Phys. Chem. A 2013, 117, 1383213842 13837
splits the aR(0,0) transition into the two closely spaced S1 and
S2 satellite peaks. The rotational potential felt by an NH
3
molecule adjacent to a vacancy or H-atom (we cannot
distinguish) has lower symmetry and lifts the M
J
degeneracy
of the NH
3
upper J, K = 1, 1 VIR state. Direct spectral evidence
that links the ortho-NH
2
species to the satellite peaks is shown
in Figure 6 by the fast decrease in the ortho-NH
2
peak that is
correlated with the initial increase in the S1 and S2 peaks right
after the photolysis laser is stopped. We include plots of the
intensities of peaks for NH, S1+S2, and ortho-NH
2
with time
for the experiment shown in Figure 1 in the Supporting
Information, Figure S3. During photolysis ortho-NH
3
in single
substitution sites and defect sites are both undergoing
photodissociation, and the S1 and S2 peak intensities reach a
constant value; NH
3
in defect sites are produced by reaction 3
but are also consumed by NH
3
photodissociation. When the
laser is turned o, this NH
3
photodissociation loss mechanism
is stopped, but the photoproduced ortho-NH
2
continues to
react with the host for a short period of time resulting in the
rapid initial increase in the satellite peak intensities. After all the
ortho-NH
2
is consumed (roughly 3 min), the S1 and S2
satellite peaks stop increasing and then start to slowly decay as
the NH
3
molecules in defect sites transform back into single
substitution sites.
Now by using dierence spectra we have been able to assign
satellite peaks for the
2
,
4
, and
3
fundamental modes as well
as the
1
+
2
and
3
+
4
combination bands. Representative
spectra and a table of the peak positions are presented in the
Supporting Information. In all cases, two satellite peaks are
observed with one at lower and one at higher wavenumbers of
the dominant aR(0,0) or a
r
R(0,0) ortho-NH
3
VIR feature for
that specic band. This is consistent with our assignment of the
dominant cause of the splitting as rotational and not vibrational
because the shifts from the monomer feature remain nearly
constant (see Supporting Information, Table S1). However, the
satellite shifts for bands that involve the
2
vibration behave
anomalously because of the large increase in the inversion
tunneling splitting for excitation of this mode. This also
supports the ortho-NH
3
assignment as the carrier of the
satellite peaks because it is highly unlikely that some species
would accidently overlap in so many regions of the spectrum.
The intensities of the two satellite features are highly positively
correlated suggesting that both transitions share a common
lower state for the same species. Finally, the sum of the
integrated intensities of these two positive satellite peaks in the
dierence spectra semiquantitatively match the intensity of the
negative ortho-NH
3
monomer peak indicating a spectroscopic
stability. This spectroscopic stability results from the fact that
the satellite peaks and the monomer peak are all for the same
spectral carrier and transition, but just for two dierent
solvation environments.
Similar satellite peaks are observed near the H
2
O
3
1
01

0
00
rovibrational transition for formic acid (HCOOH) doped
pH
2
solids that are exposed to 193 nm laser irradiation.
46,47
We
assigned these satellite peaks to H
2
O produced in radiation
damaged substitution sites that are produced by reaction of the
hydroxyl photoproduct (HCOOH + h HCO + OH) with
the pH
2
matrix. This in situ photochemical reaction produces
H
2
O in a single substitution site adjacent to either a vacancy or
a H-atom, which splits the a-type 1
01
0
00
rovibrational
transition into two satellite peaks.
46,47
However, in these
previous studies we were never able to detect the OH
intermediate using rapid scan FTIR. Presumably the OH +
H
2
H
2
O + H reaction occurs too fast under these conditions
to detect the OH intermediate using rapid scan FTIR. We show
here a similar photochemical mechanism is occurring for NH
3
photolysis, but in this case the NH
2
+ H
2
NH
3
+ H reaction
is slower (or production of ortho-NH
2
faster) such that we can
detect the ortho-NH
2
chemical intermediate while the sample is
being irradiated.
Mechanistic Role of NH
3
in Defect Sites. Once ortho-
NH
3
is created in a defect site it undergoes slow back
conversion to a single substitution site even at 1.8 K. To
demonstrate this back conversion process we show in Figure 7
a plot of the integrated intensities of the ortho-NH
3

2
aR(0,0)
transition and the summed intensities (S1+S2) of the ortho-
NH
3

2
satellite peaks. This particular photolysis experiment
was conducted early in our investigation, and the photolysis
conditions (280 J cm
2
, 20 Hz, 42 min) were at a lower laser
uence (5.6 mW cm
2
) than later experiments. As can be seen
in Figure 7, once the photolysis laser is turned on the sample
quickly reaches a photoequilibrium between ortho-NH
3
molecules in defect (S1+S2) and single substitution (ortho-
NH
3
) sites. There is approximately a 30% depletion in the
ortho-NH
3

2
aR(0,0) peak intensity during photolysis. For this
experiment there are no detectable ortho-NH
2
peaks produced
during irradiation, and the NH signal is very small after the full
42 min photolysis is complete. Once the laser is turned o,
there is a slight increase in the S1+S2 satellite intensity, but the
increase is small because of the low steady-state concentration
of ortho-NH
2
produced during photolysis at this photolysis
uence. We show the behavior for high photolysis uence in
the Supporting Information, Figure S8. After photolysis while
the sample is maintained at 1.81(2) K, the data in Figure 7
show the S1+S2 intensity decreases and the ortho-NH
3
aR(0,0)
peak increases. Also shown in Figure 7 are the results of least-
squares ts of the data to biexponential and single exponential
expressions for growth and decay, respectively. The equations
and tted parameters are presented in the Supporting
Information, Table S2. The biexponential growth of NH
3
reects fast NSC of a small amount of photoproduced para-
NH
3
and slow back conversion of ortho-NH
3
produced in
Figure 7. Kinetic plots of the integrated intensity of the ortho-NH
3

2
S1+S2 satellite peaks (red circles), the ortho-NH
3

2
aR(0,0)
monomer peak (blue circles), and the sum of these features (total,
black circles) for a NH
3
doped pH
2
sample ([NH
3
]
0
= 13 ppm, d =
0.28(2) cm). The timing and duration of a 193 nm photolysis (42 min,
5.6 mW cm
2
, 20 Hz) at 1.8 K is indicated by the yellow shaded area.
The lines represent least-squares ts of the data to biexponential
growth and single exponential decay curves.
The Journal of Physical Chemistry A Article
dx.doi.org/10.1021/jp408336n | J. Phys. Chem. A 2013, 117, 1383213842 13838
defect sites (S1+S2) to single substitution sites. The decay of
the S1+S2 integrated intensity is well modeled with a single
exponential expression except at the earliest times after
photolysis because of a fast initial increase of S1+S2 due to
production of ortho-NH
3
in defect sites via reaction 3.
Comparison of the slower rate constant extracted from the t
to ortho-NH
3
growth (k
2
= 6.66(95) 10
3
min
1
) with the
decay constant for S1+S2 (k
2
= 9.68(28) 10
3
min
1
) shows
that both conversion processes occur with approximately the
same rate constant.
Further support for the ortho-NH
3
defect to single
substitution conversion process is provided by the relative
changes in the integrated intensities of the satellite peaks and
the NH
3
monomer peak. We assert that the carriers of both
absorption signals are ortho-NH
3
; for ortho-NH
3
solvated in a
single substitution site we observe the single aR(0,0) transition
and for the defect site the transition is split into the S1 and S2
satellite peaks. For such a scenario there should be a
spectroscopic stability between the intensity of the sum of
the satellite peaks S1+S2 and the aR(0,0) transition. This is
what is observed; the black circles in Figure 7 are the total
intensity of the satellite peaks and the aR(0,0) NH
3
monomer
peak. As can be seen, after a slight increase immediately after
photolysis is stopped, the total remains relatively constant as
the satellite peaks decay and the NH
3
aR(0,0) peak grows. We
speculate that the small rise in the total NH
3

2
intensity right
after photolysis is due to NSC of the small amount of para-NH
3
produced during photolysis. From the ts of the data in Figure
7 we extract an intensity change of 0.2052(32) cm
1
for
S1+S2 compared to a +0.2646(76) cm
1
increase in the NH
3
aR(0,0) intensity. The roughly 20% discrepancy is likely due to
inaccuracies in the measured integrated intensities due to
overlap of the S1 and S2 features (mostly S2) with the NH
3

2
aR(0,0) monomer feature.
On the basis of the intensities of the satellite peaks right after
photolysis and then at longer times, we can piece together a
mechanism. The rapid decay of the ortho-NH
2
peaks after
photolysis is stopped is linked to the rapid increase in the
satellite peaks. The slower back conversion process couples the
satellite peaks to the ortho-NH
3
monomer peaks as well. We
therefore propose the following mechanism,
+
+
+
ortho NH pH
ortho NH (defect) H
ortho NH (single substitution) H
2
2
3
3
(5)
During irradiation ortho-NH
2
is produced via photodissociation
of ortho-NH
3
. Nuclear spin angular momentum is conserved
during the photodissociation, and only the ortho-NH
2
nuclear
spin isomer can be formed by removal of one H-atom from
ortho-NH
3
. The nascent ortho-NH
2
relaxes rapidly (on the
order of s) in solid pH
2
to the ground vibrational and
rotational level as evidenced by the many ortho-NH
2
transitions
observed out of the ground rotational level during irradiation.
Once the ortho-NH
2
is relaxed to the ground rotational state it
can either undergo (1) photodissociation or (2) can react with
the pH
2
host to reform ortho-NH
3
in a defect site. Now in this
case, nuclear spin conservation
28
requires that ortho-NH
2
(I =
1) and pH
2
(I = 0) should react to produce 0.66 ortho-NH
3
(I =
3/2) and 0.33 para-NH
3
(I = 1/2). However, we only detect
satellite peaks centered around the ortho-NH
3
aR(0,0)
transition. This suggests that reaction 3 via mechanism 5
results almost exclusively in ortho-NH
3
contrary to the
predictions based on nuclear spin conservation.
28
The small
production of para-NH
3
is not linked to the satellite peaks
because the intensity of the aQ(1,1) peak does not increase
right after the photolysis laser is shut o. Then with a slower
time constant, the newly produced ortho-NH
3
in defect sites
slowly converts back to single substitution sites.
These ndings have important ramications for the photo-
production of NH in solid pH
2
. Clearly for the photolysis
experiment shown in Figure 7 the conversion of NH
3
to NH is
not ecient. Even though there is roughly a 30% decrease in
the ortho-NH
3
intensity during photolysis, there is complete
recovery of the ortho-NH
3
signal with time as the intensity in
the satellite peaks is converted back to the monomer peak.
Indeed, the amount of ortho-NH
3
as measured by the intensity
of the
2
aR(0,0) peak actually increases slightly compared to
before photolysis. This slight increase is due to the photo-
conversion of ortho-NH
3
molecules in nonrotating sites
(potentially caused by (NH
3
)
n
or NH
3
(oH
2
)
n
clusters, the
small peaks in the bottom trace of Figure 7) into freely rotating
monomer sites. Nonetheless, this analysis shows that for
ecient conversion of ortho-NH
3
to NH the UV laser uence
must be signicantly greater to photodissociate the ortho-NH
2
faster than it reacts with the pH
2
host.
In Situ Photoproduction of NH. Given our mechanistic
understanding of the NH
3
in situ photolysis, we set out in a last
experiment to see how much NH could be produced using high
ArF repetition rates. We document the conversion process in
Figure 8 for this NH
3
doped pH
2
sample that was repeatedly
irradiated. As shown in Figure 8, the sample was photolyzed 7
times and all with high repetition rates (250 Hz) and laser
uences (131 J cm
2
pulse
1
). Monitoring the integrated
intensity of the NH R(0) transition throughout the process one
can see that at early photolysis times the NH growth rate is the
greatest. This behavior is due to the photokinetics of sequential
Figure 8. Kinetic plots of the integrated intensity of the NH R(0),
NH
2

2
+
3
1
01
0
00
, and oH
2
-induced Q
1
(0) absorption peaks as a
function of time for a NH
3
doped pH
2
sample at 1.8 K (same sample
as Figure 3). The timing and duration of 7 distinct 193 nm irradiations
of the same sample are indicated by the yellow shaded regions. The
photolysis conditions are as follows: 10, 20, 40, 40, and 80 min at 131
J cm
2
pulse
1
and 250 Hz followed by 20 and 40 min at 237 J cm
2
pulse
1
and 250 Hz.
The Journal of Physical Chemistry A Article
dx.doi.org/10.1021/jp408336n | J. Phys. Chem. A 2013, 117, 1383213842 13839
two-photon photodissociation and reaction 3 of the photo-
generated ortho-NH
2
with the pH
2
host. Even at these high
repetition rates reaction 3 severely limits the number of NH
2
molecules that undergo the second photolysis step to produce
NH. Thus, at early times when the NH
3
concentration in single
substitution sites is the greatest, we observe the steepest growth
in the NH peak. However, as the NH
3
concentration is
depleted and greater amounts of NH
3
occupy defect sites, less
and less NH is produced. Further, because of reaction 3 we can
never fully convert the NH
3
to NH because once the ortho-
NH
2
production rate becomes less than the rate of reaction 3,
the production of NH eectively stops.
Another interesting nding is shown by the growth in the
oH
2
signal. The oH
2
growth follows the NH growth curve
indicating rst that oH
2
is only produced when the photolysis
laser is on and second that the photochemical processes that
ultimately produce NH are strongly correlated to the processes
that produce oH
2
. The processes that produce oH
2
are
photochemical and require laser excitation of the NH
3
dopant.
That is, the reactions that produce oH
2
must involve hot H-
atom reactions. The oH
2
is not produced by cold H-atom
recombination reactions otherwise we would observe growth in
the oH
2
peak after a given irradiation step (when the laser is
turned o). The integrated intensity of the oH
2
-induced Q
1
(0)
transition is well-known such that the integrated intensity
shown in Figure 8 can be translated into a concentration. Using
the integrated intensity of the Q
1
(0) transition for the last three
points displayed in Figure 8, we determine [oH
2
] = 3050(100)
ppm. Given that this sample had an initial NH
3
concentration
of 45 ppm, such a high oH
2
concentration must result from
repeated (65-fold) photolysis of the NH
3
precursor in some
way. Thus, it is clear that reaction 3 plays a pivotal role in the in
situ photochemistry of NH
3
. We can use the measured rate
constant (k
tun
= 1.88(17) min
1
) for reaction 3 combined with
the density of solid pH
2
at liquid helium temperatures (2.600
10
22
molecules cm
3
) to calculate a second-order rate constant
for reaction 3. This procedure predicts a value of k (T = 1.8 K)
= 1.2(1) 10
24
cm
3
molecule
1
s
1
which is quite small by gas
phase standards, but leads to facile decay of ortho-NH
2
in solid
pH
2
. The facile decay is because the reaction partners are held
in constant contact and thus the chemical reaction is
transformed into a pseudo-rst-order process.
48
We write this
as k
tun
because reaction 3 has a sizable barrier and thus is a
tunneling-driven reaction. This also points to an important
distinction from gas phase reactions; namely, even though this
reaction has an exceedingly small eective second-order rate
constant the ortho-NH
2
radical only has a half-life of 22 s in
solid pH
2
at 1.8 K because the reaction partners are held in
constant contact. Also clear in Figure 8 is that NH
2
is only
detected during irradiation and with increased exposures the
amount of NH
2
produced during photolysis decreases. We
predict therefore that the generation of NH could be increased
substantially if considerably greater ArF repetition rates could
be used at early photolysis times.
All our analysis indicates that reaction 3 is playing a key role
in the in situ photokinetics of NH
3
at 193 nm. We have argued
that the main reason for this is bimolecular reaction 3 is
transformed into a pseudo-rst order reaction by the fact that
the NH
2
is produced in solid pH
2
. However, the importance of
the NH
2
+H NH
3
reaction in the measured photokinetics
(during and after irradiation) is less clear. We have shown the
rapid decay in the NH
2
peaks after the laser is shut o
correlates strongly with the rise time of the NH
3
satellite peaks,
which is why we assign the decay of NH
2
to reaction 3. It is the
reaction of the NH
2
with a neighboring pH
2
that is necessary to
create the product NH
3
molecule in a defect site. We feel the
NH
2
+H reaction is not competitive under these conditions
because it is bimolecular and the rate depends on the
concentration and mobility of the H-atoms. However, the
measured rate constant for the decay of NH
2
is approximately
10
4
times larger than the k
tun
= 2.0(2) 10
4
min
1
measured
for the CD
3
+ H
2
CD
3
H + H reaction in solid pH
2
at 5
K.
48,49
These two reactions have similar activation energies and
exothermicities and thus without detailed analysis of the barrier
widths and multidimensional tunneling dynamics for both
reactions, it is dicult to rationalize such a large dierence in
rate constants. Further, in the case of CD
3
H there is no
evidence from detailed rovibrational analysis
50
that the
molecule is created in a defect site similar to NH
3
. It could
be that NH
2
+ H NH
3
also leads to the production of ortho-
NH
3
in defect sites because the mobile H-atom occupies a
single substitution site
51
that remains after reaction. This
additional mechanism might therefore explain the larger rate
constant measured here for reaction 3 compared to the CD
3
+
H
2
CD
3
H + H reaction and the preferential production of
ortho-NH
3
which is not predicted for nuclear spin conservation
via reaction 3.
Another observation is that once the NH is produced, it is
remarkably stable. We observe no evidence of reaction of NH
with the pH
2
matrix or with H-atoms after photolysis. This
second point is surprising because in photolysis studies of other
precursor molecules (e.g., NO) that produce H-atoms, these
researchers observed cold H-atom reactions with the NO that
occurred long after the photolysis laser is stopped.
17
Possibly
the growth in the oH
2
concentration is contributing to this
strange behavior. During irradiation NH
3
and NH
2
are being
photodissociated and producing hot H-atoms that may then
recombine to generate oH
2
. Given the high concentrations of
oH
2
that are generated and that oH
2
is only produced during
irradiation, these observations may explain why we do not
observe cold H-atom reactions with the NH that is produced
even though this reaction is barrierless and the H-atoms are
mobile at these temperatures via chemical diusion.
15
Possibly
the concentration of free H-atoms are lower because of hot H
+ H H
2
reactions that eciently form oH
2
. At the end of the
experiment shown in Figure 8, we raised the temperature of the
sample to 4.3 K and observe rapid diminution in the measured
NH peak and growth of peaks we can tentatively assign to
N
2
H
2
.
52
We will report these reaction experiments in a separate
publication.

SUMMARY AND OUTLOOK


Our results for the in situ photolysis of NH
3
provide some
important lessons for studying the photochemistry of precursor
molecules in pH
2
quantum solids. Our ability to detect a
steady-state concentration of NH
2
during irradiation shows that
a signicant fraction of the photogenerated NH
2
radicals are
rapidly cooled to the lowest vibrational and rotational state.
Then on a much longer time scale, these cold NH
2
molecules
react with the pH
2
host. We argue that the decay in the NH
2
concentration after the laser is stopped is due to reactions with
the pH
2
host that occur via pseudo-rst-order kinetics rather
than NH
2
+ H reactions that are bimolecular. Regardless, when
the NH
2
radical reacts it creates ortho-NH
3
in a defect site. The
exact nature of the defect site is still unclear; however, the
presence of NH
3
in this defect site is conclusively detected
The Journal of Physical Chemistry A Article
dx.doi.org/10.1021/jp408336n | J. Phys. Chem. A 2013, 117, 1383213842 13840
using FTIR spectroscopy. We speculate that this defect site
results from the NH
2
reacting with the pH
2
host and thereby
chemically removing one of the nearest neighbor pH
2
molecules. With time the NH
3
molecules in defect sites are
transformed back into the more thermodynamically stable
single substitution site. This self-annealing of the defect sites
occurs even at 1.8 K and therefore cannot be prevented. We
can monitor this solvent relaxation either by the decay in the
signal from NH
3
molecules in defect sites (satellite peaks) or as
the growth in the NH
3
signal in single substitution sites (NH
3
monomer peaks). The detail with which we can monitor this
entire process using FTIR spectroscopy is truly remarkable.
However, these studies also raise interesting questions about
the role of H-atoms and/or vacancies in the measured reaction
kinetics. We cannot completely characterize the defect solvent
geometry because of the indirect structural information that
FTIR spectroscopy provides. We need to couple these
experimental measurements with theoretical models to try to
gure out the true nature of the defect site.
The goal to generate high concentrations of NH in solid pH
2
and characterize their stability was partially achieved. As we
have shown, if high repetition rates are employed the NH
3
can
be transformed into NH via in situ photochemistry. However,
the yield of NH is substantially reduced by tunneling reactions
of the cold NH
2
radical with the pH
2
host. Once produced the
NH radical is remarkably stable with respect to reactions with
the pH
2
host. This kinetic stability is expected because the NH
radical in a triplet state is spin forbidden from reacting with pH
2
to produce singlet NH
3
. We speculate that if even greater
repetition rates (1000 Hz ArF lasers are commercially available)
are utilized then higher conversion eciencies could be
achieved. Still in the one-photon regime, these higher repetition
rates would produce a greater probability of sequential two-
photon photodissociation and help to limit the number of NH
2
radicals that back react with the pH
2
host. What we nd curious
about the NH
3
/pH
2
photochemical system is the lack of
evidence of H-atom reactions after photolysis. In many of the
elementary steps revealed in this study H-atoms are produced
and yet we nd no evidence of H-atom reactions with the NH
radical, for example. Further, we nd that for samples that have
been extensively photolyzed, signicant concentrations of oH
2
are produced. Given the amount of oH
2
that is produced
compared with the initial NH
3
concentration, this is only
possible if the NH
3
is repeatedly cycled through photolysis.
This leads us to speculate that during irradiation a signicant
number of hot H-atom reactions are occurring which may
ultimately limit the number of free H-atoms that remain after
the laser is turned o.
Lastly, we want to emphasize that these studies uncover
fundamental unanswered questions about the microscopic
details of the in situ photochemistry of precursor molecules
in solid pH
2
. For example, what is the typical hot H-atom
trajectory after being produced by photodissociation of a
precursor molecule? We speculate that the in situ chemical
reactions of a photoproduct with the pH
2
host typically result
in the generation of cold H-atoms and/or vacancies. What role
if any do cold H-atoms and vacancies play in the chemistry
during and after irradiation? We believe that only when some of
these basic questions are answered can we begin to rationally
design conditions for future cryochemistry studies in solid pH
2
.

ASSOCIATED CONTENT
*S Supporting Information
Additional spectra of ortho-NH
2
combination band peaks, an
expanded view of the Q
1
(0) pH
2
+ R(0) NH cooperative
transition showing resolved ne structure, a kinetic plot of the
NH, ortho-NH
3
satellite, and ortho-NH
2
peak intensities
during and immediately after 193 nm photolysis, IR dierence
spectra of NH
3
satellite peaks for
4
,
3
,
1
+
2
, and
3
+
4
bands,
a table of satellite peak positions and shifts, a kinetic plot of the
ortho-NH
3

4
a
r
R(0,0) and S1+S2 satellite peaks during and
after irradiation, and a table of the tted kinetic parameters for
the ortho-NH
3

2
aR(0,0) and S1+S2 satellite peaks for the
data in Figure 7. This material is available free of charge via the
Internet at http://pubs.acs.org.

AUTHOR INFORMATION
Corresponding Author
*E-mail: danderso@uwyo.edu
Notes
The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
The authors thank the Chemistry Division at the National
Science Foundation for its generous support through Grant
CHE 08-48330.

REFERENCES
(1) Momose, T.; Shida, T. Matrix-Isolation Spectroscopy using Solid
Parahydrogen as the Matrix: Application to High-Resolution Spec-
troscopy, Photochemistry, and Cryochemistry. Bull. Chem. Soc. Jpn.
1998, 71, 115.
(2) Momose, T.; Fushitani, M.; Hoshina, H. Chemical Reactions in
Quantum Crystals. Int. Rev. Phys. Chem. 2005, 24, 533552.
(3) Yoshioka, K.; Raston, P. L.; Anderson, D. T. Infrared
Spectroscopy of Chemically Doped Solid Parahydrogen. Int. Rev.
Phys. Chem. 2006, 25, 469496.
(4) Bahou, M.; Huang, C. W.; Huang, Y. L.; Glatthaar, J.; Lee, Y. P.
Advances in Use of p-H
2
as a Novel Host for Matrix IR Spectroscopy.
J. Chin. Chem. Soc. 2010, 57, 771782.
(5) Fajardo, M. E.: Matrix Isolation Spectroscopy in Solid
Parahydrogen: A Primer. In Physics and Chemistry at Low Temper-
atures; Khriachtchev, L., Ed.; Pan Stanford Publishing Pte. Ltd.:
Singapore, 2011; pp 167202.
(6) Kenner, R. D.; Rohrer, F.; Browarzik, R. K.; Kaes, A.; Stuhl, F.
Two-Photon Formation of NH/ND(A
3
) in the 193 nm Photolysis
of Ammonia. I. Mechanism and Identification of the Intermediate
Species. Chem. Phys. 1987, 118, 141152.
(7) Kenner, R. D.; Browarzik, R. K.; Stuhl, F. Two-photon Formation
of NH/ND(A
3
) in the 193 nm Photolysis of Ammonia. II.
Photolysis of NH
2
. Chem. Phys. 1988, 121, 457471.
(8) Koplitz, B.; Xu, Z.; Wittig, C. Product Energy Distributions from
the 193 nm Photodissociation of NH
3
. Chem. Phys. Lett. 1987, 137,
505509.
(9) Sato, H. Photodissociation of Simple Molecules in the Gas Phase.
Chem. Rev. 2001, 101, 26872725.
(10) Espinosa-Garcia, J.; Corchado, J. C. Analytical Potential Energy
Surface and Kinetics of the NH
3
+H NH
2
+H
2
Hydrogen
Abstraction and the Ammonia Inversion Reactions. J. Phys. Chem. A
2010, 114, 44554463.
(11) Linder, D. P.; Duan, X.; Page, M. Ab Initio Variational
Transition State Theory Calculations for the H + NH
2
H
2
+ NH
Hydrogen Abstraction Reaction on the Triplet Potential Energy
Surface. J. Phys. Chem. 1995, 99, 1145811463.
(12) Corchado, J. C.; Espinosa-Garcia, J.; Yang, M. Kinetics and
Dynamics of the NH
3
+ H NH
2
+ H
2
Reaction Using Transition
The Journal of Physical Chemistry A Article
dx.doi.org/10.1021/jp408336n | J. Phys. Chem. A 2013, 117, 1383213842 13841
State Methods, Quasi-Classical Trajectories, and Quantum-Mechanical
Scattering. J. Chem. Phys. 2011, 135, 014303.
(13) Fueno, T.; Bonacic-Koutecky, V.; Koutecky, J. Ab Initio CI
Study of Chemical Reactions of Singlet and Triplet NH Radicals. J.
Am. Chem. Soc. 1983, 105, 55475557.
(14) Anderson, W. R. Oscillator Strengths of NH
2
and the Heats of
Formation of NH and NH
2
. J. Phys. Chem. 1989, 93, 530536.
(15) Kumada, T. Experimental Determination of the Mechanism of
the Tunneling Diffusion of H Atoms in Solid Hydrogen: Physical
Exchange Versus Chemical Reaction. Phys. Rev. B 2003, 68, 052301.
(16) Hidaka, H.; Watanabe, M.; Kouchi, A.; Watanabe, N. FTIR
Study of Ammonia Formation Via the Successive Hydrogenation of N
atoms Trapped in a Solid N
2
Matrix at Low Temperatures. Phys. Chem.
Chem. Phys. 2011, 13, 1579815802.
(17) Fushitani, M.; Momose, T. A Study on Diffusion of H Atoms in
Solid Parahydrogen. Low Temp. Phys. 2003, 29, 740743.
(18) Tam, S.; Fajardo, M. E.; Katsuki, H.; Hoshina, H.; Wakabayashi,
T.; Momose, T. High Resolution Infrared Absorption Spectra of
Methane Molecules Isolated in Solid Parahydrogen Matrices. J. Chem.
Phys. 1999, 111, 41914198.
(19) Fajardo, M. E.; Tam, S.; DeRose, M. E. Matrix Isolation
Spectroscopy of H
2
O, D
2
O, and HDO in Solid Parahydrogen. J. Mol.
Struct. 2004, 695, 111127.
(20) Fajardo, M. E.; Lindsay, C. M. Crystal Field Splitting of
Rovibrational Transitions of Water Monomers Isolated in Solid
Parahydrogen. J. Chem. Phys. 2008, 128, 014505.
(21) Fajardo, M. E.; Lindsay, C. M.; Momose, T. Crystal Field
Theory Analysis of Rovibrational Spectra of Carbon Monoxide
Monomers Isolated in Solid Parahydrogen. J. Chem. Phys. 2009, 130,
244508.
(22) Anderson, D. T.; Hinde, R. J.; Tam, S.; Fajardo, M. E. High-
Resolution Spectroscopy of HCl and DCl Isolated in Solid
Parahydrogen: Direct, Induced, and Cooperative Infrared Transitions
in a Molecular Quantum Solid. J. Chem. Phys. 2002, 116, 594607.
(23) Paulson, L. O.; Anderson, D. T. High-Resolution Vibrational
Spectroscopy of trans-Formic Acid in Solid Parahydrogen. J. Phys.
Chem. A 2009, 113, 17701778.
(24) Lee, Y. P.; Wu, Y. O.; Lees, R. M.; Xu, L. H.; Hougen, J. T.
Internal Rotation and Spin Conversion of CH
3
OH in Solid Para-
hydrogen. Science 2006, 311, 365368.
(25) Lee, Y. P.; Wu, Y. J.; Hougen, J. T. Direct Spectral Evidence of
Single-Axis Rotation and Ortho-Hydrogen-Assisted Nuclear Spin
Conversion of CH
3
F in Solid Para-Hydrogen. J. Chem. Phys. 2008,
129.
(26) Lorenz, B. D.; Anderson, D. T. Infrared Spectra of N
2
O-(ortho-
D
2
)
N
and N
2
O-(HD)
N
Clusters Trapped in Bulk Solid Parahydrogen.
J. Chem. Phys. 2007, 126, 184506.
(27) Ruzi, M.; Anderson, D. T. Matrix Isolation Spectroscopy and
Nuclear Spin Conversion of NH
3
and ND
3
in Solid Parahydrogen. J.
Phys. Chem. A 2013, 117 (39), 97129724.
(28) Fushitani, M.; Momose, T. Nuclear Spin Selection Rule in the
Photochemical Reaction of CH
3
in Solid Parahydrogen. J. Chem. Phys.
2002, 116, 1073910743.
(29) Tam, S.; Fajardo, M. E. Ortho/Para Hydrogen Converter for
Rapid Deposition Matrix Isolation Spectroscopy. Rev. Sci. Instrum.
1999, 70, 19261932.
(30) Tam, S.; Fajardo, M. E. Single and Double Infrared Transitions
in Rapid-Vapor-Deposited Parahydrogen Solids: Application to
Sample Thickness Determination and Quantitative Infrared Absorp-
tion Spectroscopy. Appl. Spectrosc. 2001, 55, 16341644.
(31) Burkholder, J. B.; Howard, C. J.; McKellar, A. R. W. Fourier
Transform Infrared Spectrum of the v
2
Band of the NH
2
Radical. J.
Mol. Spectrosc. 1988, 127, 415424.
(32) McKellar, A. R. W.; Vervloet, M.; Burkholder, J. B.; Howard, C.
J. A Combined Analysis of the v
1
, v
3
, and 2v
2
Vibrational States of the
NH
2
Radical Using Fourier Transform Absorption and Emission Data.
J. Mol. Spectrosc. 1990, 142, 319335.
(33) Zhou, S.; Xie, D.; Lin, S. Y.; Guo, H. A New ab Initio Potential-
Energy Surface for NH
2
(X
2
A) and Quantum Studies of NH
2
Vibrational Spectrum and Rate Constant for the N(
2
D)+H
2
NH
+H Reaction. J. Chem. Phys. 2008, 128, 224316.
(34) Fajardo, M. E.; Tam, S. Rapid Vapor Deposition of Millimeters
Thick Optically Transparent Parahydrogen Solids for Matrix Isolation
Spectroscopy. J. Chem. Phys. 1998, 108, 42374241.
(35) Hinde, R. J.; Anderson, D. T.; Tam, S.; Fajardo, M. E. Probing
Quantum Solvation with Infrared Spectroscopy: Infrared Activity
Induced in Solid Parahydrogen by N
2
and Ar Dopants. Chem. Phys.
Lett. 2002, 356, 355360.
(36) Paulson, L. O.; Anderson, D. T.; Lundell, J.; Marushkevich, K.;
Melavuori, M.; Khriachtchev, L. Conformation Resolved Induced
Infrared Activity: trans- and cis- Formic Acid Isolated in Solid
Molecular Hydrogen. J. Phys. Chem. A 2011, 115, 1334613355.
(37) Raston, P. L.; Anderson, D. T. Infrared-Active Vibron Bands
Associated with Rare Gas Atom Dopants Isolated in Solid Para-
hydrogen. Low Temp. Phys. 2007, 33, 487492.
(38) Sears, V. F.; Vankranendonk, J. Theory of Q Branch of Infrared
Spectrum of Solid Hydrogen. Can. J. Phys. 1964, 42, 980&.
(39) Muller, H. S. P.; Klein, H.; Belov, S. P.; Winnewisser, G.;
Morino, I.; Yamada, K. M. T.; Saito, S. Terahertz Spectroscopy of the
Amidogen Radical, NH
2
. J. Mol. Spectrosc. 1999, 195, 177184.
(40) Ram, R. S.; Bernath, P. F. Revised Molecular Constants and
Term Values for the X
3

and A
3
States of NH. J. Mol. Spectrosc.
2010, 260, 115119.
(41) Ram, R. S.; Bernath, P. F.; Hinkle, K. H. Infrared Emission
Spectroscopy of NH: Comparison of a Cryogenic Echelle Spectro-
graph with a Fourier Transform Spectrometer. J. Chem. Phys. 1999,
110, 55575563.
(42) Chackerian, C.; Guelachvili, G.; Lopez Pineriro, A.; Tipping, R.
H. Rovibrational Intensities for the v=1 Bands of the X
3

NH
Radical: Experiment and Theory. J. Chem. Phys. 1989, 90, 641649.
(43) Robinson, A.; Brown, J.; Flores-Mijangos, J.; Zink, L.; Jackson,
M. Spectroscopic Study of the
14
NH Radical in Vibrationally Excited
Levels of the X
3

State by Far Infrared Laser Magnetic Resonance.


Mol. Phys. 2007, 105, 639662.
(44) Bernath, P. F.; Amano, T. Difference Frequency Laser
Spectroscopy of the v = 1 0 Transition of NH. J. Mol. Spectrosc.
1982, 95, 359364.
(45) Klaus, T.; Takano, S.; Winnewisser, G. Laboratory Measurement
of the N = 1 0 Rotational Transition of NH at 1 THz. Astron.
Astrophys. 1997, 322, L1L4.
(46) Kufeld, K.; Wonderly, W. R.; Paulson, L. O.; Kettwich, S. C.;
Anderson, D. T. Transient H
2
O Infrared Satellite Peaks Produced in
UV Irradiated Formic Acid Doped Solid Parahydrogen. J. Phys. Chem.
Lett. 2012, 3, 342347.
(47) Wonderly, W. R.; Anderson, D. T. Transient HDO Rovibra-
tional Satellite Peaks in Solid Parahydrogen: Evidence for Hydrogen
Atoms or Vacancies? Low Temp. Phys. 2012, 38, 853859.
(48) Hoshina, H.; Fushitani, M.; Momose, T.; Shida, T. Tunneling
Chemical Reactions in Solid Parahydrogen: Direct Measurement of
the Rate Constants of R+H
2
RH+H (R = CD
3
,CD
2
H,CDH
2
,CH
3
) at
5 K. J. Chem. Phys. 2004, 120, 37063715.
(49) Momose, T.; Hoshina, H.; Sogoshi, N.; Katsuki, H.;
Wakabayashi, T. Tunneling Chemical Reactions in Solid Para-
hydrogen: A Case of CD
3
+H
2
CD
3
H+H at 5 K. J. Chem. Phys.
1998, 108, 73347338.
(50) Hoshina, H.; Fushitani, M.; Momose, T. Infrared Spectroscopy
of Rovibrational Transitions of Methyl Radicals (CH
3
, CD
3
) in Solid
Parahydrogen. J. Mol. Spectrosc. 2011, 268, 164172.
(51) Kumada, T.; Kitagawa, N.; Noda, T.; Kumagai, J.; Aratono, Y.;
Miyazaki, T. An ENDOR Spectrum of H Atoms in Solid H
2
. Chem.
Phys. Lett. 1998, 288, 755759.
(52) Hegelund, F.; Burger, H.; Polanz, O. The High-Resolution
Infrared Spectrum of the
4
,
5
, and
6
Bands of Trans-Di-Imide
Revisted. J. Mol. Spectrosc. 1994, 167, 110.
The Journal of Physical Chemistry A Article
dx.doi.org/10.1021/jp408336n | J. Phys. Chem. A 2013, 117, 1383213842 13842

Вам также может понравиться