Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

pubs.acs.

org/JPCA Article

Potential Catalytic Role of Small Heterocycles in Interstellar H2


Formation: A Laboratory Astrochemistry Study on Furan and Its
Hydrogenated Forms
Published as part of The Journal of Physical Chemistry virtual special issue “10 Years of the ACS PHYS
Astrochemistry Subdivision”.
Anita Schneiker, Gopi Ragupathy, Gábor Bazsó, and György Tarczay*
Downloaded via BENEMERITA UNIV AUTONOMA DE PUEBLA on September 24, 2023 at 17:06:05 (UTC).

Cite This: J. Phys. Chem. A 2022, 126, 2832−2844 Read Online


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: It is now well-accepted in astrochemistry that the


formation of interstellar H2 is taking place on the surface of
interstellar grains. It has also been suggested a long time ago that
polyaromatic hydrocarbons (PAHs) can catalyze this process by
subsequent H atom addition and H abstraction reactions. Recent
quantum chemical computations suggested that small heterocycles
can be better catalysts than PAHs. In this study, the reaction of H
atoms with furan, 2,3- and 2,5-dihydrofurans, and tetrahydrofuran
were studied in solid para-H2 at 3.1 K. The reactions were followed
by Fourier transform infrared (FTIR) spectroscopy. By the analysis
of spectra, 2-hydrofuran-3-yl, 3-hydrofuran-2-yl, 2,3,4-trihydrofur-
an-5-yl, and 2,3,5-trihydrofuran-4-yl radicals were identified among
the products. The experiments revealed that all the possible H
atom addition and H abstraction cycles connecting furan and tetrahydrofuran proceed effectively in both directions at a low
temperature. This indicates the possible important role of small heterocycles in interstellar H2 formation. Furthermore, it also
indicates that, in the case of H atom excess, a quasi-equilibrium exists between the c-C4HxO (x = 4−8) species, and the ratios of
these species in an astrophysical object are determined by the rate of the different H atom addition and H abstraction reaction steps.

■ INTRODUCTION
Although H2 is the smallest and the most abundant molecule
for the interstellar H2 formation rate.1−3 These include the
capture of H atoms in sites of amorphous carbon or silicate
of the interstellar medium (ISM), its formation in the present grains, temperature fluctuations leading to explosive recombi-
universe is still not fully understood. Due to the low-pressure nations in a runaway event, and chemisorption. According to
conditions in the ISM, and due to the lack of a dipole moment the latter mechanism, H atoms are chemically bonded to a
for H2, its gas-phase formation rate from H atoms both by molecule on the surface of an icy grain (Mgr) (eq 3), after
three-body collision (eq 1) and by radiative association (eq 2) which another H atom can abstract a chemisorbed H atom,
mechanisms is practically zero.1−23
reforming the original molecule and a H2 molecule (eq 4).
H + H + X → H 2 + X* (1) Thus, this mechanism can be considered as a catalytic cycle.

H + H → H 2 + hν (2) Mgr + H → MHgr (3)


Therefore, it is now widely accepted that the observed
abundance and the formation rate of H2 in the ISM can be MHgr + H → Mgr + H 2 (4)
explained only by formation on the surface or in the ice layer of
interstellar grains. However, the models considering simple
physisorption do not provide a satisfactory solution for the Received: January 14, 2022
interstellar H2 formation problem. Namely, at very low Revised: March 25, 2022
temperatures the diffusion rate of H atoms in the ice is too Published: April 25, 2022
low, while at higher temperatures the residence time of H
atoms is too short for efficient H2 formation. Several complex
solid-phase reaction mechanisms were considered to account

© 2022 American Chemical Society https://doi.org/10.1021/acs.jpca.2c00306


2832 J. Phys. Chem. A 2022, 126, 2832−2844
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

Usually the first step (eq 3) has an activation energy,4−7 and temperature. The advantages and the limitations of this
at low temperatures this reaction can proceed via H atom method to study H atom reactions relevant to astrophysical
tunneling. Most often, the second step has no barrier. environments were recently discussed in more detail.37
Among the possible molecules that can act as catalysts in In the present paper, we apply the para-H2 matrix isolation
these reactions, polyaromatic hydrocarbons (PAHs)8−12 and method to investigate the low-temperature reactions of H
small organic molecules13,14 were investigated by theoretical atoms with furan (1), 2,3-dihydrofuran (2), 2,5-dihydrofuran
and experimental methods. The role of PAHs in H2 formation (3), and tetrahydrofuran (4). (See Figure 10 for the
can be especially important in photon-dominated regions structures.) On the basis of the experiences of the present
(PDRs).15−18 Computations have revealed that, in contrast to study and those of recent similar works, we argue that the para-
graphene or graphite, the H atom addition to benzene8,9 and H2 matrix isolation method is a useful tool to screen the
pyrene10 by a tunneling mechanism has non-negligible reaction mechanisms of low-temperature H atom reactions relevant to
rates at low temperatures. From the experimental side, it was interstellar chemistry.
shown that atomic and molecular hydrogen react with benzene
and small PAHs.19−22 Mennella et al. have exposed coronene
to D atoms and observed the formation of HD and D2
■ METHODS
Experimental Details. VIZSLA, the setup used in this
molecules.11 With this experiment they proved the possible study, has been described in detail recently.38 Briefly, the
catalytic role of neutral PAH molecules in the interstellar H2 experiments were performed in a high-vacuum-compatible
formation. Schneiker et al. have suggested that PAHs with an stainless steel chamber that can be evacuated to a base pressure
imperfect aromatic system, such as phenalene, which was of 1 × 10−8 mbar at room temperature. The deposition was
investigated experimentally in the same study, can have a done onto a gold-plated silver substrate mounted on the
smaller barrier in reaction 3 or 4 and, therefore, can be better coldfinger of an RDK-415D2 cryostat (Sumitomo Heavy
catalysts than PAHs with a perfect aromatic system.12 Barrales- Industries Inc.), which allows the substrate to be cooled down
́
Martinez and Gutiérrez-Oliva have shown by computations to 3.1 K. para-H2 was prepared by flowing normal-H2 through
that for some N- and Si-doped coronenes the H addition porous Fe(III) oxide (Sigma-Aldrich, hydrated, catalyst grade,
process has no barrier at all; consequently, the H atom 30−50 mesh) at 13.9 K. The produced para-H2 was collected
chemisorption rates of these species are much higher than that and stored in a 1 L glass bulb until it was used for sample
of benzene and coronene.23 In a recent study, Miksch et al.9 preparation.
have computed the rate of the H atom addition to benzene and The sample, furan (Aldrich, ≥99%), or 2,3-dihydrofuran
small aromatic heterocycles, including pyridine, pyrrole, furan, (Aldrich, 99%), or 2,5-dihydrofuran (Aldrich, 97%), or
thiophene, silabenzene, and phosphorene, in the temperature tetrahydrofuran (Aldrich, 99.5%, purified with distillation)
range of 50−500 K. They have found that at 50 K the carbon was premixed with para-H2, in a ratio of 1:2000−1:1000 in a
atom adjacent to the heteroatom of pyrrole or furan can form a gas-mixing vacuum line. (The relatively high concentrations
covalent bond with a H atom ca. 200 times faster than that of a were used because these species and the expected products
carbon atom of benzene. They concluded that these findings have only medium and weak IR absorption bands.) The
are important not only for proving the potential role of these premixed gas and Cl2 (Messer, 99.8%; the major contaminant
species in interstellar H2 formation but also because the is O2) were introduced into the vacuum chamber at roughly a
relatively fast H atom addition rate to furan may explain the 3 cm distance from the substrate via two stainless steel
absence of furan in the ISM.24−28 capillary arrays. The flow rates were controlled by two leak
The hydrogenation reaction of pyrrole was investigated by valves. The mixture deposition rate was set to keep the
Amicangelo and Lee in solid para-H2 at 3.2 K.29 They have pressure inside the chamber at 7 × 10−5 mbar, while the flow
observed the formation of 2,3-dihydropyrrol-2-yl and 2,3- rate of Cl2 was adjusted to get a Cl2/para-H2 ratio between
dihydropyrrol-3-yl radicals. The observed 4−5:1 formation 1:500 and 1:1000. The gases were codeposited onto the
yield of these radicals was found to be consistent with the substrate at 3.1 K.
computed barrier heights and with the computed relative IR absorption spectra were recorded with a Bruker Invenio
reaction rates. Fourier transform infrared (FTIR) spectrometer working in
The para-H2 matrix, applied in the above-mentioned study, reflection−absorption mode and using a liquid-nitrogen-cooled
has several advantages for the investigation of the low- midband HgCdTe detector. For the mid-IR (MIR) spectrum,
temperature reactions of H atoms. First, the interaction of the 32−128 scans were collected between 4000 and 600 cm−1 with
investigated molecules with the para-H2 host is negligible. a resolution of 0.5 cm−1, and for the near-IR (NIR) spectrum,
Therefore, these species have sharp spectral bands, which 32 scans were collected between 9000 and 600 cm−1 with a
facilitates spectral assignments. Second, H atoms can easily be resolution of 0.5 cm−1. The purity of the para-H2 was checked
generated in situ in a relatively high mixing ratio in this host. by collecting the NIR spectrum of the deposited matrix and
The method of generating H atoms in a para-H2 matrix (see measuring the absorption bands of ortho-H2 and para-H2.39
the details in the Experimental Details section) was originally Before deposition, 256 scans taken from the pure substrate
established by the groups of Anderson and Lee during which were averaged for the MIR and NIR background spectra using
numerous astrochemically relevant H atom addition, as well as the same settings as detailed above.
H atom abstraction, reactions have been investigated.30−36 All the UV and NIR irradiations were carried out by an
Although solid para-H2 is not a typical astrophysical environ- optical parametric oscillator (OPO, GWU/Spectra Physics
ment, as it is colder and less polar than typical astrophysical VersaScan MB 240, fwhm ≈ 5 cm−1) equipped with a
ices, it provides excellent conditions for sensitive monitoring of frequency-doubling unit (Spectra Physics uvScan). The OPO
these processes at first glance. It is expected that, if a reaction is was pumped by a pulsed Nd:YAG laser (Spectra Physics
taking place at around 3−4 K in solid para-H2, then it can also Quanta Ray Lab 150, P ≈ 2.1−2.2 W, λ = 355 nm, f = 10 Hz,
proceed in an astrophysical ice at a somewhat higher pulse duration = 2−3 ns).
2833 https://doi.org/10.1021/acs.jpca.2c00306
J. Phys. Chem. A 2022, 126, 2832−2844
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

Figure 1. H addition and H abstraction PESs as computed at the B3LYP/cc-pVTZ level of theory without (normal characters) and with (bold
characters) ZPVE correction. Values are in kJ mol−1. (a) H atom addition reactions of furan (1). (b) H atom abstraction reactions of
tetrahydrofuran (4). (c) H atom addition reactions of 2,3-dihydrofuran (2) and 2,5-dihydrofuran (3). (d) H atom abstraction reactions of 2 and 3.

H atoms were generated in the matrix by a slightly modified which wavelength the photon flux of the laser is much lower,
approach of Anderson’s and Lee’s groups.29−36 Accordingly, the matrix was irradiated for 60−125 min. The changes during
the sample was exposed to 365 nm light for 90−120 min (∼75 the irradiations were monitored by taking spectra of 32 scans
μJ cm−2 at 10 Hz at the substrate). This irradiation breaks the every minute. Experiments without Cl2 were also carried out,
bond of Cl2, and the two Cl atoms get separated and do not with photolysis at 400, 365, 330, 300, 270, 240, and 216 nm.
recombine in para-H2 due to the diminishing cage effect.34 To estimate the mixing ratios, the recipe given in ref 39 was
followed.
Cl 2 → 2Cl (5)
Quantum Chemical Computations. The interpretation
To monitor this process, and estimate the Cl mixing ratio of experimental results was aided by quantum chemical
(from the 5095 cm−1 absorption band of Cl atoms)40 in the computations, all done by the Gaussian 09 (rev. D01) program
matrix, NIR spectra were taken every 30 min during this package.41 Geometry optimizations of stationary points
irradiation. Between the measurements, the IR light of the (minima and transition structures) on the potential energy
spectrometer was blocked to minimize IR irradiation of the surface were performed at the B3LYP42,43/cc-pVTZ44 level of
sample. Before each experiment, blank experiments were theory. To ensure that the transition structures connect the
carried out on samples deposited without Cl2, in order to right minima, intrinsic reaction coordinate (IRC) computa-
check that 365 nm irradiation does not photolyze the tions45 were performed. Anharmonic vibrational frequencies
heterocycle in the matrix. and intensities were computed by vibrational perturbation
Only a small fraction of Cl atoms react with solid para-H2. theory (VPT2)46,47 at the B3LYP/cc-pVTZ level.
However, this reaction can be initiated by vibrationally exciting
the para-H2 molecules. In the present work, 90 min of 2217
nm (∼700 μJ cm−2 at 10 Hz at the substrate) irradiation was
■ RESULTS AND DISCUSSION
In order to study both the H atom addition and the H atom
applied for this purpose. The excited para-H2 molecules readily abstraction reactions, a comparable set of experiments was
react with the Cl atoms: performed for 1 and for its hydrogenated closed-shell
Cl + H 2* → HCl + H (6)
analogues, 2, 3, and 4. The computed structure, anharmonic
vibrational wavenumbers, and intensities of all of the discussed
It should be also noted that the radiation from the source of species are summarized in Table S1.
the IR spectrometer can also induce this reaction. The H atom As discussed in the Methods section, H atoms were
generation was followed by obtaining spectra of 32 scans every generated after deposition of the Cl2-doped matrix by the
minute and detecting the HCl absorption in the MIR region. photochemical reactions. In the first step, Cl2 is dissociated to
The materials were held in darkness overnight after completing Cl atoms by 365 nm irradiation (eq 5). In the next step 2217
the NIR laser irradiation, and 128-scan MIR spectra were nm para-H2 molecules are vibrationally excited, which react
obtained every 30 min. with Cl atoms, yielding H atoms and HCl (eq 6). Then, the
After the dark processes, secondary photolyses at 400, 365, matrix is kept in the dark, and H atoms can diffuse in the
330, 300, 270, 240, and 216 nm were carried out. With the matrix and can react with the investigated molecules. These
exception of the 216 nm irradiation, at each wavelength, the latter two are the astrochemically relevant steps, and we
matrix was irradiated for 30 min. In the case of 216 nm, at observed the same reactions during these steps. The difference
2834 https://doi.org/10.1021/acs.jpca.2c00306
J. Phys. Chem. A 2022, 126, 2832−2844
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

is that, unlike in the dark process, some suprathermal H atoms be formed in much less amount. The second H atom addition
might be present in the matrix during the 2217 nm irradiaton. can yield 2, with a minor product of 3.
While H atoms can move in the matrix by quantum diffusion The low-frequency region of the spectra recorded at
in the dark period,31,32 the diffusion of HCl (and that of the different phases of the experiment is displayed in Figure 2;
remaining Cl atoms) is hindered at 3.1 K. Therefore, HCl
molecules (and Cl atoms) remain isolated from the
investigated molecules. In the final step, a secondary UV
photolysis was performed to help the assignments of the
spectral bands of the products of H atom reactions. In the
discussion, to confirm the spectral assignments, we show the
deposited spectra and the four difference spectra correspond-
ing to the above-mentioned steps: 365 nm photolysis, 2217 nm
photolysis, the dark process, and secondary photolysis. The
spectra corresponding to blank experiments (i.e., matrixes
deposited without Cl2) are provided in the Supporting
Information.
Reaction of H Atoms with Furan (1). The H addition
potential energy surface (PES) of 1 computed at the B3LYP/
cc-pVTZ level of theory is displayed in Figure 1a. The first H
atom addition reactions have small barriers; the zero-point
vibrational energy (ZPVE)-corrected values are 5 and 16 kJ
mol−1 for the C atom in the ortho position (or position 2) and
the C atom in the meta position (or position 3), respectively.
These reactions result in the formation of 2-hydrofuran-3-yl
(R1) and 3-hydrofuran-2-yl (R2) radicals, respectively.
Although, at the MPWB1K/def2-TZVP level of theory, Miksch
et al. have computed larger barriers, 15 and 27 kJ mol−1,9
respectively, their results agree qualitatively well with ours,
regarding the energy order of the transition states and those of
the products. According to former studies, barriers with a
similar or even larger height are permeable to H atoms even at
3−4 K by tunneling, and the reaction can be observed on the
laboratory time scale.13 Miksch et al. have computed the Figure 2. Low-frequency region of the IR spectra corresponding to
reaction rate to be 10−17 and 10−20 cm3 s−1 at 50 K for the experiments with furan (1). (a) The IR spectrum of 1 codeposited
paths yielding R1 and R2, respectively.9 with Cl2 in a para-H2 matrix at 3.1 K. (b) The difference IR spectrum
As concluded by Miksch et al., the H atom addition to the O obtained by subtraction of the spectrum recorded after deposition
atom is highly endothermic, with a reaction energy of +159 kJ from the spectrum measured after 120 min of 365 nm irradiation. (c)
mol−1 at the MPWB1K/def2-TZVP level of theory.9 There- The difference IR spectrum obtained by subtraction of the spectrum
fore, it is unfeasible under low-temperature conditions. recorded after the 365 nm irradiation from the spectrum recorded
Because of the same reason, the H atom abstractions can after a subsequent 30 min of 2217 nm irradiation. (d) The difference
also be disregarded during the analysis of the experiments, as at IR spectrum obtained by subtraction of the spectrum recorded after
the 2217 nm irradiation from the spectrum recorded after keeping the
the ZPVE-corrected B3LYP/cc-pVTZ level of theory we
sample in the dark for 13 h. (e) The difference IR spectrum obtained
computed a reaction energy of +55 kJ mol−1 for the abstraction by subtraction of the spectrum recorded before from the one recorded
of the H atom at both the ortho and at the meta positions. after 30 min of secondary 270 nm photolysis. Δ, water contamination;
As also noted by Miksch et al.,9 the second H atom addition ○ , bands of HO 2 radical; ×, band of ClO 2 ; *, bands of
reaction is barrierless, as it is a radical−radical recombination uncompensated atmospheric CO2. For clarity, only the most intense
process. In this reaction, the R1 radical can form a covalent bands of 1 and those of the products are marked, and the bands
bond with a H atom with its carbon atoms in the 3, 4, and 5 tentatively assigned to HCl complexes are unmarked. See the enlarged
ring positions. Among these paths, the reaction with the C bands of 2 and 3 observed after the H atom reaction in a different
atom at position 4 would yield a high-energy biradical, which is measurement with a higher H atom mixing ratio in Figure S6.
unlikely to be formed, and therefore, it has been omitted from
the present study. The other two reaction paths of the R1 + H see the full-scale spectra in Figures S1 and S2. Figure 2a shows
reaction yield 2 and 3, respectively. 2 is a thermodynamically the spectrum recorded after deposition. All the intense bands
more favored product than 3; the energy difference between can be assigned to 1. (See the assignments in Table 1 and the
the two isomers is 13 kJ mol−1 at the ZPVE-corrected B3LYP/ complete list of the computed frequencies and intensities of
cc-pVTZ level of theory. The R2 + H reaction results in the the fundamental modes in Table S1.) Some very weak bands
formation of 2 if the H atom is reacting with the C atom at belong to water contamination at 3765.4 and 1631.8 cm−1.52
position 2. In this case, the reaction of a H atom with C atoms As can be seen from the difference spectrum in Figure 2b,
at positions 4 or 5 would yield higher-energy biradicals, which upon 365 nm irradiation the bands of 1 decreased and some
are not considered in this study. weak intensity new bands appeared in the spectrum. Note that
From the above computational results, in line with the some HCl formation was observed even during this irradiation
conclusions of Miksch et al.,9 it can be predicted that the main (see Figure S1), which indicated that some H atoms were
primary product of the 1 + H reaction is R1; R2 is expected to already present in the matrix during this phase. (This can likely
2835 https://doi.org/10.1021/acs.jpca.2c00306
J. Phys. Chem. A 2022, 126, 2832−2844
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

Table 1. Computed Anharmonic and Experimental Table 2. Computed Anharmonic and Experimental
Vibrational Wavenumbers (in cm−1) and IR Intensities (in Vibrational Wavenumbers (in cm−1) and IR Intensities (in
km mol−1) of Furan (1) km mol−1) of 2-Hydrofuran-3-yl Radical (R1)
computed (C2v) exptl computed (C1) exptl
a b c
B3LYP/cc-pVTZ gas phase Ar matrix para-H2 matrix assignment B3LYP/cc-pVTZ para-H2 matrix assignment
613.6 (24) 603 603.0 (m) ν14 622.3 (21) 621.0 (m) ν21
749.1 (99) 745 744.1 746.2 (vs) ν13 753.3 (6) 745.7 (m) ν20
876.5 (17) 871 869.1 869.5 (ms) ν8 839.5 (5) 840.4 (w) ν19
882.0 (1) 877.1 (w) ν21 870.5 (7) 862.7 (w)a ν18
997.5 (41) 995 993.6 994.4 (vs) ν7 902.9 (12) 913.8 (m) ν17
1065.1 (11) 1066 1065.0 1065.6 (s) ν6 928.9 (61) 935.0 (s) ν16
1163.7 (17) 1180 1177.7 1178.2 (s) ν19 1017.7 (7) 1012.1 (m) ν14
1268.5 (<1) 1267 1265.7 (w) ν18 1131.7 (36) 1143.5 (m) ν12
1381.9 (4) 1384 1383.3 (w) ν4 1235.8 (22) 1239.9 (w) 2ν21
1476.1 (13) 1491 1486.5 (m) ν3 1319.3 (12) 1329.4 (m) ν9
1557.1 (<1) 1556 1554.6 (w)d ν17 2804.7 (47) 2812.3 (w) ν4
3121.1 (2) 3126.3 (m) ν16 a
Uncertain due to low intensity.
3130.2 (1) 3129 3127.2 (w) ν2
3141.2 (3) 3140 3136.7 (vw)d ν15 Table 3. Computed Anharmonic and Experimental
3149.9 (<1) 3161 3156.7 (w) ν1 Vibrational Wavenumbers (in cm−1) and IR Intensities (in
a
Ref 48. bRefs 49−51. cThis work. dUncertain due to low intensity. km mol−1) of 3-Hydrofuran-2-yl Radical (R2)
computed (C1) exptl
be explained by the IR irradiation by the source of the IR
spectrometer.) In the blank experiment, when the sample was B3LYP/cc-pVTZ para-H2 matrix assignment
deposited without Cl2, no new bands appeared upon 365 nm 708.4 (40) 713.0 (m) ν21
irradiation (see Figure S3), which proves that 1 is not 831.0 (12) 830.9 (w) ν19
photolyzed at 365 nm. Some very weak bands that are 881.5 (3) 881.9 (w)a ν18
produced by the reaction of H atoms can be identified easily 924.5 (16) 933.4 (w) ν16
because they increased much more prominently in the next 1006.0 (37) 1012.1 (m) ν14
phase of the experiments, i.e., during and after the H atom 1078.6 (4) 1072.1 (w)a ν13
generation; see below. The other new bands should belong to 1152.9 (39) 1165.2 (s) ν11
either chlorinated products or to an HCl complex of 1. Among 1261.2 (12) 1273.6 (vw) ν9
these, on the basis of literature value (Ne matrix: 1442.5 1611.1 (19) 1622.9 (w)a ν6
cm−1),53 the most intense one at 1442.2 cm−1 can be assigned 2754.9 (52) 2767.3, 2769.6 (w)b ν4
a
to ClO2, a product of Cl atoms with O2 contaminants. The Uncertain due to low intensity. bUncertain due to nearby bands of
next two most intense bands observed at 1133.5 and 992 cm−1 HCl oligomers.
might be assigned to a 1·HCl complex (where HCl forms a
hydrogen bond with the oxygen atom of 1), as two intense atoms with the O2 contaminant from the Cl2 cylinder),54 water
bands of this complex were computed to be at 1157.2 and contamination (at 3765.4 and 1631.8 cm−1), and uncompen-
994.0 cm−1 (see Table S1). The formation of the 1·HCl sated atmospheric CO2 (at 720.5, 617.8, and 667.9 cm−1)52
complex can be rationalized by the fact that the Cl2 mixing and might be tentatively assigned to HCl complexes.
ratio in the sample was relatively high, and upon absorbing the Furthermore, in experiments with high H mixing ratios, we
photons by the Cl2 molecules, the matrix warms up, which could observe some of the higher-intensity, well-resolved bands
increases the diffusion rate. It can also be seen that the positive of 2 (at 1140.6 and 1625.3 cm−1) and 3 (at 897.8 and 1090.8
bands in the spectrum of Figure 2b are much weaker than cm−1). These bands are enlarged in Figure S6, and the
those of the negative ones. This might be explained by the assignments of 2 and 3 are given in the section after the next
same warming effect, which resulted in the partial sublimation one.
of the matrix. The remaining bands are weak and cannot be After the H atom generation by 2217 nm irradiation, the
safely assigned. It should be noted that the intensity of these matrix was kept in the dark. According to former studies, H
unassigned weak bands hardly changed during the 2217 nm atoms are still present in this period in the matrix, and if there
irradiation and in the dark process. are no other reaction partners, they are typically completely
Upon the generation of H atoms by 2217 nm irradiation, consumed in the H + H recombination process in ca. 10−12 h.
some further bands appeared in the spectrum (positive bands As can be seen in Figures 2d, S4, and S5, the bands that
in the difference spectrum in Figure 2c), while the bands of 1 appeared upon 2217 nm photolysis and were assigned to R1
decreased. Upon the different temporal behaviors, these bands and R2 or to their HCl complexes are further increased in this
could be assigned to different groups. On the comparison of period, proving that these are products of a H atom reaction
computed spectra, the group with some intense bands could be and that the reaction proceeds at 3.1 K. (Note that upon 2217
assigned to R1 (see Table 2 and Figure S4), while another nm photolysis, in reaction 6, the heat of reaction releases in the
group of much less intense bands was assigned to R2 (see form of the kinetic energy of the H and HCl products.
Table 3 and Figure S5). The remaining medium and most of Therefore, before dissipation of the kinetic energy into the
the small-intensity bands belong to HO2 radical (at 1409.7, matrix, some suprathermal H atoms might be present, which
1392.4, and 1100.1 cm−1; the product of the reaction of H can react with the sample more readily. The extra energy is
2836 https://doi.org/10.1021/acs.jpca.2c00306
J. Phys. Chem. A 2022, 126, 2832−2844
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

rapidly and completely dissipated into the matrix when the abstraction PES of 4. The abstraction of a H atom from the
2217 nm irradiation is switched off.) ortho position, resulting in 2,3,4-trihydrofuran-5-yl (R3)
Finally, secondary UV irradiations were consecutively radical, has a barrier of only 5 kJ mol−1 without ZPVE
carried out to help the assignments at 400, 365, 330, 300, correction at the B3LYP/cc-pVTZ level of theory, and the
270, 240, and 216 nm. In the cases when Cl2 was completely barrier completely disappears when the ZPVE correction is
bleached in the primary 365 nm photolysis, no significant accounted for. The H atom abstraction from the meta position
change was observed in the spectrum upon 400, 365, 330, and of the ring, yielding 2,3,5-trihydrofuran-4-yl (R4) radical, has a
300 nm photolysis. The decrease of the bands assigned to R1 somewhat higher, but still relatively low, computed barrier; it is
and R2 was observed with a higher rate when the irradiation only 20 and 13 kJ mol−1 without and with ZPVE correction,
wavelength was decreased to 270 nm, which is displayed in respectively. Considering that similar, slightly higher barriers
Figure 2e. At the same time, the bands of 1 increased. This were computed for the reaction of H atom addition to furan, it
observation can be explained by the back-formation of 1 (and a is expected that the H atom abstraction reaction from 4 can
H atom) from R1 and R2 radicals and also by the also proceed at 3.1 K.
decomposition of the 1·HCl complex. The different photo- The second H abstraction from the C atom in position 4 of
chemical decomposition rates of R1, R2, and the complexes at R3 results in the formation of 2. The H atom abstraction from
270 and 240 nm further supported their assignments. (The the C atom in position 3 of R4 leads to the formation of 3,
spectrum of the secondary 216 nm irradiation was not used for while that from the C atom in position 5 leads to the formation
the analysis. At this wavelength the decomposition of furan was of 2. We have found all of these three H atom abstraction
observed.) reactions to be barrierless, which is consistent with former
The changes of the mixing ratios of 1, R1, R2, 2, and 3 upon computational studies for other small organic radicals when the
the 2217 nm photolysis and the dark period are shown in H atom abstraction yields a low-energy closed-shell mole-
Figures 3 and S7. In good agreement with the computational cule.13 Any other H atom abstractions from R3 or R4 would
yield high-energy biradicals, which are not considered in this
study.
The low-frequency region of the IR spectrum of a 4/Cl2
mixture deposited in para-H2 is shown in Figure 4a, and the
full-scale spectra can be seen in Figure S8. Each band was
observed in the spectrum of 4 deposited in a para-H2 matrix
without Cl2 (see Figure S9). The high- and medium-intensity
bands are assigned to the fundamental modes of the lower-
energy (Cs) ring puckering conformer of 4 (see Table 4). A
weak band of the water contaminant can be seen in the
displayed spectrum region at 1631.8 cm−1.52 The remaining
unassigned weak bands most likely belong to combination and
overtone modes of the Cs conformer of 4 or might be the most
intense bands of the C2 puckering form of 4.
In the case of the 365 nm irradiation of the matrix
containing only 4, just minor changes were observed in band
shapes. Upon the photolysis of the sample also containing Cl2,
the bands of 4 decreased (Figure 4b). This can be explained by
the formation of some H atoms, formed in the reaction of H2
molecules with Cl, which is induced by the IR light of the
Figure 3. Temporal profiles of the estimated mixing ratios of 1, R1, source of the spectrometer. Therefore, the H atoms have
R2, 2, and 3 after 365 nm irradiation of 1 and Cl2 codeposited in
consumed 4 already in this phase of the experiment. Some of
para-H2 at 3.1 K. The vertical line indicates the end of the NIR laser
irradiation. The estimated mixing ratio of total H atoms generated was the newly appearing bands (i.e., the positive bands in the
820 ± 140 ppm. difference spectrum), which increased also in the next phases
of the experiments, should be assigned to the products of the
H atom reaction; see below. Similar to the experiment with 1,
predictions, the major product is R1. The mixing ratio of this some remaining bands should belong to complexes of HCl
species is clearly increasing even in the dark period. R2 is with 4, and we cannot exclude that some byproducts are
formed in a considerably smaller amount, and mostly during formed in the reaction of Cl atoms with 4. Furthermore, the
the 2217 nm photolysis. 2 and 3 are formed in an even smaller significantly more intense negative bands compared to the
amount. Since the second H atom addition reaction has no positive ones in the difference spectrum of Figure 4b reveal
barrier, this observation can be explained most likely by the that some of the matrix material was sublimed by the laser.
relatively small amount of R1 and R2 formed and partially by This is consistent with the observation that the background
the H abstraction process from 2 and 3. (See the next pressure in the vacuum chamber slightly increases during
sections.) In conclusion, these experiments prove that the H irradiations.
addition reaction to 1 proceeds with a non-negligible rate at Upon H atom generation by 2217 nm irradiation (Figure
3.1 K. 4c) and keeping the matrix subsequently in the dark (Figure
Reaction of H Atoms with Tetrahydrofuran (4). 4d), some of the bands that appeared during the 365 nm
Following the increasing complexity order, we continue the photolysis further increased. These bands can be grouped to
discussion with the analysis of the computational and different species based on their different intensity ratios in the
experimental results on 4. Figure 1b shows the H atom former three spectra (i.e., Figure 4b−d). Some bands can be
2837 https://doi.org/10.1021/acs.jpca.2c00306
J. Phys. Chem. A 2022, 126, 2832−2844
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

Table 4. Computed Anharmonic and Experimental


Vibrational Wavenumbers (in cm−1) and IR Intensities (in
km mol−1) of Tetrahydrofuran (4)
computed (Cs) exptl
a
B3LYP/cc-pVTZ gas phase para-H2 matrixb assignment
780.2 (4) 835.6 834.8 (w) ν15
843.7 (5) 868.8 869.6 (w) ν31
883.1 (6) 898.9 901.4 (s) ν14
910.5 (2) 903.2 903.1 (m) ν30
919.5 (28) 921.7 922.8 (m) ν13
932.1 (<1) 952.5 952.0 (w) ν12
1018.9 (5) 1028.0 1024.6 (w) ν11
1058.6 (85) 1072.0 1077.7 (s) ν29
1125.6 (2) 1130.2 1130.7 (w) ν28
1189.7 (4) 1174.3 1171.4 (mw) ν10
1203.1 (5) 1195.8 1195.4 (w) ν27
1224.8 (2) 1211.8 1207.7 (w) ν26
1232.3 (2) 1238.0 1239.8 (m) ν9
1267.9 (1) 1292.0 1292.6 (w) ν8
1292.2 (1) 1315.0 1314.0 (w) ν25
1325.1 (<1) 1339.0 1339.5 (w) ν24
1363.1 (4) 1368.0 1365.2 (w) ν7
1451.6 (1) 1441.0 1438.9 (w) ν23
1464.8 (3) 1466.0 1467.7 (w) ν6
1469.4 (<1)/1469.4 (<1) 1487.0 1485.9 (w) ν22/ν5
2810.1 (37) 2851.7 2850.0 (m, sh) ν21
2813.4 (73) 2854.3 2853.9 (s) ν4
Figure 4. Low-frequency region of IR spectra corresponding to 2870.5 (19) 2862.2 2868.8 (s) ν20
experiments with tetrahydrofuran (4). (a) The IR spectrum of 4 2937.6 (4) 2934c 2929.9 (m) ν19
codeposited with Cl2 in a para-H2 matrix at 3.1 K. (b) The difference 2964.6 (40) 2941.0c 2941.2 (m) ν2
IR spectrum obtained by subtraction of the spectrum recorded after
2959.2 (71) 2962.0c 2958.9 (m) ν1
deposition from the spectrum measured after 120 min of 365 nm
irradiation. (c) The difference IR spectrum obtained by subtraction of 2953.4 (84) 2981.3 2980.3 (s) ν18
a
the spectrum recorded after the 365 nm irradiation from the spectrum Refs 55 and 56. bThis work. cLiquid phase, ref 56.
recorded after a subsequent 90 min of 2217 nm irradiation. (d) The
difference IR spectrum obtained by subtraction of the spectrum Table 5. Computed Anharmonic and Experimental
recorded after the 2217 nm irradiation from the spectrum recorded Vibrational Wavenumbers (in cm−1) and IR Intensities (in
after keeping the sample in the dark for 11.3 h. (e) The difference IR km mol−1) of 2,3,4-Trihydrofuran-5-yl (R3)
spectrum obtained by subtraction of the spectrum recorded before
from the one recorded after 30 min of secondary 270 nm photolysis. computed (C1) exptl
Δ, water contamination; ○, bands of HO2 radical; ×, band of ClO2. B3LYP/cc-pVTZ para-H2 matrix assignment
For clarity, only the most intense bands of 4 and those of the products
are marked, and the bands tentatively assigned to HCl complexes are 1148.0 (69) 1156.0 (m) ν18
unmarked. 1175.9 (23) 1185.1 (m) ν16
1374.6 (9) 1374.7 (w)a ν11
a
assigned to HO2 radical (at 1409.7, 1392.4, and 1100.1 Uncertain due to low intensity.
cm −1 ) 54 generated from O 2 contamination, to water
contamination (at 3765.4 and 1631.8 cm−1),52 and to Table 6. Computed Anharmonic and Experimental
uncompensated atmospheric CO2 (at 720.5, 617.8, and 667.9 Vibrational Wavenumbers (in cm−1) and IR Intensities (in
cm−1). Unfortunately, as the computations reveal, the expected km mol−1) of 2,3,5-Trihydrofuran-4-yl (R4)
products of the reaction of 4 with H atoms, R3 and R4, have
computed (C1) exptl
only a few bands with medium IR absorption, which makes the
assignment less straightforward. However, with the help of the B3LYP/cc-pVTZ para-H2 matrix assignment
above-mentioned grouping, by the comparison of these with 894.9 (36) 908.5 (m) ν24
computed spectra and by the aid of further experiments (270 945.4 (2) 943.0 (vw)a ν21
nm secondary photolysis and reaction of H atoms with 2 and 1055.9 (78) 1072.1 (w) ν18
3), some bands were assigned to R3 and R4 with high 1155.3 (3) 1144.1 (w) ν17
confidence (see Tables 5 and 6 and Figures S10 and S11). The 2825.7 (85) 2835.8 (m) ν5
most intense bands of one of the possible second H atom a
Uncertain due to low intensity.
abstraction products, 2, are also clearly identifiable among the
positive bands of the difference spectra of Figure 4, parts c and
d (enlarged in Figure S12), while the bands of 3 cannot be
clearly identified. (See the assignments of these species in the complexes of 4, R3, and R4. It is possible that some unassigned
next section.) Some remaining bands might belong to HCl weak bands belong to some chlorinated byproducts, but the
2838 https://doi.org/10.1021/acs.jpca.2c00306
J. Phys. Chem. A 2022, 126, 2832−2844
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

complexity of the spectra and the signal-to-noise level do not Reaction of H Atoms with 2,3-Dihydrofuran (2) and
allow making further assignments. 2,5-Dihydrofuran (3). Although 2 and its isomer, 3, are
Upon secondary 270 nm UV photolysis, the bands assigned topologically different, both of them are connecting 1 and 4 on
to R3 and R4 are slowly bleached (see Figure 4e). Most likely, the H atom abstraction/addition PES; therefore, we discuss
this photolysis results in a H atom loss from R3 and R4, which the computational results and the experiments carried out for
leads to the formation of 2 and 3. This assumption is these two complexes together.
consistent with the observation that, indeed, the bands of 2 The computed PESs of the H atom addition reactions of 2
and 3 are increasing during this irradiation. Somewhat and 3 are shown in Figure 1c. All of the possible H additions to
surprisingly, in the difference spectrum corresponding to this 2 and 3 have very low energy barriers. In detail, the H atom
irradiation, the most intense negative bands are not the ones addition to the C atom in position 5 of 2, leading to R4, has a
that were assigned to R3 and R4. These negative bands cannot barrier of 5 kJ mol−1 at the ZPVE-corrected B3LYP/cc-pVTZ
belong to the higher-energy puckering form of 4 because they level of theory, while the H atom addition to the C atom in
appeared as small bands upon the 365 nm and the 2217 nm position 4 of 2, resulting in the formation of R3, has no barrier
photolysis and they are not bleached when the sample at all. The sp2 C atoms of 3 are in symmetrical positions; the H
deposited without Cl2 was irradiated at 270 nm. These atom addition to either of these C atoms results in the
bands are the ones that appeared mostly upon the initial 365 formation of R4. This reaction has a computed barrier of 7 kJ
nm photolysis and can be tentatively assigned to either HCl mol−1. R3 cannot be formed directly from 3. The H atom
complexes of 4 or to chlorinated products. This interpretation addition reactions leading to 4 from both R3 and R4 are
is consistent with the observation upon the 270 nm secondary barrierless radical−radical recombination processes.
photolysis of the experiment carried out with 1, and the The H atom abstraction PESs of 2 and 3 are displayed in
photodecomposition of 4·HCl explains the formation of 4, Figure 1d. Similar to the H atom addition, the H atom
which appears with positive bands on the difference spectrum abstraction can lead to two different radicals, R1 and R2, in the
in Figure 4e. Among the photolysis products, 1 can also be case of 2, and only one radical product, R1, for 3. The latter
identified in a small amount. This might be explained by reaction has a barrier of only 2 kJ mol−1 without ZPVE
assuming that possibly a trace amount of R1 and/or R2, that correction at the B3LYP/cc-pVTZ level of theory and has no
could not be identified, is formed via 2 (and 3) in the reaction barrier if the PES is corrected by ZPVE. Almost the same
values were obtained for the barrier from 2 to the same radical,
of 4 with H atoms, since R1 and R2 decompose to 1 upon 270
R1, while the other path from 2 to R2 has higher, but still low,
nm irradiation.
computed barriers of 9 and 3 kJ mol−1 with and without ZPVE
Summing up the results of the experiment with 4, both
correction, respectively. The H abstraction from an sp2 C leads
products of the expected first H atom abstraction reactions, R3
to high-energy products; therefore, these are not considered in
and R4, and one of the possible products of the second H
the analysis. According to our PES scans, the second H atom
atom abstraction reactions, 2, are identified. The temporal
abstractions from the sp3 C atoms of R1 and R2, leading to 1,
profiles of the mixing ratios of these products are plotted in have no barriers.
Figure 5. As can be seen, the mixing ratios of all these species Summing up the computational predictions and the
are monotonically increasing even in the dark period, which topological considerations, the H atom reactions with 2 can
clearly proves that the H atom abstraction reactions from 4, lead to primary products of R1, R2, R3, and R4, while in the
R3, and R4 are feasible and relatively efficient even at 3.1 K. case of 3, only the formation of R1 and R4 is expected. In the
case of large H atom excess, the formation of 1 and 4 might be
feasible in both cases, and consecutive addition−abstraction or
abstraction−addition reactions can cause 2 ↔ 3 isomerization.
However, it should be noted that these simple computations
cannot predict the relative rate of these reactions; therefore,
the quasi-equilibrium ratios of the products are also
unpredictable from these computations.
The low-frequency part of the spectra belonging to the same
set of experiments, which were discussed in the previous two
sections, are plotted in Figures 6 and 7 for the experiments
with 2 and 3, respectively. (See the full-scale spectra for two
parallel experiments for 2 in Figures S13 and S14 and those for
3 in Figures S15 and S16 and the spectra corresponding to the
blank experiment without Cl2 in Figures S17 and S18.) The
spectra of 2 and 3 deposited in a para-H2 matrix are shown in
panels a of Figures 6 and 7. Except for very small traces of
water, and very small traces of 2 in the deposited sample of 3,
and 3 in the deposited sample of 2, only bands of the pure
samples were identified. The assignments of 2 and 3 are given
in Tables 7 and 8. As mentioned, these assignments were used
Figure 5. Temporal profiles of the estimated mixing ratios of 4, R3,
R4, and 2 after 365 nm irradiation of tetrahydrofuran (4) and Cl2 to identify 2 or 3 among products of the reaction of H atoms
codeposited in para-H2 at 3.1 K. The vertical line indicates the end of with 1 and 4.
the NIR laser irradiation. The estimated mixing ratio of total H atoms At the first look, the difference spectra corresponding to the
generated was 760 ± 160 ppm. Note that some R3 was produced irradiation experiments and to the changes during the dark
before the NIR laser irradiation during the 365 nm photolysis. period are rather congested for 2 and simpler for 3. This
2839 https://doi.org/10.1021/acs.jpca.2c00306
J. Phys. Chem. A 2022, 126, 2832−2844
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

Figure 6. Low-frequency region of IR spectra corresponding to Figure 7. Low-frequency region of IR spectra corresponding to
experiments with 2,3-dihydrofuran (2). (a) The IR spectrum of 2 experiments with 2,5-dihydrofuran (3). (a) The IR spectrum of 3
codeposited with Cl2 in a para-H2 matrix at 3.1 K. (b) The difference codeposited with Cl2 in a para-H2 matrix at 3.1 K. (b) The difference
IR spectrum obtained by subtraction of the spectrum recorded after IR spectrum obtained by subtraction of the spectrum recorded after
deposition from the spectrum measured after 90 min of 365 nm deposition from the spectrum measured after 120 min of 365 nm
irradiation. (c) The difference IR spectrum obtained by subtraction of irradiation. (c) The difference IR spectrum obtained by subtraction of
the spectrum recorded after the 365 nm irradiation from the spectrum the spectrum recorded after the 365 nm irradiation from the spectrum
recorded after a subsequent 90 min of 2217 nm irradiation. (d) The recorded after a subsequent 70 min of 2217 nm irradiation. (d) The
difference IR spectrum obtained by subtraction of the spectrum difference IR spectrum obtained by subtraction of the spectrum
recorded after the 2217 nm irradiation from the spectrum recorded recorded after the 2217 nm irradiation from the spectrum recorded
after keeping the sample in the dark for 11.8 h. (e) The difference IR after keeping the sample in the dark for 11.5 h. (e) The difference IR
spectrum obtained by subtraction of the spectrum recorded before spectrum obtained by subtraction of the spectrum recorded before
from the one recorded after 30 min of secondary 270 nm photolysis. from the one recorded after 30 min of secondary 270 nm photolysis.
Δ, water contamination; ○, bands of HO2 radical; ×, band of ClO2. Δ, water contamination; ○, bands of HO2 radical; ×, band of ClO2; *,
For clarity, only the most intense bands of 2 and those of the products bands of uncompensated atmospheric CO2. For clarity, only the most
are marked, and the bands tentatively assigned to HCl complexes are intense bands of 3 and those of the products are marked, and the
unmarked. bands tentatively assigned to HCl complexes are unmarked.

immediately suggests that in the case of 2 more products were During the H atom generation process (Figures 6c and 7c)
formed than for 3. These spectra, even the congested ones, can and the dark period (Figures 6d and 7d) all the possible four
be relatively straightforwardly interpreted by checking the primary product radicals are generated from 2: R1 (identified
bands of the possible reaction products already assigned in the at 1329.4, 1143.5, 1012.1, 935.0 cm−1 and weak bands at
former sections. As in the cases of both 1 and 4, we have 1239.9, 913.8, 840.4, and 621.0, cm−1; see Figure S19), R2 (at
already observed changes during the 365 nm irradiation. 1165.2 and 933.4 cm−1, with weak bands at 1273.6, 1012.1,
Recalling the interpretation given in the previous section, it is and 830.9 cm−1; see Figure S20), R3 (at 1374.7, 1185.1, and
possibly due to the excitation by the source of the IR 1156.0 cm−1; see Figure S21), and R4 (at 1144.1 and 908.5
spectrometer, which can initiate the reaction of Cl atoms with cm−1; see Figure S22). On the basis of the spectra we cannot
H2. Among the products (positive bands), besides the radicals, confirm the formation of 1 and 4. A small amount of 3 might
some bands should be tentatively assigned either to the be formed from 2, but its most intense band is hardly larger
complexes of 2 or 3 with HCl or to chlorinated byproducts. than the noise level (Figure S23).
Once again, the complex formation can be active during this In the case of the same experiments with 3, we could identify
irradiation because relatively high energy is pumped into the among the radical products R1 (at 1143.4, 935.0 and less
matrix via the photon absorption by Cl2 molecules, increasing intense bands at 1329.4, 1239.9, 1012.1, 840.4, and 621.0
the diffusion rate. Therefore, the bands of the complexes and cm−1) and R4 (at 1144.1, 1072.1, and 908.5 cm−1); see Figures
those of the chlorinated byproducts do not change S24 and S25. As expected, we could not identify R2 and R3
considerably during the next phases of the experiments. among the products, which, as discussed above, is consistent
2840 https://doi.org/10.1021/acs.jpca.2c00306
J. Phys. Chem. A 2022, 126, 2832−2844
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

Table 7. Computed Anharmonic and Experimental


Vibrational Wavenumbers (in cm−1) and IR Intensities (in
km mol−1) of 2,3-Dihydrofuran (2)
computed (C1) exptl
a
B3LYP/cc-pVTZ gas phase para-H2 matrixb assignment
707.0 (36) 705.0 705.3, 706.5 (ms)c ν24
826.3 (3) 828.4 828.4 (m) ν23
831.1 (3) 885.2 886.5 (m) ν22
902.2 (141) 916 913 (m, br) ν21
928.4 (42) 919 (sh) ν18 + ν27
903.9 (54) 925 921.8 (ms) ν20
910.9 (18) 925.9 (ms) ν19
981.2 (3) 996 996.4 (ms) ν18
1060.9 (24) 1063 1065.6 (s) ν17
1052.7 (11) 1061.3 (m) ν16
1142.2 (33) 1141.6 1140.6 (s) ν15
1183.2 (3) 1188.6 1189.5 (w) ν14
1225.1 (1) 1226.7 (w) ν13
1271.5 (15) 1280.5 (mw) ν12 Figure 8. Temporal profiles of the estimated mixing ratios of 2, R1,
1377.7 (3) 1378.0 1377.5 (mw) ν10 R2, R3, and R4 after 365 nm irradiation of 2,3-dihydrofuran (2) and
1459.5 (<1) 1458.7 1457.9 (w) ν9 Cl2 codeposited in para-H2 at 3.1 K. The vertical line indicates the
end of the NIR laser irradiation. The estimated mixing ratio of total H
1642.0 (19) 1624.5 1625.3 (vs) ν7
atoms generated was 760 ± 140 ppm. Note that some R1 and R2
2874.6 (27) 2872.6 (ms) ν5
were produced before the NIR laser irradiation during the 365 nm
2891.8 (82) 2900.7, 2903.9 (m)c ν4 photolysis.
2943.5 (39) 2930.4 2936.3 (m) ν3
3101.6 (7) 3110.0 3105.3, 3107.7 (w)c ν1
a
Ref 57. Bands are reassigned. bThis work. cSplit band.

Table 8. Computed Anharmonic and Experimental


Vibrational Wavenumbers (in cm−1) and IR Intensities (in
km mol−1) of 2,5-Dihydrofuran (3)
computed (Cs) exptl
a
B3LYP/cc-pVTZ gas phase para-H2 matrixb assignment
673.5 (33) 661.2 663.7, 664.7 (m)c ν18
741.5 (7) 739.7 740.9 (m) ν9
800.8 (4) 801.0 801.4 (mw) ν27
888.1 (20) 901.5 897.8 (s) ν8
968.2 (7) 982.6 982.9 (m) ν7
1007.8 (19) 1011.2 1010.0 (s) ν17
1064.3 (89) 1093.0 1090.8 (vs) ν25
1152.8 (2) 1166.8 1167.1 (w) ν16
1303.5 (<1) 1306.0 1305.7 (vw) ν24
1341.8 (4) 1349.0 1349.4 (m) ν23
1354.4 (3) 1362.0 1363.0 (w) ν5 Figure 9. Temporal profiles of the estimated mixing ratios of 3, R1,
and R4 after 365 nm irradiation of 2,5-dihydrofuran (3) and Cl2
1481.7 (1) 1489.0 1487.8 (mw) ν4
codeposited in para-H2 at 3.1 K. The vertical line indicates the end of
2812.5 (151) 2864.4 2856.4 (s) ν21
the NIR laser irradiation. The estimated mixing ratio of total H atoms
2827.9 (113) 2883.0 2883.7 (ms) ν15 generated was 580 ± 100 ppm. Note that some R1 was produced
3081.2 (4) 3091.0 3086.9, 3088.4 (w)c ν20 before the NIR laser irradiation during the 365 nm photolysis.
3087.6 (10) 3095.8 3094.7 (w) ν1


a
Ref 57. Some bands are reassigned. bThis work. cSplit band.
CONCLUSIONS AND ASTROCHEMICAL
with the topological expectations. The formation of 1 and 2 is RELEVANCE
highly uncertain; most band intensities of these species are The present and other recent studies13,14,29,33−39,58−63
close to the noise level (Figure S26). demonstrate that astrochemically relevant H atom addition
The temporal profiles of the mixing ratios for these and abstraction reactions can be effectively studied by the
experiments are plotted in Figures 8, 9, and S27−S29. The para-H2 matrix isolation method. Although the para-H2 matrix
species that are formed in very small amounts are omitted from environment used for the current study is apolar and colder
these figures. Similar to the experiments with 1 and 4, the (<4 K) than typical interstellar ice environments, if a reaction
increase of the mixing ratios of the radicals during the dark proceeds under the investigated conditions, the reaction
period clearly proves that these reactions can proceed channel is very likely open under astrophysical conditions.
effectively at 3.1 K. Although in real astrophysical environments at typically higher
2841 https://doi.org/10.1021/acs.jpca.2c00306
J. Phys. Chem. A 2022, 126, 2832−2844
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

temperatures (>10 K) the reaction rates, including the relative The present work and that of Amicangelo and Lee29 clearly
rates of the H atom addition and abstraction processes, can be prove experimentally the former theoretical predictions that
different, this method provides an important tool for heterocycles can effectively catalyze the formation of H2. Thus,
monitoring whether the reaction is feasible at low temperatures heterocycles might play a more important role in interstellar
either because of tunneling or because it is a barrierless H2 formation than PAHs. Since all the processes are taking
process. It also provides a direct comparison with quantum place in both the H atom addition and the H atom abstraction
chemical computations performed for isolated molecules. The directions, a quasi-equilibrium exists between each species: 1,
method has a great advantage that, due to sharp, well-resolved 2, 3, and 4 (closed-shell molecules) and R1, R2, R3, and R4
vibrational bands, the assignments are reliable, and the reaction (radicals). Although the equilibrium mixing ratios are expected
mechanisms can straightforwardly be deduced. In addition, it is to depend highly on the temperature, the polarity, and the
very sensitive because H atoms can be generated in a large composition of the ice, calling for the investigation of these
mixing ratio, the H atoms can effectively move in the para-H2 processes in astrophysical analogue ices, the present experi-
matrix by quantum diffusion, and because of the very high ments might allow drawing conclusions relevant to low
signal-to-noise ratio. Therefore, it enables the study of slow temperatures, e.g., for conditions of dark molecular clouds.
processes on the laboratory time scale. In our view, none of the However, in our experiments we were not able to produce
enough H atoms to reach the equilibrium, considering that we
laboratory measurements can simulate perfectly the astrophys-
could clearly observe 2 starting from 1 and 4, while in very
ical conditions; the para-H2 matrix isolation method can be
similar experiments with 2 and 3, we could hardly identify 1
used to sensitively prescreen the reactions, especially those
and 4; it indicates that under equilibrium conditions 2 is the
taking place on the surface of grains. Then, these reactions can most abundant form. This experimental observation, in
be verified by less sensitive experiments performed on the agreement with the computational predictions of Miksch et
surface or in the bulk astrophysical analogue ices. al., 9 might give an explanation for the unsuccessful
In this work, the H atom addition and abstraction reactions attempts24−27 for the astrophysical observations of 1 and
of 1, 2, 3, and 4 were studied by the above-discussed para-H2 calls for radio astronomical search of 2 in the ISM.


matrix isolation method. The analysis of IR spectra taken after
H atom generation during the dark processes revealed that, in
ASSOCIATED CONTENT
agreement with theoretical predictions, 1 and 4 effectively
react with H atoms in H atom addition and H atom * Supporting Information

abstraction processes, respectively. Furthermore, 2 and 3 can The Supporting Information is available free of charge at
take part in both types of H atom reactions. In each case, the https://pubs.acs.org/doi/10.1021/acs.jpca.2c00306.
corresponding radicals, formed in the first step of the reactions,
B3LYP/cc-pVTZ structures, energies, and vibrational
were identified. In addition, in the case of 1 and 4, the
frequencies and full-scale spectra and temporal profiles
products of the second step of the reactions (i.e., a second H
of some parallel experiments with different mixing ratios
atom addition and second H atom abstraction, respectively) (PDF)
could also be unambiguously identified. These processes are


summarized in Figure 10.
AUTHOR INFORMATION
Corresponding Author
György Tarczay − MTA-ELTE Lendület Laboratory
Astrochemistry Research Group, Institute of Chemistry,
Laboratory of Molecular Spectroscopy, Institute of Chemistry,
and Centre for Astrophysics and Space Science, ELTE Eötvös
Loránd University, H-1518 Budapest, Hungary;
orcid.org/0000-0002-2345-1774;
Email: gyorgy.tarczay@ttk.elte.hu
Authors
Anita Schneiker − George Hevesy Doctoral School, ELTE
Eötvös Loránd University, H−1518 Budapest, Hungary;
MTA-ELTE Lendület Laboratory Astrochemistry Research
Group, Institute of Chemistry, ELTE Eötvös Loránd
University, H-1518 Budapest, Hungary
Gopi Ragupathy − MTA-ELTE Lendület Laboratory
Astrochemistry Research Group, Institute of Chemistry, ELTE
Eötvös Loránd University, H-1518 Budapest, Hungary
Gábor Bazsó − Wigner Research Centre for Physics, H-1525
Budapest, Hungary
Complete contact information is available at:
Figure 10. H atom addition and abstraction reactions connecting 1
with 4 via R1, R2, 2, 3, R3, and R4. The thick reaction arrows
https://pubs.acs.org/10.1021/acs.jpca.2c00306
represent the paths which were observed experimentally. The thin
arrows are the ones that are predicted to proceed theoretically but Notes
could not be observed directly in the experiments. The authors declare no competing financial interest.
2842 https://doi.org/10.1021/acs.jpca.2c00306
J. Phys. Chem. A 2022, 126, 2832−2844
The Journal of Physical Chemistry A


pubs.acs.org/JPCA Article

ACKNOWLEDGMENTS (18) Skov, A. L.; Thrower, J. D.; Hornekær, L. L. Polycyclic


Aromatic Hydrocarbons − Catalysts for Molecular Hydrogen
The support of the Lendület program of the Hungarian Formation. Faraday Discuss. 2014, 168, 223−234.
Academy of Sciences is acknowledged. The authors acknowl- (19) Petrie, S.; Javahery, G.; Bohme, D. K. Gas-Phase Reactions of
edge Dr. Sándor Góbi for helpful discussions. This work was Benzenoid Hydrocarbon Ions with Hydrogen Atoms and Molecules:
also supported by the ELTE Institutional Excellence Program Uncommon Constraints to Reactivity. J. Am. Chem. Soc. 1992, 114,
(Grant TKP2021-NKTA-64). 9205−9206.


(20) Scott, G. B.; Fairley, D. A.; Freeman, C. G.; McEwan, M. J.;
Adams, N. G.; Babcock, L. M. CmHn+ Reactions with H and H2: An
REFERENCES Experimental Study. J. Phys. Chem. A 1997, 101, 4973−4978.
(1) Williams, D. A.; Hartquist, T. W. The Cosmic-Chemical Bond: (21) Le Page, V.; Keheyan, Y.; Bierbaum, V. M.; Snow, T. P.
Chemistry from the Big Bang to Planet Formation; RSC Publishing: Chemical Constraints on Organic Cations in the Interstellar Medium.
London, U.K., 2013. J. Am. Chem. Soc. 1997, 119, 8373−8374.
(2) Vidali, G. H2 Formation on Interstellar Grains. Chem. Rev. 2013, (22) Snow, T. P.; Le Page, V.; Keheyan, Y.; Bierbaum, V. M. The
113, 8762−8782. Interstellar Chemistry of PAH Cations. Nature 1998, 391, 259−260.
(3) Wakelam, V.; Bron, E.; Cazaux, S.; Dulieu, F.; Gry, C.; Guillard, (23) Barrales-Martínez, C.; Gutiérrez-Oliva, S. The Effect of
P.; Habart, E.; Hornekær, L.; Morisset, S.; Nyman, G.; et al. H2 Heteroatoms in Carbonaceous Surfaces: Computational Analysis of
Formation on Interstellar Dust Grains: The Viewpoints of Theory, H Chemisorption on to a PANH and Si-Doped PAH. Mon. Not. R.
Experiments, Models and Observations. Mol. Astrophys 2017, 9, 1− Astron. Soc. 2019, 490, 172−180.
36. (24) Dezafra, R. L.; Thaddeus, P.; Kutner, M.; Scoville, N.;
(4) Rauls, E.; Hornekær, L. Catalyzed Routes to Molecular Solomon, P. M.; Weaver, H.; Williams, D. R. W. Search for Interstellar
Hydrogen Formation and Hydrogen Addition Reactions on Neutral Furan and Imidazole. Astrophys. Lett. 1972, 10, 1−3.
Polycyclic Aromatic Hydrocarbons Under Interstellar Conditions. (25) Kutner, M. L.; Machnik, D. E.; Tucker, K. D.; Dickman, R. L.
Astrophys. J. 2008, 679, 531. Search for Interstellar Pyrrole and Furan. Astrophys. J. 1980, 242,
(5) Bonfanti, M.; Casolo, S.; Tantardini, G. F.; Martinazzo, R. 541−544.
Surface Models and Reaction Barrier in Eley−Rideal Formation of H2 (26) Lattelais, M.; Ellinger, Y.; Matrane, A.; Guillemin, J.-C. Looking
on Graphitic Surfaces. Phys. Chem. Chem. Phys. 2011, 13, 16680− for Heteroaromatic Rings and Related Isomers as Interstellar
16688. Candidates. Phys. Chem. Chem. Phys. 2010, 12, 4165−4171.
(6) Barrales-Martínez, C.; Cortés-Arriagada, D.; Gutiérrez-Oliva, S. (27) Barnum, T. J.; Lee, K. L. K.; McGuire, B. A. Chirped-Pulse
Molecular Hydrogen Formation in the Interstellar Medium: the Role Fourier Transform Millimeter-Wave Spectroscopy of Furan, Isotopo-
of Polycyclic Aromatic Hydrocarbons Analysed by the Reaction Force logues, and Vibrational Excited States. ACS Earth Space Chem. 2021,
and Activation Strain Model. Mon. Not. R. Astron. Soc. 2018, 481, 5, 2986−2994.
3052−3062. (28) Etim, E. E.; Adelagun, R. O. A.; Andrew, C.; Oluwole, O. E.
(7) Cazaux, S.; Tielens, A. G. G. M. H2 Formation on Grain Optimizing the Searches for Interstellar Heterocycles. Adv. Space Res.
Surfaces. Astrophys. J. 2004, 604, 222. 2021, 68, 3508−3520.
(8) Goumans, T. P. M.; Kästner, J. Hydrogen-Atom Tunneling (29) Amicangelo, J. C.; Lee, Y. P. Hydrogenation of Pyrrole: Infrared
Could Contribute to H2 Formation in Space. Angew. Chem. Int. Ed Spectra of the 2,3-Dihydropyrrol-2-yl and 2,3-Dihydropyrrol-3-yl
2010, 49, 7350−7352. Radicals Isolated in Solid Para-Hydrogen. J. Chem. Phys. 2020, 153,
(9) Miksch, A. M.; Riffelt, A.; Oliveira, R.; Kästner, J.; Molpeceres, 164302.
G. Hydrogenation of small aromatic heterocycles at low temperatures. (30) Kufeld, K. A.; Wonderly, W. R.; Paulson, L. O.; Kettwich, S. C.;
Mon. Not. R. Astron. Soc. 2021, 505, 3157−3164. Anderson, D. T. Transient H2O Infrared Satellite Peaks Produced in
(10) Goumans, T. P. M. Hydrogen Chemisorption on Polycyclic UV Irradiated Formic Acid Doped Solid Parahydrogen. J. Phys. Chem.
Aromatic Hydrocarbons via Tunneling. Mon. Not. R. Astron. Soc. Lett. 2012, 3, 342−347.
2011, 415, 3129−3134. (31) Ruzi, M.; Anderson, D. T. Quantum Diffusion-Controlled
(11) Mennella, V.; Hornekær, L.; Thrower, J.; Accolla, M. The Chemistry: Reactions of Atomic Hydrogen with Nitric Oxide in Solid
Catalytic Role of Coronene for Molecular Hydrogen Formation. Parahydrogen. J. Phys. Chem. A 2015, 119, 12270−12283.
Astrophys. J. Lett. 2012, 745, L2. (32) Balabanoff, M. E.; Ruzi, M.; Anderson, D. T. Signatures of a
(12) Schneiker, A.; Csonka, I. P.; Tarczay, G. Hydrogenation and Quantum Diffusion Limited Hydrogen Atom Tunneling Reaction.
Dehydrogenation Reactions of the Phenalenyl Radical/1H-Phenalene Phys. Chem. Chem. Phys. 2018, 20, 422−434.
System at low Temperatures. Chem. Phys. Lett. 2020, 743, 137183. (33) Bahou, M.; Wu, Y. J.; Lee, Y. P. Infrared Spectra of Protonated
(13) Haupa, K. A.; Tarczay, G.; Lee, Y.-P. Hydrogen Abstraction/ Coronene and Its Neutral Counterpart in Solid Parahydrogen:
Addition Tunneling Reactions Elucidate the Interstellar H2NCHO/ Implications for Unidentified Interstellar Infrared Emission Bands.
HNCO Ratio and H2 Formation. J. Am. Chem. Soc. 2019, 141, Angew. Chem., Int. Ed 2014, 53, 1021−1024.
11614−11620. (34) Tsuge, M.; Tseng, C. Y.; Lee, Y. P. Spectroscopy of Prospective
(14) Schneiker, A.; Góbi, S.; Joshi, P. R.; Bazsó, G.; Lee, Y. P.; Interstellar Ions and Radicals Isolated in Para-Hydrogen Matrices.
Tarczay, G. Non-Energetic, Low-Temperature Formation of Cα- Phys. Chem. Chem. Phys. 2018, 20, 5344−5358.
Glycyl Radical, a Potential Interstellar Precursor of Natural Amino (35) Sundararajan, P.; Tsuge, M.; Baba, M.; Sakurai, H.; Lee, Y. P.
Acids. J. Phys. Chem. Lett. 2021, 12, 6744−6751. Infrared Spectrum of Hydrogenated Corannulene Rim-HC20H10
(15) Habart, E.; Boulanger, F.; Verstraete, L.; Pineau des Forêts, G.; Isolated in Solid Para-Hydrogen. J. Chem. Phys. 2019, 151, 044304.
Falgarone, E.; Abergel, A. H2 Infrared Line Emission Across the (36) Amicangelo, J. C.; Lee, Y. P. Infrared Spectra of the 1,1-
Bright Side of the ρ Ophiuchi Main Cloud. Astron. Astrophys. 2003, Dimethylallyl and 1,2-Dimethylallyl Radicals Isolated in Solid Para-
397, 623−634. Hydrogen. J. Chem. Phys. 2018, 149, 204304.
(16) Habart, E.; Boulanger, F.; Verstraete, L.; Walmsley, C. M.; (37) Tarczay, G.; Haupa, K.; Lee, Y. P. On the correlation of the
Pineau des Forêts, G. Some Empirical Estimates of the H2 Formation abundances of HNCO and NH2CHO: Advantages of solid para-H2 to
Rate in Photon-Dominated Regions. Astron. Astrophys. 2004, 414, study astrochemical H-atom addition and abstraction reactions. Proc.
531−544. Int. Astron. Union 2019, 15, 394−396.
(17) Le Page, V.; Snow, T. P.; Bierbaum, V. Molecular Hydrogen (38) Bazsó, G.; Csonka, I. P.; Góbi, S.; Tarczay, G. VIZSLA
Formation Catalyzed by Polycyclic Aromatic Hydrocarbons in the Versatile Ice Zigzag Sublimation Setup for Laboratory Astrochemistry.
Interstellar Medium. Astrophys. J. 2009, 704, 274−280. Rev. Sci. Instrum. 2021, 92, 124104.

2843 https://doi.org/10.1021/acs.jpca.2c00306
J. Phys. Chem. A 2022, 126, 2832−2844
The Journal of Physical Chemistry A pubs.acs.org/JPCA Article

(39) Tsuge, M.; Lee, Y.-P. Spectroscopy of Molecules Confined in Atom with Acetamide (CH 3 CONH 2 ) and Photolysis of
Solid Para-Hydrogen. In Molecular and Laser Spectroscopy; Gupta, V. •CH2CONH2 to form Ketene (CH2CO) in Solid para-Hydrogen.
P., Ozaki, Y., Eds.; Elsevier: Amsterdam, Netherlands, 2020; pp 167− Phys. Chem. Chem. Phys. 2020, 22, 6192−6201.
215. (61) Joshi, P. R.; How, K. C. Y.; Lee, Y. P. Hydrogen Abstraction of
(40) Raston, P. L.; Kettwich, S. C.; Anderson, D. T. Infrared Studies Acetic Acid by Hydrogen Atom to Form Carboxymethyl Radical
of Ortho-Para Conversion at Cl-atom and H-atom Impurity Centers •CH2C(O)OH in Solid para-Hydrogen and Its Implication in
in Cryogenic Solid Hydrogen. Low Temp. Phys. 2010, 36, 392−399. Astrochemistry. ACS Earth Space Chem. 2021, 5, 106−117.
(41) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; (62) Góbi, S.; Csonka, I. P.; Bazsó, G.; Tarczay, G. Successive
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, Hydrogenation of SO and SO2 in Solid para-H2: Formation of Elusive
B.; Petersson, G. A., et al. Gaussian 09, rev. D.01; Gaussian Inc.: Small Oxoacids of Sulfur. ACS Earth Space Chem. 2021, 5, 1180−
Wallingford, CT, 2013. 1195.
(42) Becke, A. D. Density-functional Thermochemistry. III. The (63) Feldman, V. I.; Ryazantsev, S. V.; Kameneva, S. V. Matrix
Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648−5652. Isolation in Laboratory Astrochemistry: State-of-the-art, Implications
(43) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle- and Perspective. Russ. Chem. Rev. 2021, 90, 1142−1165.
Salvetti Correlation-Energy Formula into a Functional of the Electron
Density. Phys. Rev. B 1988, 37, 785−789.
(44) Dunning, T. H. Gaussian Basis Sets for Use in Correlated
Molecular Calculations. I. The Atoms Boron through Neon and
Hydrogen. J. Chem. Phys. 1989, 90, 1007−1023.
(45) Fukui, K. The Path of Chemical Reactions − The IRC
Approach. Acc. Chem. Res. 1981, 14, 363−368.
(46) Nielsen, H. H. The Vibration−Rotation Energies of Molecules.
Rev. Mod. Phys. 1951, 23, 90−136.
(47) Allen, W. D.; Yamaguchi, Y.; Császár, A. G.; Clabo, D. A., Jr.;
Remington, R. B.; Schaefer, H. F. III. A Systematic Study of Molecular
Vibrational Anharmonicity and Vibration-Rotation Interaction by
Self- Consistent-Field Higher-Derivative Methods. Linear Polyatomic
Molecules. Chem. Phys. 1990, 145, 427−466.
(48) Shimanouchi, T. Tables of Molecular Vibrational Frequencies
Consolidated, Vol. I; National Bureau of Standards, U.S. Government
Printing Office: Washington, DC, 1972, pp 1−160.
(49) Sánchez-García, E.; Mardyukov, A.; Tekin, A.; Crespo-Otero,
R.; Montero, L. A.; Sander, W.; Jansen, G. Ab Initio and Matrix
Isolation Study of the Acetylene−Furan Dimer. Chem. Phys. 2008,
343, 168−185.
(50) Sánchez-García, E.; Mardyukov, A.; Studentkowski, M.;
Montero, L. A.; Sander, W. Furan−Formic Acid Dimers: An ab Recommended by ACS
Initio and Matrix Isolation Study. J. Phys. Chem. A 2006, 110, 13775−
13785. Quantum Tunneling Mediated Low-Temperature Synthesis
(51) Lockwood, S. P.; Fuller, T. G.; Newby, J. J. Structure and of Interstellar Hemiacetals
Spectroscopy of Furan: H2O Complexes. J. Phys. Chem. A 2018, 122,
7160−7170. Jia Wang, Ralf I. Kaiser, et al.
(52) Bahou, M.; Huang, C.-W.; Huang, Y.-L.; Glatthaar, J.; Lee, Y.-P. JUNE 26, 2023
THE JOURNAL OF PHYSICAL CHEMISTRY LETTERS READ
Advances in Use of p-H2 as a Novel Host for Matrix IR Spectroscopy.
J. Chin. Chem. Soc. 2010, 57, 771−782.
(53) Lai, L. H.; Liu, C. P.; Lee, Y. P. Laser Photolysis of OClO in New Insights into the Formation of CH3OCH3 and CH3SCH3
Solid Ne, Ar, and Kr. II. Site Selectivity, Mode Specificity, and Effects without and with the Assistance of Na+ Ions and Some
of Matrix Hosts. J. Chem. Phys. 1998, 109, 988. Implications for Interstellar Chemistry: An In Silic...
(54) Jacox, M. E.; Milligan, D. E. Spectrum and Structure of the Sarvesh Kumar Pandey, Elangnnan Arunan, et al.
HO2 Free Radical. J. Mol. Spectrosc. 1972, 42, 495−513. FEBRUARY 06, 2023
(55) Evseeva, L. A.; Sverdlov, L. M. Analysis and Interpretation of ACS EARTH AND SPACE CHEMISTRY READ
Vibrational Spectra of Tetrahydrofuran and its Deutero Derivatives.
Sov. Phys. J. 1968, 11, 87−90. 1,3-Butadiene on Titan: Crystal Structure, Thermal
(56) Cadioli, B.; Gallinella, E.; Coulombeau, C.; Jobic, H.; Berthier, Expansivity, and Raman Signatures
G. Geometric Structure and Vibrational Spectrum of Tetrahydrofur-
an. J. Phys. Chem. 1993, 97, 7844−7856. Tuan H. Vu, Robert Hodyss, et al.
OCTOBER 05, 2022
(57) Klots, T. D.; Collier, W. B. Vibrational Assignment and Analysis
ACS EARTH AND SPACE CHEMISTRY READ
for 2,3-Dihydrofuran and 2,5-Dihydrofuran. Spectrochim. Acta Part A:
Mol. Spectrosc 1994, 50, 1725−1748.
(58) Haupa, K. A.; Tielens, A. G. G. M.; Lee, Y. P. Reaction of H + Time-Dependent Density Functional Study of Nitrogen-
HONO in Solid para-Hydrogen: Infrared Spectrum of •ONH(OH). Substituted Polycyclic Aromatic Hydrocarbons and Diffuse
Phys. Chem. Chem. Phys. 2017, 19, 16169−16177. Interstellar Bands
(59) Haupa, K. A.; Strom, A. I.; Anderson, T. A.; Lee, Y. P. Nishant Shukla, Gazi A. Ahmed, et al.
Hydrogen-Atom Tunneling Reactions with Methyl Formate in Solid NOVEMBER 21, 2022
para-Hydrogen: Infrared Spectra of the Methoxy Carbonyl [•C(O)- ACS EARTH AND SPACE CHEMISTRY READ
OCH3] and Formyloxy Methyl [HC(O)OCH2•] Radicals. J. Chem.
Phys. 2019, 151, 234302.
Get More Suggestions >
(60) Haupa, K. A.; Ong, W. S.; Lee, Y. P. Hydrogen Abstraction in
Astrochemistry: Formation of •CH2CONH2 in the Reaction of H

2844 https://doi.org/10.1021/acs.jpca.2c00306
J. Phys. Chem. A 2022, 126, 2832−2844

You might also like