Download as pdf or txt
Download as pdf or txt
You are on page 1of 56

Subscriber access provided by UNIV TEXAS MEDICAL BRANCH

A: Spectroscopy, Molecular Structure, and Quantum Chemistry


Infrared Spectroscopy of Protonated Phenol-Water Clusters
Marusu Katada, and Asuka Fujii
J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.8b04446 • Publication Date (Web): 20 Jun 2018
Downloaded from http://pubs.acs.org on June 23, 2018

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a service to the research community to expedite the dissemination
of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in
full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully
peer reviewed, but should not be considered the official version of record. They are citable by the
Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore,
the “Just Accepted” Web site may not include all articles that will be published in the journal. After
a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web
site and published as an ASAP article. Note that technical editing may introduce minor changes
to the manuscript text and/or graphics which could affect content, and all legal disclaimers and
ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or
consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W.,


Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 Infrared Spectroscopy of Protonated Phenol-
Phenol-Water Clusters
7
8
9 Marusu Katada and Asuka Fujii*
10
11
Department of Chemistry, Graduate School of Science, Tohoku University,
12
13
14 Sendai 980-8578, Japan
15
16
17 *Corresponding author
18
19 E-mail: asukafujii@m.tohoku.ac.jp
20
21 ORCID Asuka Fujii: 0000-0002-6854-9636
22
23
24
25
26 Abstract
27
28
29 To explore the microhydration structures of protonated phenol, size-selective infrared
30
31
32 spectroscopy of protonated phenol-(water)n clusters (n = 1 – 5) was performed.
33
34
35 Protonation of phenol can occur either at the phenyl ring or the hydroxy group. The
36
37
38 coexistence of the two isomer types separated by the high isomerization barrier was
39
40
reconfirmed for bare protonated phenol. Preferential hydration of the hydroxy group
41
42
43 initially occurs in both the two isomer types of protonated phenol. Development of the
44
45
46 water hydrogen-bond network is localized around the hydroxy group up to n = 2.
47
48
49 Intracluster proton transfer from the phenol moiety to the water moiety was observed in
50
51
52 n ≥ 3~4. The water moiety with the H3O+ ion core resides on the phenyl ring, and the
53
54
55 water moiety is bound to the phenyl ring with a π-hydrogen bond. Such a structure is
56
57 1
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 2 of 55

1
2
3
4
5
6 in striking contrast to those of phenol+-(water)n radical cation clusters, in which the
7
8
9 water moiety is located away from the phenyl ring even when intracluster proton
10
11
transfer occurs.
12
13
14
15
16
17 1. Introduction
18
19
20 Protonated aromatic molecules have attracted much interest in the diverse fields of
21
22
23 chemistry and biochemistry. 1-7 Protonated aromatics are well-known crucial
24
25
26 intermediates in electrophilic aromatic substitution reactions. 1 The presence of
27
28
protonated aromatics in interstellar space has been suggested. 2,3 Moreover, protonated
29
30
31 aromatics are expected to play important roles in biochemical processes since the proton
32
33
34 is one of the major carriers of positive charge in such systems. 4 Therefore,
35
36
37 characterization of protonated aromatic molecules and their microsolvated clusters in
38
39
40 the gas phase has been extensively performed by spectroscopic, mass spectrometric, and
41
42
43 theoretical approaches. 5-45 Not only the structure determination, e.g., identification of
44
45
the protonated site and their microsolvation structures, but also (excited state) proton
46
47
48 transfer mechanisms have been studied.
49
50
51 The simplest prototype of protonated aromatics is protonated benzene. 8-16

52
53
54 Protonated benzene has the arenium (benzenium) ion structure, in which the
55
56
57 2
58
59
60 ACS Paragon Plus Environment
Page 3 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 protonated carbon atom has the sp3 hybridization and the resonance among the π
7
8
9 orbitals is partly broken. This type of protonated species is also called σ-complex or
10
11
Wheland intermediate. Though the proton affinity (PA) of benzene (750 kJ/mol) is larger
12
13
14 than water (PA = 691 kJ/mol), 46 the hydration of protonated benzene with only a single
15
16
17 water molecule induces the proton transfer reaction to form protonated water
18
19
20 (hydronium ion, H3O+). 13 This is due to the large solvation energy of the hydronium
21
22
23 ion by benzene with the strong charge-induced dipole interaction.
24
25
26 On the other hand, in substituted benzene molecules, 19-45 the protonation of the
27
28
substituent competes with the protonation of the phenyl ring, and the former process
29
30
31 can be exclusive when the “local” proton affinity of the substituent is much larger than
32
33
34 that of the phenyl ring moiety. Such a case has been found in benzaldehyde. 27-31 In
35
36
37 this context, protonated phenol, H+PhOH, is of great interest because competition
38
39
40 between the protonation of the substituent (hydroxy (OH) group) and the phenyl ring
41
42
43 has been observed. 19-26 Solcà and Dopfer have applied infrared (IR) dissociation
44
45
spectroscopy to H+PhOH. 19-21 They found that at least two isomers of H+PhOH coexist
46
47
48 under the jet-cooled condition. In the most stable structure, the excess proton is located
49
50
51 at the para (p-) position of the phenyl ring, resulting in an arenium ion. In the second
52
53
54 stable structure, the excess proton is located at the ortho (o-) position of the phenyl ring.
55
56
57 3
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 4 of 55

1
2
3
4
5
6 Protonation of the meta (m-) and ipso positions is much less stable. Since the arenium
7
8
9 ions produced by the protonation at the p- and o-positions are difficult to distinguish by
10
11
spectroscopy, hereafter we call them both Ph-type. Solcà and Dopfer have reported
12
13
14 that the excess proton can also be located at the lone pair electrons of the OH group and
15
16
17 this forms an oxonium ion. Hereafter, this isomer is called O-type. Though the
18
19
20 relative energy of the O- type is over 70 kJ/mol higher than that of the most stable
21
22
23 Ph-type, Solcà and Dopfer have observed the competitive production of the O-type
24
25
26 isomer under the supersonic jet expansion cooling condition, and they attributed it to
27
28
the high potential energy barrier (~160 kJ/mol) in the isomerization from the O-type to
29
30
31 the Ph-type.
32
33
34 The microsolvation of H+PhOH is also of great interest. Solcà and Dopfer
35
36
37 performed IR spectroscopic studies on the microsolvation by inert gas species (Ne, Ar,
38
39
40 and N2). 19-21 They showed that the solvation begins with hydrogen (H-) bond formation
41
42
43 to the acidic OH group(s) and then proceeds to π-bond formation, while intermolecular
44
45
binding to the aliphatic CH2 group in the protonated phenyl ring is not preferred. On
46
47
48 the other hand, studies on the microsolvation of H+PhOH by water have been very
49
50
51 scarce in spite of its special importance in relation to the protonation of tyrosine in
52
53
54 biological environments. H-bonded structures of H+PhOH-(water)n (H+PhOH-Wn)
55
56
57 4
58
59
60 ACS Paragon Plus Environment
Page 5 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 clusters and the size dependence of their intracluster proton transfer have been studied
7
8
9 by quantum chemical computations and energy-resolved collision-induced dissociation
10
11
(CID) measurements. 24, 26 However, the computations were performed for only a small
12
13
14 number of hydrated structures based on the Ph-type isomer. 24 It is also difficult to
15
16
17 definitively infer H-bonded structures of clusters from CID experiments because
18
19
20 rearrangements of clusters can occur prior to dissociation.
21
22
23 On the other hand, neutral phenol-water clusters (PhOH-Wn) and phenol-water
24
25
26 radical cation clusters (PhOH+-Wn) have been extensively studied in the gas phase. 47-66
27
28
Neutral PhOH-Wn has been investigated by size-selective IR spectroscopy as a model of
29
30
31 water H-bond networks. 47-59 It has been demonstrated that the H-bond network of
32
33
34 PhOH-Wn develops from ring structures to cage structures as the cluster size increases.
35
36
37 47-57 Further processes to form bulk-like structures including fully solvated water
38
39
40 molecules have been discussed. 58,59 The H-bonded structures of PhOH+-Wn have also
41
42
43 been investigated by IR and electronic spectroscopies, and the occurrence of intracluster
44
45
proton transfer from the phenol cation to the water moiety has been confirmed in n ≥ 3 ~
46
47
48 4. 60-66 We should note that both the H-bonded networks of neutral and cationic
49
50
51 hydrated phenols develop around the phenolic OH group and a direct interaction
52
53
54 between the water network and the phenyl ring has not yet been observed in these
55
56
57 5
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 6 of 55

1
2
3
4
5
6 systems.
7
8
9 As mentioned above, the two isomer types of H+PhOH coexist under the jet-cooled
10
11
condition. 19-21 The differences in the microhydration structures between these two
12
13
14 types and interconversion between the isomers with the development of hydration are of
15
16
17 great interest. In the present study, we apply IR dissociation spectroscopy to the
18
19
20 H+PhOH-Wn (n = 1- 5) clusters to explore the development of the H-bond network
21
22
23 structures of these clusters with the help of quantum chemical calculations and by
24
25
26 comparison with related systems. The IR spectrum of the H+PhOH monomer was
27
28
measured by the Ar-tagging method to determine the isomer ratio under the
29
30
31 experimental conditions as well as to confirm the presence or absence of radical cation
32
33
34 species. 67-69
35
36
37
38
39
40 2. Experimental
Experimental and Computational Methods
41
42
43 Size-selective IR spectra of H+PhOH-Ar and H+PhOH-W1-5 in the OH and CH stretching
44
45
vibrational region (2800 - 3900 cm-1) were measured by IR dissociation spectroscopy.
46
47
48 The mass selection of the clusters was performed by a tandem-type quadrupole mass
49
50
51 spectrometer. 68, 69 Protonated clusters were generated by electron ionization with an
52
53
54 electron gun (Omegatron Co.) followed by supersonic jet expansion cooling. The sample
55
56
57 6
58
59
60 ACS Paragon Plus Environment
Page 7 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 of phenol (Kanto Chemical Co.) was used without further purification. Phenol was
7
8
9 heated to ~280 K in the sample container of a high pressure supersonic jet valve
10
11
(Even−Lavie valve). 70 The phenol vapor was seeded in the carrier gas of He/Ar/H2
12
13
14 (80:10:10) or He/H2 (95:5) containing trace water vapor. The total stagnation pressure
15
16
17 of the jet expansion was 60-80 atm. The electron beam accelerated at a voltage of 200 V
18
19
20 was crossed with the jet expansion in the collisional region. The protonated cluster of
21
22
23 interest was selected by the first quadrupole mass filter and the cluster was introduced
24
25
26 into the octopole ion guide. The size-selected cluster was irradiated by an IR laser in the
27
28
ion guide. A fragment ion is generated via predissociation when the IR frequency is
29
30
31 resonant on a vibrational transition of the cluster. The second quadrupole mass filter
32
33
34 was set to pass only the fragment ion and the IR spectrum of the cluster was measured
35
36
37 by monitoring the intensity of the fragment ion while scanning the IR frequency. The IR
38
39
40 light was generated by an OPO/OPA system (LaserVision), and the typical IR power was
41
42
∼2 mJ/pulse.
43
44
45
The IR spectrum of the H+PhOH monomer was measured by using the He/Ar/H2
46
47
48 carrier gas and by the Ar-tagging method. On the other hand, the IR spectra of the
49
50
51 H+PhOH-Wn clusters were measured by using the He /H2 carrier gas and by monitoring
52
53
54 the water-loss channel. The dissociation of PhOH was not detected in the present
55
56
57 7
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 8 of 55

1
2
3
4
5
6 measurements.
7
8
9 In the present experiment, the protonated cluster ion intensity was not high
10
11
12 enough to achieve a mass resolution of ∆m/z <1. Under the spectral measurement
13
14 conditions, the mass resolution was ∆m/z ~2 and we could not completely separate the
15
16
17 protonated cluster from the radical cation cluster of the same cluster size. Therefore, we
18
19
20 compared the observed IR spectrum of H+PhOH-Ar with that of PhOH+-Ar, 71,72 and
21
22
23 confirmed that spectral contamination by the radical cations was negligible under the
24
25
26 present cluster source condition.
27
28
Quantum chemical calculations were performed on the H+PhOH-Ar and
29
30
31 H+PhOH-Wn (n = 1 – 5) clusters to evaluate their H-bonded network structures based on
32
33
34 the observed IR spectra. As will be shown later, the interaction between the water
35
36
37 moiety and phenyl ring (π-electrons) is the key to understanding the microhydration
38
39
40 structures of H+PhOH-Wn. Dispersion should be important in such a type of interaction.
41
42
43 Since MP2 level computations of the present clusters were too time-consuming for our
44
45
computational resources, we employed the dispersion-corrected density functional
46
47
48 theory (DFT) method. For each cluster size, energy-optimized structures were
49
50
51 calculated at the ωB97X-D/6-311++G(d, p) level, which is expected to have a good
52
53
54 balance between computational cost and accuracy. The relative energies of the stable
55
56
57 8
58
59
60 ACS Paragon Plus Environment
Page 9 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 structures with the zero-point energy correlation were evaluated and their IR spectra
7
8
9 were simulated under the harmonic oscillator approximation. The scaling factor was set
10
11
to 0.932 to fit the calculated OH stretch frequency of the most stable isomer of
12
13
14 H+PhOH-Ar (Ph-type) to the most intense band in the observed IR spectrum. The
15
16
17 atomic charge was evaluated by Mulliken population analysis to examine the origin of
18
19
20 the free OH shift of the water moiety. All calculations were performed by the Gaussian
21
22
23 16 program suite. 73
24
25
26
27
28
3. Results and Discussion
29
30
31 A. H+PhOH-
PhOH-Ar
32
33
34 Figure 1(a) shows the observed IR spectrum of H+PhOH-Ar produced in the present
35
36
37 cluster source. The band at 2863 cm-1 is attributed to the aliphatic CH stretch, 5, 7, 9-11,

38
39
40 19-21 and this band indicates the production of the (p- and/or o-) Ph-type isomer(s), in
41
42
43 which the protonation occurs on the phenyl ring. On the other hand, the multiple bands
44
45
in the OH stretch region demonstrate the coexistence of the O-type isomer, which is
46
47
48 protonated at the hydroxy group. The observed bands are the same as those previously
49
50
51 reported by Solcà and Dopfer, 19-21 but the intensity distribution is somewhat different,
52
53
54 reflecting the different isomer population ratio.
55
56
57 9
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 10 of 55

1
2
3
4
5
6 Figures 1(c)-(e) represent the IR spectral simulations based on the Ph-type (p-),
7
8
9 Ph-type (o-), and O-type isomers, respectively. The most intense peak at 3499 cm-1 in
10
11
the observed spectrum is assigned to the Ar-bound OH stretch of the Ph-type isomer.
12
13
14 The Ph-type (p-) isomer is much more stable than the Ph-type (o-) isomer, but a high
15
16
17 isomerization barrier (~90 kJ/mol) is predicted for these isomers by Solcà and Dopfer.
18
19
20 19-21 The Ph-type (p-) and Ph-type (o-) isomers show almost identical spectra, and we
21
22
23 cannot experimentally distinguish between them. Though the coexistence of these two
24
25
26 isomers is not excluded, we focused only on the more stable Ph-type (p-) isomer in the
27
28
following discussions for simplicity. The relatively broadened band at 3336 cm-1 and
29
30
31 the second intense band at 3533 cm-1 are attributed to the Ar-bound and free OH
32
33
34 stretches of the O-type isomer, respectively. A weak band is seen at 3554 cm-1 in the
35
36
37 observed spectrum. Solcà and Dopfer have assigned this band to the free OH stretch of
38
39
40 the Ph-type isomer with the π-bound Ar tag. 19-21
41
42
43 The relative energy of the O-type is much higher (+73 kJ/mol) than the Ph-type,
44
45
but the O-type clearly dominates. This is due to the high barrier in the isomerization
46
47
48 process, as discussed by Solcà and Dopfer. 19-21 Figure 1(b) shows the Lorentzian fitting
49
50
51 of the observed spectrum after slight smoothing (the upper trace shows the residual in
52
53
54 the fitting). On the basis of the integrated band intensities estimated by the fitting
55
56
57 10
58
59
60 ACS Paragon Plus Environment
Page 11 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 and the calculated transition intensities, the relative population of the Ph-type (p- and
7
8
9 o-) and O-type is evaluated as ~4:1. In this measurement, we employed the jet
10
11
expansion of the stagnation pressure at 60 atm, whereas Solcà and Dopfer reported a
12
13
14 ratio of ~5:1 in the jet expansion (with electron ionization) at 12 atm. 20 The higher
15
16
17 production yield of the less stable O-type isomer under the higher pressure jet
18
19
20 expansion condition might be due to kinetic trapping to the local minimum with the
21
22
23 rapid cooling process.
24
25
26 Since the mass resolution of the present experimental condition was not high
27
28
enough (∆m/z ~2) to completely separate H+PhOH-Ar from the PhOH+-Ar radical cation,
29
30
31 we examined its spectral contamination by comparison of the observed IR spectra. The
32
33
34 IR spectrum of PhOH+-Ar produced by an electron ionization source is shown in Fig. 1(f),
35
36
37 which is taken from ref. 72. It has been known that PhOH+-Ar has two stable isomers
38
39
40 with the Ar-binding sites. The Ar-bound OH stretch band of PhOH+-Ar is seen at 3464
41
42
43 cm-1 in the H-bonded isomer, while the band at 3536 cm-1 is attributed to the free OH
44
45
stretch of the π-bonding isomer. 71, 72 With the electron ionization at the collisional
46
47
48 region of the jet expansion, the more stable H-bonded isomer is the major product. 72
49
50
51 In the present spectrum (Fig. 1(a)), the Ar-bound OH stretch band of PhOH+-Ar at 3464
52
53
54 cm-1 is nearly absent. Therefore, we conclude that the spectral contamination by the
55
56
57 11
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 12 of 55

1
2
3
4
5
6 radical cation species is essentially negligible in the present measurements. In our
7
8
9 previous study of the PhOH+-Wn radical cation clusters, the ions were produced by
10
11
resonance multiphoton ionization of the phenol moiety and low pressure jet expansion
12
13
14 with the neat He carrier gas, and the radical cation species were dominant products. 61
15
16
17 In the present condition, we employed electron ionization and high pressure jet
18
19
20 expansion, and also added H2 to the carrier gas to promote the protonation. These
21
22
23 factors would be effective to efficiently quench the radical cation species. In the
24
25
26 following discussion, we ignore the contribution of the radical cations and focus on the
27
28
protonated species.
29
30
31 B. H+PhOH-
PhOH-W1
32
33
34 Figure 2(a) shows the observed IR spectrum of H+PhOH-W1 in the OH and CH stretch
35
36
37 region. In the spectrum, the CH stretch band of the aliphatic CH2 group appears at 2864
38
39
40 cm-1, which is essentially the same frequency as that in the monomer. Moreover, a
41
42
43 remarkably broadened absorption is seen around 2900 cm-1 and this band is attributed
44
45
to a H-bonded OH stretch. These spectral features suggest a cluster structure in which
46
47
48 the Ph-type isomer is solvated at the OH group by a water molecule. The simulated IR
49
50
51 spectrum and schematic structure of the minimum energy structure is shown in Fig.
52
53
54 2(c). The phenolic OH group of the Ph-type isomer acts as a proton donor to the water
55
56
57 12
58
59
60 ACS Paragon Plus Environment
Page 13 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 moiety. This structure is essentially the same as that proposed in the previous CID and
7
8
9 theoretical studies, 24, 26 and its simulated spectrum well reproduces the observed
10
11
spectrum in the region below 3500 cm-1. However, the observed spectrum shows three
12
13
14 free OH stretch bands at 3585, 3626, and 3713 cm-1. While the latter two bands can be
15
16
17 assigned to the symmetric and antisymmetric (ν1 and ν3) stretching vibrations of the
18
19
20 water moiety, the former band at 3585 cm-1 cannot be attributed to the most stable
21
22
23 Ph-type isomer. Therefore, this band indicates the coexistence of the O-type isomer
24
25
26 shown in Fig. 2(d), in which the O-type monomer is hydrated at the protonated phenolic
27
28
hydroxy group. Its free phenolic OH would be the spectral carrier of the 3585 cm-1
29
30
31 band (the H-bonded phenolic OH stretch band is calculated below 2600 cm-1). The
32
33
34 relative energy of the O-type isomer is much higher (+43.1 kJ/mol) than the most stable
35
36
37 Ph-type isomer. However, this large energy difference essentially comes from that
38
39
40 between the monomer isomers (here, we should note that the energy difference in the
41
42
43 hydrated clusters is significantly reduced, indicating a stronger interaction with water
44
45
in the O-type isomer). The coexistence of the monomer isomers by the high potential
46
47
48 barrier rationalizes the coexistence of the hydrated isomers. We estimated the
49
50
51 population ratio of the Ph- and O-type isomers based on the band area evaluation by
52
53
54 Lorentzian fitting (Fig. 2(b)) and the computed vibrational transition intensities. As a
55
56
57 13
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 14 of 55

1
2
3
4
5
6 result, we evaluated the ratio between the Ph- and O-types to be ~4:1. This ratio is the
7
8
9 same as that estimated for the isomers of H+PhOH-Ar, and suggests that the high
10
11
isomerization barrier between the Ph-type and O-type is still present in H+PhOH-W1.
12
13
14 The binding of the water molecule to the protonated CH2 moiety in the phenyl
15
16
17 ring also forms a few stable structures. In such a structure, the water molecule becomes
18
19
20 a single acceptor and therefore such a structure can be called CH-O type. The most
21
22
23 stable structure among this type of isomer and its IR spectral simulation are shown in
24
25
26 Fig. 2(e). This isomer is, however, much higher in energy (+38.4 kJ/mol) than the most
27
28
stable Ph-type isomer. In the CH-O type isomer, the strong H-bonded aliphatic CH
29
30
31 stretch band is predicted at ~2600 cm-1 (the lower frequency side of the free aliphatic
32
33
34 CH stretch). However, the observed broadened absorption is well represented by a
35
36
37 single peak centered at ~2900 cm-1. Moreover, the free phenolic OH stretch frequency
38
39
40 of the CH-O type isomer is reasonably expected to be very close to that in π-bound
41
42
43 H+PhOH-Ar (Ph-type), which is 3554 cm-1 in the observed spectrum (see Fig. 1(a)).
44
45
However, such a band is missing in the observed spectrum. Therefore, we conclude
46
47
48 that the production of this type of isomer is negligible. This is also consistent with the
49
50
51 results of the microsolvation of H+PhOH by inert gas species; the solvation of the OH
52
53
54 group is far superior to that of the CH2 group. 19-21 The absence of the CH-O type
55
56
57 14
58
59
60 ACS Paragon Plus Environment
Page 15 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 isomer indicates that the barrier to the water transfer is lower than that of the proton
7
8
9 transfer.
10
11
In the observed spectrum, weak bands are seen above 3750 cm-1. The
12
13
14 frequencies of these bands are too high to be fundamental vibrations. Therefore, these
15
16
17 bands are attributed to combination bands of the water free OH bands and
18
19
20 intermolecular vibrations.
21
22
23 The observed H-bonded structures of H+PhOH-W1 are similar to those reported
24
25
26 for neutral PhOH-W1 and PhOH+-W1. 48, 49, 51, 55, 61-65 In these clusters, the water
27
28
molecule acts as a proton acceptor to the phenolic OH. The ν1 and ν3 frequencies of the
29
30
31 water moiety are 3650 and 3748 cm-1 in neutral PhOH-W1, 48 while they are 3626 and
32
33
34 3709 cm-1 in PhOH+-W1, respectively. 64 The small red-shifts of the bands in PhOH+-W1
35
36
37 are attributed to the influence of the excess charge, which also clearly enhances the
38
39
40 intensity of the ν1 band, whose transition dipole moment is parallel to the electric field
41
42
43 of the excess charge. The ν1 and ν3 frequencies of H+PhOH-W1 (Ph-type) are 3626 and
44
45
3713 cm-1, respectively, and they are very close to those of PhOH+-W1. The charge
46
47
48 distribution of H+PhOH estimated by the Mulliken population analysis is discussed in
49
50
51 the Supporting Information (SI). The Mulliken charge (0.264) of the OH group of
52
53
54 H+PhOH (Ph-type) is much closer to that of PhOH+ radical cation (0.29) than neutral
55
56
57 15
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 16 of 55

1
2
3
4
5
6 PhOH (0.038). This charge evaluation is consistent with the trend of the observed
7
8
9 frequency shifts of the water free OH bands.
10
11
C. H+PhOH-
PhOH-W2
12
13
14 The observed IR spectrum and spectral simulations of H+PhOH-W2 are shown in Fig. 3.
15
16
17 The very intense and broadened H-bonded OH band seems to disappear from the
18
19
20 observed region at this cluster size. Instead, a moderately intense H-bonded OH band is
21
22
23 seen at ~3240 cm-1. This band indicates the presence of a H-bond between neutral
24
25
26 molecules. Three free OH stretch bands also appear in the high frequency region. The
27
28
bands at 3638 and 3725 cm-1 are assigned to the ν1 and ν3 bands of a single acceptor
29
30
31 water molecule, respectively, and the band at 3685 cm-1 is from a dangling OH of an
32
33
34 acceptor-donor (two-coordinated) water molecule. These band features suggest a
35
36
37 linear-chain type H-bonded structure with a Ph-type core as shown in the inset of Fig.
38
39
40 3(c). Its simulation well reproduces the observed spectral features, indicating that this
41
42
43 is the major isomer of H+PhOH-W2. This structure has been predicted in the previous
44
45
theoretical study by Ataelahi and Omidyan. 24 In the present simulation, an intense
46
47
48 H-bonded OH stretching band is predicted at around 2600 cm-1, but this band is missing
49
50
51 in the observed spectrum. This band is due to the strongly H-bonded phenolic OH
52
53
54
55
56
57 16
58
59
60 ACS Paragon Plus Environment
Page 17 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 stretch, and the present harmonic approximation would result in a large overestimation
7
8
9 of the vibrational frequency.
10
11
A CH---O type isomer is also stable in this size. However, its energy is much
12
13
14 higher (+25 kJ/mol) than the most stable Ph-type isomer shown in Fig. 3(c). The absence
15
16
17 of the free phenolic OH stretching band at around 3550 cm-1 supports the negligible
18
19
20 population of this type of isomer.
21
22
23 In neutral PhOH-W2, the H-bonded network forms a cyclic structure while the
24
25
26 chain type structure is preferred in PhOH+-W2. 48-51,53,-55, 56, 57, 61, 62, 65 This structural
27
28
difference can be attributed to the enhanced acidity of the phenolic OH in the cation,
29
30
31 which results in a preference for the straight OH---O H-bond angle. The reduced
32
33
34 basicity of the phenolic OH in the cation also prohibits it from accepting a proton to
35
36
37 form a cyclic structure. The observed structure of H+PhOH-W2 is similar to that of
38
39
40 PhOH+-W2, and this reflects the similar charge distribution in the OH groups in PhOH+
41
42
43 and H+PhOH.
44
45
The band width of the ν1 and ν3 bands of the observed spectrum of H+PhOH-W2
46
47
48 seems larger than that of the dangling OH band. This suggests overlap of bands due to
49
50
51 coexistence of a minor isomer. The structure and simulated IR spectrum of the most
52
53
54 stable isomer with the O-type protonated core are shown in Fig. 3(d). In this structure,
55
56
57 17
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 18 of 55

1
2
3
4
5
6 both the water molecules are directly bound to the protonated site (OH2+ group), and
7
8
9 their ν1 and ν3 bands are shifted to lower frequency by ~20 cm-1 in comparison with
10
11
those of the Ph-type isomer. The overlap of these bands well explains the larger
12
13
14 bandwidth of the observed spectrum. The relative population ratio between the
15
16
17 Ph-type and O-type isomers is roughly estimated to be ~7:1 based on the band fitting
18
19
20 shown in Fig. 3(b) and the computational transition intensities. Though the energy
21
22
23 difference between the Ph-type and O-type isomers becomes smaller with the progress
24
25
26 of the hydration, they are still separated by the high isomerization barrier. The smaller
27
28
relative population of the O-type isomer at this size may reflect the small population of
29
30
31 its precursor, H+PhOH-W0-1.
32
33
34 D. H+PhOH-
PhOH-W3
35
36
37 Figure 4(a) shows the observed IR spectrum of H+PhOH-W3 measured by monitoring
38
39
40 the water-loss channel. No signal was detected by monitoring the PhOH-loss channel.
41
42
43 H-bonded OH stretch bands are seen at ~3200 and ~3400 cm-1, and two free OH
44
45
stretching bands assigned to the ν1 and ν3 bands of single acceptor water appear at 3646
46
47
48 and 3733 cm-1, respectively. On the other hand, the dangling OH band, which is
49
50
51 intense in the spectrum of H+PhOH-W2, is remarkably suppressed. A weak band at
52
53
54 ~3800 cm-1 is attributed to a combination band of free OH stretch and intermolecular
55
56
57 18
58
59
60 ACS Paragon Plus Environment
Page 19 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 vibration. These features clearly indicate that the H-bonded water network is no longer
7
8
9 a simple linear chain type, and terminal water molecules in the network should be
10
11
single acceptors. Multiple interpretations are possible for the observed spectrum, and
12
13
14 we examine each interpretation in the following discussion.
15
16
17 The Ph-type structure shown in Fig. 4(c) is reasonably derived from the major
18
19
20 isomer of H+PhOH-W2. This structure satisfies the above-mentioned requirements
21
22
23 and its IR simulation qualitatively reproduces the observed spectral pattern. The
24
25
26 observed H-bonded OH stretch bands at ~3200 and ~3400 cm-1 are attributed to the
27
28
symmetric and antisymmetric OH stretches, respectively, of the water molecule directly
29
30
31 bound to the phenolic OH group. The symmetric OH stretch of such a water molecule
32
33
34 bound to the charged site frequently shows a larger intensity enhancement than the
35
36
37 antisymmetric OH stretch. 74 This trend agrees well with the observed intensity
38
39
40 distribution. However, the simulation fails to reproduce the observed intensity
41
42
43 distribution. The structure of this Ph-type isomer is similar to that of PhOH+-W3 but
44
45
largely different from neutral PhOH-W3 which has a cyclic structure. 48,49,51,53,55,61,62,64
46
47
48 We searched stable isomers of H+PhOH-W3, and found that the proton transferred
49
50
51 structure shown in Fig.4(d) is the most stable isomer at the ωB97X-D/6-311++G(d, p)
52
53
54 level. This isomer is derived from an initial O-type isomer with the proton transfer
55
56
57 19
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 20 of 55

1
2
3
4
5
6 from the PhOH moiety to the water moiety. The previous theoretical study on
7
8
9 H+PhOH-Wn by Ataelahi and Omidyan lacked the proton-transferred structure at this
10
11
size since their main focus was on the excited electronic state dynamics and their initial
12
13
14 structures in the energy optimization were limited to those with the Cs symmetry. 24
15
16
17 Though this isomer is more stable than the Ph-type isomer (Fig. 4(c), +13.9 kJ/mol), the
18
19
20 band fitting shown in Fig. 4(b) demonstrates that its contribution might be minor. The
21
22
23 isomer shown in Fig. 4(e) is another proton-transferred isomer also derived from the
24
25
26 O-type isomer. Though this isomer is higher in energy (+8.6 kJ/mol), its band at 3400
27
28
cm-1 (the water OH stretch bound to the phenyl oxygen) agrees well with the observed
29
30
31 band, and its partial contribution seems plausible.
32
33
34 In the previous study on the protonated benzene water clusters, Duncan and
35
36
37 coworkers demonstrated that DFT methods (including dispersion-corrected ones) have
38
39
40 trouble in evaluating the relative energy between the proton-transferred and
41
42
43 non-proton transferred species. 13 Because of the similarity between the systems, we
44
45
cannot exclude the possibility of large errors in the present energy evaluation. Moreover,
46
47
48 in H+PhOH-Wn, the isomer population does not necessarily reflect the relative energy
49
50
51 because of the potential barrier, as clearly demonstrated in the monomer. Duncan and
52
53
54 coworkers showed that comparison with the spectra of protonated water is helpful to
55
56
57 20
58
59
60 ACS Paragon Plus Environment
Page 21 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 identify the proton-transferred structure when the energy evaluation is useless. 13, 75

7
8
9 Thus, the IR spectrum of H+(H2O)3-benzene is compared with the present IR spectrum
10
11
of H+PhOH-W3. Fig. 4(f) is the reproduction of the IR spectrum of H+(H2O)3-benzene
12
13
14 taken from ref. 13. In this cluster, the excess proton is in the water moiety, and the
15
16
17 resulting hydronium ion (H3O+) is π-H-bonded to the phenyl ring. The π-H-bonded OH
18
19
20 stretch band of the hydronium ion appears at 3192 cm-1. The spectrum of
21
22
23 H+(H2O)3-benzene is quite similar to that of H+PhOH-W3, and this suggests the proton
24
25
26 transfer from phenol to the water moiety occurs at this size. However, the 3400 cm-1
27
28
band of H+PhOH-W3 is absent in the spectrum of H+(H2O)3-benzene. If this band is
29
30
31 attributed to the water OH stretch bound to the phenyl oxygen, as seen in the isomer of
32
33
34 Fig. 4(e), one of the neutral water molecules becomes a donor-acceptor (two-coordinated)
35
36
37 site, and its free OH should show the dangling OH band at ~3700 cm-1. However, as
38
39
40 seen in the band fitting in Fig. 4(b), such a band should be very weak. Therefore, any
41
42
43 single isomer cannot explain all the spectral features. Since the Ph- and O-types
44
45
coexist in H+PhOH-Wn<3, which might be precursors in the H+PhOH-W3 production, the
46
47
48 coexistence of both the non-proton-transferred isomers (the Ph-type) and
49
50
51 proton-transferred isomer (from the O-type) also seems highly plausible in H+PhOH-W3.
52
53
54
55
56
57 21
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 22 of 55

1
2
3
4
5
6 At the present stage, unequivocal determination of the isomer structure of H+PhOH-W3
7
8
9 is difficult.
10
11
E. H+PhOH-
PhOH-W4
12
13
14 The observed IR spectrum of H+PhOH-W4 is shown in Fig. 5(a). The number of band
15
16
17 features suddenly increases, and the spectrum is much more complicated. This suggests
18
19
20 a large change in the H-bonded network structure (of the major isomer) at this size. A
21
22
23 characteristic feature of the spectrum is the band at 3540 cm-1. This is the region where
24
25
26 the free phenolic OH stretch of H+PhOH (Ph-type) is expected. 19-21 However, the
27
28
phenolic OH is preferentially H-bonded by water, and the free phenolic OH band is not
29
30
31 actually seen in the spectra of H+PhOH-Wn (n = 1- 3). Therefore, this band is
32
33
34 attributed to a weakly H-bonded OH stretch of the neutral water moiety. 76 The most
35
36
37 stable isomer we found at this size is shown in Fig. 5(b). This isomer has a proton
38
39
40 transferred structure with the H3O+ ion core, and a neutral water moiety is bound to the
41
42
43 phenyl ring by a π-H-bond. Its simulated spectrum well reproduces the observed
44
45
spectral features above 3400 cm-1, and the observed band at 3540 cm-1 is assigned to the
46
47
48 π-H-bonded OH stretch (here, we should note that if the H3O+ moiety is directly bound
49
50
51 to the phenyl ring, its π-H-bonded OH is shifted to much lower frequency (~3200 cm-1),
52
53
54 13 as seen in the spectrum of H+(H2O)3-benzene in Fig. 4(f)). However, the agreement in
55
56
57 22
58
59
60 ACS Paragon Plus Environment
Page 23 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 the low frequency region is only qualitative. This can be partly attributed to the large
7
8
9 influence of anharmonicity and significant line broadening expected for the H-bonded
10
11
OH bands lower than ~3000 cm-1. However, contributions of other isomers are also
12
13
14 suggested. Some low energy isomers that also have the protonated water core (H3O+)
15
16
17 and π-H-bond are found (Figs. 5(c)-(d)), and the observed features might be partially
18
19
20 attributed to these coexisting isomers.
21
22
23 In the n = 1 – 3 clusters, preferential hydration of the phenolic OH group is
24
25
26 seen, and the water moiety is rather isolated from the protonated CH2 site in the major
27
28
Ph-type. Therefore, as discussed for n = 3, the proton transfer might occur only in the
29
30
31 minor O-type. At the present size (n = 4), the water network can be long enough to link
32
33
34 the phenyl ring to the phenolic OH group. Therefore, proton transfer can occur from
35
36
37 both the Ph- and O-types, and this results in the dominance of the proton transferred
38
39
40 structures. When the water moiety is still separated from the CH2 site, the proton
41
42
43 transfer from the phenolic OH can occur but the resulting structure is estimated to be
44
45
much higher in energy (+32.9 kJ/mol) as shown in Fig. 5(e).
46
47
48 It is very interesting that the protonated water moiety prefers to be bound to
49
50
51 the phenyl ring through the π-H-bond. In PhOH+-Wn, though the proton transfer from
52
53
54 PhOH+ to the water moiety also occurs at n = 3 or 4, the water H-bond network is
55
56
57 23
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 24 of 55

1
2
3
4
5
6 extended in front of the phenoxy oxygen atom, and the π-H-bond formation has never
7
8
9 been observed. 60-62, 64, 66 Once the proton is transferred to the water moiety, neutral
10
11
phenol is formed in H+PhOH-Wn, while phenoxy radical is formed in PhOH+-Wn. The
12
13
14 hydroxy oxygen in neutral phenol has sp3 hybridization, and the H-bond to the phenolic
15
16
17 oxygen directs the water moiety out of the plane of the phenyl ring (see Figs. 5(b)-(d)).
18
19
20 This location of the water moiety gives it good accessibility to the phenyl ring plane to
21
22
23 form the π-H-bond. On the other hand, the oxygen atom in the phenoxy radical can be
24
25
26 regarded as that of a carbonyl group and has sp2 hybridization. Therefore, the water
27
28
moiety H-bonded to the oxygen atom resides in the plane of the phenyl ring (see Fig.
29
30
31 5(e) for a similar case of a H-bond to a carbonyl oxygen). In this case, the accessibility
32
33
34 to the phenyl ring plane is much worse. This may prohibit the formation of a
35
36
37 π-H-bonded structure in PhOH+-Wn.
38
39
40 F. H+PhOH-
PhOH-W5
41
42
43 Figure 6 shows (a) observed IR spectrum and (b)-(d) low energy stable isomer structures
44
45
of H+PhOH-W5 and their simulated IR spectra. Since the parent ion intensity of this
46
47
48 cluster size is quite weak, the quality of the observed spectrum is worse than the other
49
50
51 sizes. The spectral features are similar to that of H+PhOH-W4. The π-H-bonded OH
52
53
54 stretch band is still seen at around 3550 cm-1, and a broad absorption of H-bonded OH
55
56
57 24
58
59
60 ACS Paragon Plus Environment
Page 25 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 stretches appears below 3450 cm-1. In the free OH stretch region, the dangling OH
7
8
9 stretch band of an acceptor-donor (two-coordinated) water site is clearly seen at 3708
10
11
12 cm-1. The ν1 band of single acceptor water also appears at 3641 cm-1, and the ν3 band
13
14 becomes a shoulder of the dangling OH band.
15
16
17 The calculated low energy stable isomers have the proton transferred
18
19
20 structures with the H3O+ ion core, and the water network is located on the phenyl ring.
21
22
23 The observed spectral features, especially the π-H-bonded OH band at ~3550 cm-1, are
24
25
26 consistent with these structures of the low energy isomers. Similar to the case of the n =
27
28
4 cluster, such structures are strikingly different from those of PhOH+-W5, in which the
29
30
31 proton is transferred to the water moiety but the water network develops away from the
32
33
34 phenyl ring. 66 The difference in the H-bond coordination property of the O atom in the
35
36
37 aromatic moiety would be the origin of the structure difference. In H+PhOH-W5, the
38
39
40 energy separation among the isomers is very small, and the simulated IR spectra of the
41
42
43 isomers are more or less similar to each other. Therefore, definitive determination of
44
45
the structure of the observed cluster is difficult.
46
47
48 G. Ar-
Ar-tagged clusters of H+PhOH-
PhOH-Wn
49
50
51 The IR spectra of Ar-tagged clusters of H+PhOH-Wn (n = 1 -4) were also measured by
52
53
54 monitoring the Ar-loss channel. The observed spectra are summarized in Figs. S2-S5
55
56
57 25
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 26 of 55

1
2
3
4
5
6 with the comparison of the observed spectra of the corresponding bare clusters and
7
8
9 simulations. Ar-tagging is a highly useful technique to select low energy isomers and
10
11
simplify spectra of ionic clusters. 67-69 However, in the present system, the low energy
12
13
14 isomer selection by the Ar-tagging is not effective for the multiple isomers separated by
15
16
17 the high energy barrier. We found that in the size range of n = 1 - 4, the observed band
18
19
20 features of the Ar-tagged clusters were essentially the same as those of the bare clusters,
21
22
23 and no clear difference in the isomer distribution occurred by the tagging. In addition,
24
25
26 so many possible Ar-binding sites exist in these clusters, and the coexistence of multiple
27
28
isomers (due to both the bare cluster structures and the Ar-binding sites) makes the
29
30
31 observed spectra even more complex, especially in the free OH stretch region.
32
33
34 Therefore, essentially no new information was obtained from the measurement of the
35
36
37 Ar-tagged spectra.
38
39
40 4. Concluding remarks
remarks
41
42
43 The hydrogen bond structures of H+PhOH-Wn (n = 1 - 5) were studied by IR
44
45
spectroscopy in the OH and CH stretch region. The protonation of phenol occurs on the
46
47
48 phenyl ring (Ph-type) and hydroxy group (O-type), as previously reported. 19-21 The
49
50
51 coexistence of both isomer types even under the high pressure supersonic jet expansion
52
53
54 condition was reconfirmed by IR spectroscopy of H+PhOH-Ar. In both the isomer types,
55
56
57 26
58
59
60 ACS Paragon Plus Environment
Page 27 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 the initial hydration preferentially occurs at the hydroxy group, and the water cluster
7
8
9 develops exclusively around the hydroxy group up to n = 2. These observed hydrated
10
11
structures are similar to those of the PhOH+-Wn radical cation clusters of the
12
13
14 corresponding size. The proton transferred structure with the H3O+ ion core was
15
16
17 estimated to be the most stable isomer at n = 3, but the coexistence of the
18
19
20 proton-transferred and non-proton-transferred isomers was suggested at this size.
21
22
23 The former comes from the O-type while the latter corresponds to the Ph-type. The
24
25
26 geometrical separation of the protonated CH2 site and the water moiety around the
27
28
hydroxy group prevents the proton transfer in the Ph-type isomer at this size. At n ≥ 4,
29
30
31 the proton-transferred structures become dominant. The water moiety is bound to the
32
33
34 phenyl ring by a π-H-bond, and resides in the ring plane. These structures are in
35
36
37 striking contrast to those of PhOH+-Wn (n ≥ 4), in which the proton transferred water
38
39
40 moiety is located away from the phenyl ring plane. This structural difference is
41
42
43 interpreted in terms of the difference of the hydrogen bond coordination properties of
44
45
the O atom in neutral phenol and phenoxy radical.
46
47
48 Because of the presence of the high energy barrier among multiple isomers, the
49
50
51 minimum energy isomer is not necessarily the dominantly observed isomer in
52
53
54 H+PhOH-Wn. In addition, the problem of the accuracy of DFT computations on a similar
55
56
57 27
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 28 of 55

1
2
3
4
5
6 protonated system has been pointed out. 13 Therefore, spectroscopic observations are
7
8
9 highly important to elucidate the actual microhydration process in H+PhOH-Wn.
10
11
Though we have focused on the OH stretching vibrational region in the present study,
12
13
14 observation of the fingerprint region is promising to obtain more structural information.
15
16
17 It has been reported that the protonation of benzene causes strong IR activity in carbon
18
19
20 ring distortion modes and CH bends.7, 11 Similar behaviors are expected for phenol,
21
22
23 and observation of the fingerprint region will be an important future task for further
24
25
26 unequivocal isomer structure determination. Strongly H-bonded OH stretches of the
27
28
charged site are also expected to appear below 2500 cm-1, and their observation is highly
29
30
31 important to unequivocally identify proton transfer in the clusters.
32
33
34
35
36
37 Supporting Information
38
39
40 Definitions of numbering of atoms in protonated phenol, phenol cation, and neutral
41
42
43 phenol. Charge distribution of protonated phenol, phenol cation, and neutral phenol.
44
45
Infrared spectra of protonated phenol-(water)n-Ar clusters (n = 1 -4). Complete author
46
47
48 list of ref.73.
49
50
51
52
53
54 Acknowledgements
55
56
57 28
58
59
60 ACS Paragon Plus Environment
Page 29 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 We would like to acknowledge Dr. Toshihiko Maeyama and Dr. Yoshiyuki Matsuda for
7
8
9 their helpful discussions. This study was partly supported by a Grant-in-Aid for
10
11
Scientific Research (Project No. 18H01931) from JSPS.
12
13
14
15
16
17 Notes
18
19
20 The authors declare no competing financial interest.
21
22
23
24
25
26 References
27
28
1. Carey, F. A.; Sunderg, R. J. Advanced Organic Chemistry; Plenum: New York, 1993.
29
30
31 2. A. Pathak, A.; and P. J. Sarre, P. J. Protonated PAHs as Carriers of Diffuse
32
33
34 Interstellar Bands. Mon. Not. R. Astron. Soc. 2008,
2008 391, L10–L14.
35
36
37 3. Dopfer, O. Laboratory Spectroscopy of Protonated PAH molecules Relevant for
38
39
40 Interstellar Chemistry. EAS Publications Series 2011,
2011 46, 103-108.
41
42
43 4. Berg, J. M.; Tymoczko, J. T.; Stryer, L. Biochemistry, 5th Ed.; W. H. Freeman: New
44
45
York, 2002.
46
47
48 5. Dopfer, O. IR Spectroscopic Strategies for the Structural Characterization of Isolated
49
50
51 and Microsolvated Arenium Ions. J. Phys. Org. Chem. 2006,
2006 19, 540-551.
52
53
54
55
56
57 29
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 30 of 55

1
2
3
4
5
6 6. Maksić, Z. B.; Kovačević, B.; Vianello, R. Advances in Determining the Absolute
7
8
9 Proton Affinities of Neutral Organic Molecules in the Gas Phase and Their
10
11
Interpretation: A Theoretical Account. Chem. Rev. 2012,
2012 112, 5240–5270.
12
13
14 7. Duncan, M. A. Infrared Laser Spectroscopy of Mass-Selected Carbocations. J. Phys.
15
16
17 Chem. A, 2012,
2012 116, 11477–11491.
18
19
20 8. Freiser, B. S.; Beauchamp, J. L. Acid-Base Properties of Molecules in Excited
21
22
23 Electronic States Utilizing Ion Cyclotron Resonance Spectroscopy. J. Am. Chem. Soc.,
24
25
26 1977,
1977 99, 3214-3225.
27
28
9. Solcà, N.; Dopfer, O. Protonated Benzene: IR Spectrum and Structure of C6H7+.
29
30
31 Angew. Chem. Int. Ed. 2002,
2002 41, 3628-3631.
32
33
34 10. Solcà, N.; Dopfer, O. Interaction of the Benzenium Ion with Inert Ligands: IR
35
36
37 spectra of C6H7+-Ln Cluster Cations (L = Ar, N2, CH4, H2O). Chemistry - A Eur. J. 2003,
2003 9,
38
39
40 3154-3163.
41
42
43 11. Douberly, G. E.; Ricks, A. M.; Schleyer, P. v. R.; Duncan, M. A. Infrared Spectroscopy
44
45
of Gas Phase Benzenium Ions: Protonated Benzene and Protonated Toluene, from 750 to
46
47
48 3400 cm-1. J. Phys. Chem. A, 2008,
2008 112, 4869–4874.
49
50
51 12. Bandyopadhyay, B.; Cheng, T. C.; Wheeler, S. E.; Duncan, M. A. Vibrational
52
53
54 Spectroscopy and Theory of the Protonated Benzene Dimer and Trimer. J. Phys. Chem.
55
56
57 30
58
59
60 ACS Paragon Plus Environment
Page 31 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 A, 2012,
2012 116, 7065–7073.
7
8
9 13. Cheng, T. C.; Bandyopadhyay, B.; Mosley, J. D.; Duncan, M. A. IR Spectroscopy of
10
11
Protonation in Benzene–Water Nanoclusters: Hydronium, Zundel, and Eigen at a
12
13
14 Hydrophobic Interface. J. Am. Chem. Soc., 2012,
2012 134, 13046–13055.
15
16
17 14. Rode, M. F.; Sobolewski, A. L.; Dedonder, C.; Jouvet, C.; Dopfer, O. Computational
18
19
20 Study on the Photophysics of Protonated Benzene. J. Phys. Chem. A, 2009,
2009 113, 5865–
21
22
23 5873.
24
25
26 15. Chakraborty, S.; Omidyan, R.; Alata, I.; Nielsen, I. B.; Dedonder, C.; Broquier, M.;
27
28
Jouvet, C. Protonated Benzene Dimer: An Experimental and Ab Initio Study. J. Am.
29
30
31 Chem. Soc., 2009,
2009 131, 11091–11097.
32
33
34 16. Esteves-López,N.; Dedonder-Lardeux, C.; Jouvet, C. Excited State of Protonated
35
36
37 Benzene and Toluene. J. Chem. Phys. 2015,
2015 143, 074303.
38
39
40 17. Alata, I.; Omidyan, R.; Broquier, M.; Dedonder, C.; Dopfer, O.; Jouvet, C. Effect of
41
42
43 Protonation on the Electronic Structure of Aromatic Molecules: NaphthaleneH+. Phys.
44
45
Chem. Chem. Phys. 2010,
2010 12, 14456-14458.
46
47
48 18. Patzer, A.; Schütz, M.; Jouvet, C.; Dopfer, O. Experimental Observation and
49
50
51 Quantum Chemical Characterization of the S1 ← S0 Transition of Protonated
52
53
54 Naphthalene-Argon Clusters. J. Phys. Chem. A, 2013,
2013 117, 9785–9793.
55
56
57 31
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 32 of 55

1
2
3
4
5
6 19. Solcà N.; Dopfer, O. Protonation of Aromatic Molecules: Competition between Ring
7
8
9 and Oxygen Protonation of Phenol (Ph) Revealed by IR Spectra of PhH+-Arn. Chem.
10
11
12
Phys. Lett. 2001,
2001 342, 191-199.
13
14 20. Solcà N.; Dopfer, O. Spectroscopic Identification of Oxonium and Carbenium Ions of
15
16
17 Protonated Phenol in the Gas Phase: IR Spectra of Weakly Bound C6H7O+-L Dimers (L
18
19
20 = Ne, Ar, N2). J. Am. Chem. Soc., 2004,
2004 126, 1716–1725.
21
22
23 21. Solcà N.; Dopfer, O. Isomer-Selective Detection of Microsolvated Oxonium and
24
25
26 Carbenium Ions of Protonated Phenol: Infrared Spectra of C6H7O+-Ln Clusters
27
28
(L=Ar/N2, n≤6). J. Chem. Phys. 2004,
2004 120, 10470.
29
30
31 22. Solcà N.; Dopfer, O. Selective Infrared Photodissociation of Protonated
32
33
34 para-Fluorophenol Isomers: Substitution Effects in Oxonium and Fluoronium Ions. J.
35
36
37 Chem. Phys. 2004,
2004 121, 769.
38
39
40 23. Azizkarimi, S.; Omidyan, R.; Azimi, G. Electronically Excited States of Protonated
41
42
43 Phenol and para-Substituted Phenol. Chem. Phys. Lett. 2013,
2013 555, 19-25.
44
45
24. Ataelahi, M.; Omidyan, R. Microhydration Effects on the Electronic Properties of
46
47
48 Protonated Phenol: A Theoretical Study. J. Phys. Chem. A, 2013,
2013 117, 12842–12850.
49
50
51 25. Škríba, A.; Janková, Š.; Váňa, J.; Barták, P.; Bednář, P.; Fryčák, P.: Kučera, L.:
52
53
54 Kurka, O.; Lemr, K.; Macíková, P.; Marková, E.; Nováková, P.: Papoušková, B.;
55
56
57 32
58
59
60 ACS Paragon Plus Environment
Page 33 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 Skopalová, J.; Švecová, H.; Roithová, J. Protonation Sites and Fragmentations of
7
8
9 para-Aminophenol. Int. J. Mass Spectrom. 2013,
2013 337, 18-23.
10
11
26. Haupert L. J.; Wenthold, P. G. Hydration Energies of Aromatic Ions in the Gas
12
13
14 Phase. J. Phys. Chem. A, 2013,
2013 117, 1164–1170.
15
16
17 27. Chiavarino, B.; Crestoni, M. E.; Fornarini, S. IR Spectroscopic Features of Gaseous
18
19
20 C7H7O+ Ions: Benzylium versus Tropylium Ion Structures. J. Phys. Chem. A, 2006,
2006 110,
21
22
23 9352–9360.
24
25
26 28. Alata, I.; Omidyan, R.; Dedonder-Lardeux, C.; Broquierac, M.; Jouvet, C.
27
28
Electronically Excited States of Protonated Aromatic Molecules: Benzaldehyde. Phys.
29
30
31 Chem. Chem. Phys. 2009,
2009 11, 11479-11486.
32
33
34 29. Chakraborty, S.; Patzer, A.; Dopfer. O. IR Spectra of Protonated Benzaldehyde
35
36
37 Clusters, C7H7O+-Ln (L=Ar, N2; N≤2): Ion-Ligand Binding Motifs of the Cis and Trans
38
39
40 Oxonium Isomers. J. Chem. Phys. 2010,
2010 133, 044307.
41
42
43 30. Patzer, A.; Zimmermann, M.; Alata, I.; Jouvet, C.; Dopfer, O. Electronic Spectra of
44
45
Protonated Benzaldehyde Clusters with Ar and N2: Effect of ππ* Excitation on the
46
47
48 Intermolecular Potential, J. Phys. Chem. A 2010,
2010 114, 12600-12604.
49
50
51 31. Dopfer, O.; Patzer, A,; Chakraborty, S,; Alata, I.; Omidyan, R.; Broquier, M.;
52
53
54 Dedonder, C.; Jouvet, C. Electronic and Vibrational Spectra of Protonated
55
56
57 33
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 34 of 55

1
2
3
4
5
6 Benzaldehyde-Water Clusters, [BZ-(H2O)n≤5]H+: Evidence for Ground-State Proton
7
8
9 Transfer to Solvent for n ≥ 3. J. Chem. Phys. 2014,
2014 140, 124314.
10
11
32. Pasker, F. M.; Solcà N.; Dopfer, O. Spectroscopic Identification of Carbenium and
12
13
14 Ammonium Isomers of Protonated Aniline (AnH+): IR Spectra of Weakly Bound
15
16
17 AnH+-Ln Clusters (L = Ar, N2). J. Phys. Chem. A 2006,
2006 110, 12793-12804.
18
19
20 33. Zhao, D.; Langer, J.; Oomens, J.; Dopfer, O. Infrared Spectra of Protonated
21
22
23 Polycyclic Aromatic Hydrocarbon Molecules: Azulene. J. Chem. Phys. 2009,
2009 131,
24
25
26 184307.
27
28
34. Lagutschenkov L.; Dopfer, O. Infrared Spectrum of a Protonated Fluorescence Dye:
29
30
31 Acridine Orange. J. Mol. Spectrosc. 2011,
2011 268, 66-77.
32
33
34 35. Alata, I.; Brouquier, M.; Dedonder, C.; Jouver, C.; Marceca, E. Electronic Excited
35
36
37 States of Protonated Aromatic Molecules: Protonated Fluorene. Chem. Phys. 2012,
2012 393,
38
39
40 25-31.
41
42
43 36. Alata, I.; Omidyan, R.; Broquier, M.; Dedonder, C.; Jouvet, C. Protonated
44
45
Salicylaldehyde: Electronic Properties. Chem. Phys. 2012,
2012 399, 224-231.
46
47
48 37. Féraud, G.; Domenianni, L.; Marceca, E.; Dedonder-Lardeux, C.; Jouvet, C.
49
50
51 Photodissociation Electronic Spectra of Cold Protonated Quinoline and Isoquinoline in
52
53
54 the Gas Phase. J. Phys. Chem. A, 2017,
2017 121, 2580–2587.
55
56
57 34
58
59
60 ACS Paragon Plus Environment
Page 35 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 38. Omidyan, R. Protonation Effect on the Electronic Structure of Small PAHs:
7
8
9 Acenaphthylene and Acenaphthene. Chem. Phys. Lett. 2011,
2011 518, 15-20.
10
11
39. Omidyan, R.; Heidari, Z.; Salehi, M.; Azimi, G. Electronically Excited States of
12
13
14 Neutral, Protonated α-Naphthol and Their Water Clusters: A Theoretical Study. J. Phys.
15
16
17 Chem. A, 2015,
2015 119, 6650–6660.
18
19
20 40. Omidyan, R.; Omidyana M.; Mohammadzadeh, A. Electronically Excited State of
21
22
23 Neutral/Protonated, Indole/5-Hydroxyinodole-Water Clusters: A Theoretical Study. RSC
24
25
26 Advances, 2016,
2016 6, 33148-33158.
27
28
41. Ishiuchi, S. –I.; Wako, H.; Kato, D.; Fujii, M. High-Cooling-Efficiency Cryogenic
29
30
31 Quadrupole Ion Trap and UV-UV Hole Burning Spectroscopy of Protonated Tyrosine. J.
32
33
34 Mol. Spectrosc. 2017,
2017 332, 45-51.
35
36
37 42. Wako, H.; Ishiuchi, S. –I.; Kato, D.; Féraud, G.; Dedonder-Lardeux, C.; Jouvetb C.;
38
39
40 Fujii, M. A Conformational Study of Protonated Noradrenaline by UV–UV and IR Dip
41
42
43 Double Resonance Laser Spectroscopy Combined with an Electrospray and a Cold Ion
44
45
Trap Method. Phys. Chem. Chem. Phys. 2017,
2017 19, 10777-10785.
46
47
48 43. Marian, C.; Nolting, D.; Weinkauf, R. The Electronic Spectrum of Protonated
49
50
51 Adenine: Theory and Experiment. Phys. Chem. Chem. Phys. 2005,
2005 7, 3306-3316.
52
53
54 44. Hansen, C. S.; Kirk, B. B.; Blanksby, S. J.; Trevitt, A. J. Ultraviolet
55
56
57 35
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 36 of 55

1
2
3
4
5
6 Photodissociation of the N-Methylpyridinium Ion: Action Spectroscopy and Product
7
8
9 Characterization. J. Phys. Chem. A 2013,
2013 117, 10839-10846.
10
11
45. Valencia, D.; Whiting, G. T.; Buloa, R. E.; Weckhuysen, B. M. Protonated
12
13
14 Thiophene-Based Oligomers as Formed within Zeolites: Understanding Their Electron
15
16
17 Delocalization and Aromaticity. Phys. Chem. Chem. Phys. 2016,
2016 18, 2080-2086.
18
19
20 46. Hunter, E. P.; Lias, S. G. Evaluated Gas Phase Basicities and Proton Affinities of
21
22
23 Molecules: An Update. J. Phys. Chem. Ref. Data 1998,
1998 27, 413−457.
24
25
26 47. Tanabe, S.; Ebata, T.; Fujii, M.; Mikami, N. OH Stretching Vibrations of
27
28
Phenol-(H2O)n (n=1-3) Complexes Observed by IR-UV Double Resonance Spectroscopy.
29
30
31 Chem. Phys. Lett. 1993,
1993 215, 347-352.
32
33
34 48. Watanabe, T.; Ebata, T.; Tanabe, S,; Mikami, N. Size-Selected Vibrational Spectra of
35
36
37 Phenol ‐ (H2O)n (n=1 – 4) Clusters Observed by IR – UV Double Resonance and
38
39
40 Stimulated Raman‐UV Double Resonance Spectroscopies. J. Chem. Phys. 1996,
1996 105,
41
42
43 408-419.
44
45 49. Watanabe H.; Iwata, S. Theoretical Studies of Geometric Structures of Phenol‐
46
47
48 Water Clusters and Their Infrared Absorption Spectra in the O–H Stretching Region. J.
49
50
51 Chem. Phys. 1996,
1996 105, 420-431.
52
53
54 50. Gerhards, M.; Kleinermanns, K. Structure and Vibrations of Phenol(H2O)2. J. Chem.
55
56
57 36
58
59
60 ACS Paragon Plus Environment
Page 37 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 Phys. 1995,
1995 103, 7392-7400.
7
8
9 51. Roth, W.; Schmitt, M.; Jacoby, Ch.; Spangenberg, D.; Janzen, Ch.; Kleinermanns, K.
10
11
Double Resonance Spectroscopy of Phenol(H2O)1-12: Evidence for Ice-Like Structures in
12
13
14 Aromate-Water clusters? Chem. Phys. 1998,
1998 239, 1-9.
15
16
17 52. Janzen, Ch.; Spangenberg, D.; Roth, W.; Kleinermanns, K. Structure and Vibrations
18
19
20 of Phenol(H2O)7,8 Studied by Infrared-Ultraviolet and Ultraviolet-Ultraviolet
21
22
23 Double-Resonance Spectroscopy and Ab Initio Theory. J. Chem. Phys. 1999,
1999 110,
24
25
26 9898-9907.
27
28
53. Lüchow, A.; Spangenberg, D.; Janzen, Ch.; Jansen, A.; Gerhards, M.; Kleinermanns,
29
30
31 K. Structure and Energetics of Phenol(H2O) ,n ≤ 7: Quantum Monte Carlo Calculations
32
33
34 and Double Resonance Experiments. Phys. Chem. Chem. Phys. 2001,
2001 3, 2771-2780.
35
36
37 54. Fang, W.-H.; Liu, R. -Z. Theoretical Characterization of the Structures and
38
39
40 Properties of Phenol-(H2O)2 Complexes. J. Chem. Phys. 2000,
2000 113, 5253−5258.
41
42
43 55. Kryachko, E.; Nakatsuji, H. Ab Initio Study of Lower Energy Phenol-Water 1≤n≤4
44
45
Complexes: Interpretation of Two Distinct Infrared Patterns in Spectra of Phenol-Water
46
47
48 Tetramer. J. Phys. Chem. A 2002,
2002 106, 731−742.
49
50
51 56. Shimamori T.; Fujii, A., Infrared Spectroscopy of Warm and Neutral Phenol-Water
52
53
54 Clusters. J. Phys. Chem. A 2015,
2015 119, 1315–1322.
55
56
57 37
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 38 of 55

1
2
3
4
5
6 57. Samala N. R.; Agmon, N. Structure, Spectroscopy, and Dynamics of the
7
8
9 Phenol-(Water)2 Cluster at Low and High Temperatures. J. Chem. Phys. 2017,
2017 147,
10
11
234307.
12
13
14 58. Mizuse, K.; Hamashima T.; Fujii, A. Infrared Spectroscopy of Phenol-(H2O)n>10:
15
16
17 Structural Strains in Hydrogen Bond Networks of Neutral Water Clusters. J. Phys.
18
19
20 Chem. A, 2009,
2009 113, 12134–12141.
21
22
23 59. Hamashima, T.; Mizuse, K.; Fujii. A. Spectral Signatures of Four-Coordinated Sites
24
25
26 in Water Clusters: Infrared Spectroscopy of Phenol−(H2O)n (∼20 ≤ n ≤ ∼50). J. Phys.
27
28
Chem. A, 2011,
2011 115, 620–625.
29
30
31 60. Sato, S.; Mikami, N. Size Dependence of Intracluster Proton Transfer of
32
33
34 Phenol−(H2O)n (n = 1−4) Cations. J. Phys. Chem. 1996,
1996 100, 4765–4769.
35
36
37 61. Sawamura, T.; Fujii, A.; Sato, S.; Ebata, T.; Mikami., N. Characterization of the
38
39
40 Hydrogen-Bonded Cluster Ions [Phenol−(H2O)n]+ (n = 1−4), (Phenol)2+, and
41
42
43 (Phenol−Methanol)+ As Studied by Trapped Ion Infrared Multiphoton Dissociation
44
45
Spectroscopy of Their OH Stretching Vibrations. J. Phys. Chem. 1996,
1996 100, 8131–8138.
46
47
48 62. Re, S.; Osamura, Y. Size-Dependent Hydrogen Bonds of Cluster Ions between
49
50
51 Phenol Cation Radicals and Water Molecules: A Molecular Orbital Study. J. Phys. Chem.
52
53
54 A, 1998,
1998 102, 3798–3812.
55
56
57 38
58
59
60 ACS Paragon Plus Environment
Page 39 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 63. Chapman, D. M.; Hompf, F. J.; Müller-Dethlefs, K.; Waterstradt E.; Hobza, P.; ,
7
8
9 Špirko, V. Structure and Dynamics of the Phenol-Water-Argon Cation Radical. Chem.
10
11
12
Phys. 1998,
1998 239, 417-428.
13
14 64. Kleinermanns, K.; Janzen, Ch.; Spangenberg, D,; Gerhards, M. Infrared
15
16
17 Spectroscopy of Resonantly Ionized (Phenol)(H2O)n+. J. Phys. Chem. A, 1999,
1999 103, 5232–
18
19
20 5239.
21
22
23 65. Gerhards, M.; Jansen, A.; Unterberg, C.; Kleinermanns, K. OH Stretching
24
25
26 Vibrations of the Phenol(H2O)1+ Cation. Chem. Phys. Lett. 2001,
2001 344, 113-119.
27
28
66. Ishikawa, H.; Kurusu, I.; Yagi, R.; Kato, R.; Kasahara, Y. Quantitative Temperature
29
30
31 Dependence of the Microscopic Hydration Structures Investigated by Ultraviolet
32
33
34 Photodissociation Spectroscopy of Hydrated Phenol Cations. J. Phys. Chem. Lett. 2017,
2017
35
36
37 8, 2541–2546.
38
39
40 67. Okumura, M.; Yeh, L. I.; Myers, J. D.; Lee, Y. T. Infrared Spectra of the Solvated
41
42
43 Hydronium Ion: Vibrational Predissociation Spectroscopy of Mass-Selected
44
45
H3O+⋅(H2O)n⋅(H2)m. J. Phys. Chem. 1990,
1990 94, 3416-3427.
46
47
48 68. Li, Y. − C.; Hamashima, T.; Yamazaki, R.; Kobayashi, T.; Suzuki, Y.; Mizuse, K.; Fujii,
49
50
51 A.; Kuo, J. − L. Hydrogen-Bonded Ring Closing and Opening of Protonated Methanol
52
53
54 Clusters H+(CH3OH)n (n = 4−8) with the Inert Gas Tagging. Phys. Chem. Chem. Phys.
55
56
57 39
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 40 of 55

1
2
3
4
5
6 2015,
2015 17, 22042−22053.
7
8
9 69. Mizuse, K.; Fujii, A. Infrared Photodissociation Spectroscopy of H+ (H2O)6 —Mm (M =
10
11
Ne, Ar, Kr, Xe, H2, N2, and CH4): Messenger- Dependent Balance between H3O+ and
12
13
14 H5O2+ Core Isomers. Phys. Chem. Chem. Phys. 2011,
2011 13, 7129−7135.
15
16
17 70. Even, U.; Jortner, J.; Noy, D.; Lavie, N.; Cossart-Magos, C. Cooling of Large
18
19
20 Molecules below 1 K and He Clusters Formation. J. Chem. Phys. 2000,
2000 112, 8068−8071.
21
22
23 71. Fujii, A.; Sawamura, T.; Tanabe, S.; Ebata, T.; Mikami, N. Infrared Dissociation
24
25
26 Spectroscopy of the OH stretching Vibration of Phenol-Rare Gas van der Waals Cluster
27
28
Ion. Chem. Phys. Lett. 1994,
1994 225, 104-107.
29
30
31 72. Solcà, N.; Dopfer, O. Infrared Spectra of the Phenol–Ar and Phenol–N2 Cations:
32
33
34 Proton-Bound versus π-Bound Structures. Chem. Phys. Lett. 2000,
2000 325, 354-359.
35
36
37 73. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman,
38
39
40 J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; et al. Gaussian 16,
41
42
43 Revision A.03, Gaussian, Inc., Wallingford CT, 2016.
44
45
74. Choi, J. –H.; Kuwata, K. T.; Haas, B. –M.; Cao, Y.; Johnson, M. S. Okumura, M.
46
47
48 Vibrational Spectroscopy of NO+(H2O)n: Evidence for the Intracluster Reaction
49
50
51 NO+(H2O)n →H3O+(H2O)n-2(HONO) at n≥4. J. Chem. Phys. 1994,
1994 100, 7153-7165.
52
53
54 75. Mauney, D. T.; Maner, J. A.; Duncan, M. A. IR Spectroscopy of Protonated
55
56
57 40
58
59
60 ACS Paragon Plus Environment
Page 41 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6 Acetylacetone and Its Water Clusters: Enol-Keto Tautomers and Ion  Solvent Proton
7
8
9 Transfer. J. Phys. Chem. A 2017,
2017 121, 7059-7069.
10
11
76. Pribble, R. N.; Zwier, T. S. Size-Specific Infrared Spectra of Benzene-(H2O)n Clusters
12
13
14 (n = 1 through 7); Evidence for Noncyclic (H2O)n Structures. Science, 1994,
1994 265, 75-79.
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57 41
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 42 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43 Figure 1 (a) Observed IR spectrum of H+PhOH-Ar. (b) Lorentzian fitting of the
44
45 observed spectrum after slight smoothing. The upper trace is the residual in the
46 fitting. (c)-(e) Simulated IR spectra of the Ph-type (p-), Ph-type (o-), and O-type isomers
47
of H+PhOH-Ar, respectively. All the simulations were performed at the
48
49 ωB97X-D/6-311++G(d, p) level and a scaling factor of 0.932 was applied to the
50 vibrational frequencies. Relative energies of the isomers are shown in kJ/mol. (f) IR
51
52 spectrum of PhOH+-Ar. Reprinted with permission from ref. 72. Copyright 2000
53 Elsevier.
54
55
56
57 42
58
59
60 ACS Paragon Plus Environment
Page 43 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42 Figure 2 (a) Observed IR spectrum of H+PhOH-W1. (b) Lorentzian fitting of the
43 observed spectrum after slight smoothing. The upper trace is the residual in the
44
45 fitting. (c)-(e) Simulated IR spectra of the Ph-type (p-), O-type, and CH-O-type isomers
46 of H+PhOH-W1, respectively. Relative energies of the isomers are shown in kJ/mol.
47
48
49
50
51
52
53
54
55
56
57 43
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 44 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45 Figure 3 (a) Observed IR spectrum of H+PhOH-W2. (b) Lorentzian fitting of the
46 observed spectrum after slight smoothing. The upper trace is the residual in the
47
fitting. (c)-(d) Simulated IR spectra of the Ph-type and O-type isomers of H+PhOH-W2,
48
49 respectively. Relative energies of the isomers are shown in kJ/mol.
50
51
52
53
54
55
56
57 44
58
59
60 ACS Paragon Plus Environment
Page 45 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46 Figure 4 (a) Observed IR spectrum of H+PhOH-W3. (b) Lorentzian fitting of the
47
observed spectrum after slight smoothing. The upper trace is the residual in the
48
49 fitting. (c)-(e) Simulated IR spectra of stable isomers of H+PhOH-W3. Relative energies
50 of the isomers are shown in kJ/mol. (f) IR spectrum of H+(H2O)3-benzene. Reprinted
51
52 with permission from ref. 13. Copyright 2012 American Chemical Society.
53
54
55
56
57 45
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 46 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
Figure 5 (a) Observed IR spectrum of H+PhOH-W4. (b)-(e) Simulated IR spectra of
48
49 stable isomers of H+PhOH-W4. Relative energies of the isomers are shown in kJ/mol.
50
51
52
53
54
55
56
57 46
58
59
60 ACS Paragon Plus Environment
Page 47 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39 Figure 6 (a) Observed IR spectrum of H+PhOH-W5. (b)-(d) Simulated IR spectra of
40
low energy stable isomers of H+PhOH-W5. Relative energies of the isomers are shown
41
42 in kJ/mol.
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57 47
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 48 of 55

1
2
3
4
5 TOC figure
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57 48
58
59
60 ACS Paragon Plus Environment
Page 49 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45 Figure 1
46
47 187x241mm (300 x 300 DPI)
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 50 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45 Figure 2
46
47 182x247mm (300 x 300 DPI)
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 51 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45 Figure 3
46
47 160x248mm (300 x 300 DPI)
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 52 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45 Figure 4
46
47 163x275mm (300 x 300 DPI)
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 53 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45 Figure 5
46
47 153x267mm (300 x 300 DPI)
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 54 of 55

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45 Figure 5
46
47 163x216mm (300 x 300 DPI)
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 55 of 55 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Figure for Table of Contents
34
35 120x99mm (300 x 300 DPI)
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment

You might also like