You are on page 1of 26

Subscriber access provided by HACETTEPE UNIVERSITESI KUTUPHANESI

Article
Reaction of Aniline with Singlet Oxygen (O #g)
2 1

Jomana Al-Nu’airat, Mohammednoor K. Altarawneh, Xiangpeng


Gao, Phillip R. Westmoreland, and Bogdan Z Dlugogorski
J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.7b00765 • Publication Date (Web): 13 Apr 2017
Downloaded from http://pubs.acs.org on April 14, 2017

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 free 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 accessible to all
readers and 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.

The Journal of Physical Chemistry A 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 25 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 85x88mm (150 x 150 DPI)
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 2 of 25

1
2
3 Reaction of Aniline with Singlet Oxygen (O2 1∆g)
4
5
6
7 Jomana Al-Nu’airat1, Mohammednoor Altarawneh*1, Xiangpeng Gao1,
8 Phillip R. Westmoreland2 and Bogdan Z. Dlugogorski1
9
10
11 1
12 School of Engineering and Information Technology
13 Murdoch University, 90 South Street, Murdoch, WA 6150, Australia
14
15
16 2
17 Department of Chemical and Biomolecular Engineering
18
19 North Carolina State University, Raleigh NC 27695-7905, USA
20
21
22 *
Corresponding author
23
24 Phone: (+61) 8 9360 7507, E-mail: M.Altarawneh@murdoch.edu.au
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
58
59
60 1
ACS Paragon Plus Environment
Page 3 of 25 The Journal of Physical Chemistry

1
2
3 Abstract
4
5
6 Dissolved organic matter (DOM) acts as an effective photochemical sensitiser that produces
7
8 the singlet delta state of molecular oxygen (O2 1∆g), a powerful oxidiser that removes aniline
9
10 from aqueous solutions. However, the exact mode of this reaction, the para to ortho-
11 iminobenzoquinone ratio, and the selectivity of one over the other remain largely speculative.
12
13 This contribution resolves these uncertainties. We report, for the first time, a comprehensive
14
15 mechanistic and kinetic account of the oxidation of aniline with the singlet delta oxygen
16 using B3LYP and M06 functionals in both gas and aqueous phases. Reaction mechanisms
17
18 have been mapped out at E, H and G scales. The 1,4-cycloaddition of O2 1∆g to aniline forms
19
20 a 1,4-peroxide intermediate (M1), which isomerises via a closed-shell mechanism to generate
21 a para-iminobenzoquinone molecule. On the other hand, the O2 1∆g ene-type reaction forms
22
23 an ortho-iminobenzoquinone product when the hydroperoxyl bond breaks, splitting hydroxyl
24
25 from the 1,2-hydroperoxide (M3) moiety. The gas-phase model predicts the formation of
26 both para and ortho iminobenzoquinones. In the latter model, the M1 adduct displays the
27
28 selectivity of up to 96 %. A water-solvation model predicts that M1 decomposes further,
29
30 forming only para-iminobenzoquinone with a rate constant of k = 1.85 × 109 (L/(mol s)) at T
31 = 313 K. These results corroborate the recent experimental findings of product concentration
32
33 profile in which para-iminobenzoquinonine represents the only detected product.
34
35
36
37
38 Key words: B3LYP, M06, Aniline, DOM, Singlet oxygen, Para-iminobenzoquinone, and
39
40 Ortho-iminobenzoquinone.
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 2
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 4 of 25

1
2
3 1. Introduction
4
5
6 Aniline (phenylamine, C6H5NH2) acts as an important precursor or intermediate in
7
8 manufacturing of isocyanates, polymers, dyes, rubbers, pesticides, pharmaceuticals and
9
10 explosives.1,2 However, its hazardous properties, once aniline enters the environment
11 overshadow its industrial merits, especially in aquatic media.3,4 Concentrations of aniline, as
12
13 low as a few ppb in water, induce severe effects on ecosystems and on human health.3,5,6
14
15 Industrial discharges, forest fires, tobacco smoking, and automobile exhaust fumes represent
16 the primary routes for aniline pollution.1,7 Aniline constitutes a hazardous organic compound
17
18 according to US Environmental Protection Agency8 because of its high toxicity and its
19
20 chemical and biochemical stability9 that renders the traditional water treatment
21 technologies10–12 ineffective, such as ozone oxidation and biological treatments. This
22
23 stability makes it necessary to deploy other techniques, such as photodegradation2,13,14 that
24
25 relies on reactive oxygen species (ROS) to remove aniline from water.
26
27
15–18
28 In aquatic systems, dissolved organic matter (DOM) functions as a photoreaction and
29
30 photodegradation agent capable of removing free aniline even at temperatures close to a
31 freezing point.14–22 DOM consists of humic materials derived from the dead cells of plants,
32
33 bacteria, and algae that facilitate the generation of ROS, most notably singlet delta oxygen
34 .
35 (O2 1∆g), hydroxyl radical (OH), superoxide anion (O2 -), and hydrogen peroxide (H2O2), 23

36
37 among which the singlet delta oxygen is the most reactive.24,25 The formation of ROS
38 involves light excitation of complex molecules, such as DOM or rose bengal dye, which
39
40 decay to their ground state in processes that generate ROS. Singlet delta oxygen resides 94.0
41
42 kJ/mol (∆GHo298) above the ground-state triplet oxygen (O2 3Σ-g). Literature documents well
43 the appearance of singlet oxygen in natural water containing high loading of humic
44
45 substances, with a quantum yield varying from 1 to 3 %.26–28
46
47
48 Prior experimental studies have demonstrated the efficient decomposition of aniline by
49
50 singlet delta oxygen.29–31 For example, the photo-induced oxidation of aniline with singlet
51
52 delta oxygen29,32,33 yields para and ortho iminobenzoquinones (C6H5NO), as detected in
53 experiments. However, conflicting findings exist concerning the final product ratio and the
54
55 selectivity of one over the other, as well as the presence of other competing pathways that
56 .
57 might involve superoxide radical anion (O2 -).
58
59
60 3
ACS Paragon Plus Environment
Page 5 of 25 The Journal of Physical Chemistry

1
2
3
4
5 Briviba et al.29 have suggested para-iminobenzoquinone as a key product of the reaction of
6 aniline with singlet delta oxygen in water, apparently consistent with the findings of
7
8 Karunakaran and Karuthapandian who have also reported para-iminobenzoquinone as the
9 .
10 sole product of the oxidation of diphenylamine34 with the superoxide radical-anion (O2 -).
11 29
12 This, however, raises the question of validity of the Briviba et al. conclusion of the
13 formation of the para-iminobenzoquinone via a singlet delta oxygen rather than the
14
15 superoxide radical-anion pathway.
16
17
18 Regarding products variety and selectivity, Booth and Boyland32 and Parke 35
have detected
19
20 both the ortho and para-aminophenol isomers arising by photoreaction of aniline with singlet
21
22 delta oxygen in rat and cat liver microsomes. However, Parke 35 published inconsistencies in
23 the para to ortho-aminophenol product ratio varying from 2:5 in a cat to 15:l in gerbil. Given
24
25 that, the studies of Booth and Boyland as well as of Parke were carried out in a biological
26
27 framework, catalytic reactions of enzymes might have influenced the product distribution.
28
29
30 The highly transient nature of singlet oxygen, especially its lifetime of 4 µs, makes the
31
32 experimental determination of the precise mode of the reaction between aniline and singlet
33 oxygen a challenging task.36 The involvement of other ROS and emergence of parallel
34
35 reactions render it even harder to elucidate precise pathways driven solely by singlet oxygen.
36
37 This motivates the present in silico study, in which we report a comprehensive mechanistic
38 and kinetic account of pathways encountered in the oxidative transformation of aniline by
39
40 singlet delta oxygen. We map out reaction routes and assess their relative feasibility based on
41
42 computed reaction rate constants. The experimental profiles of product species reported in
43 literature29 show very good agreement with the results of the present calculations.
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 4
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 6 of 25

1
2
3 2. Methodology
4
5
6 The Gaussian 0937 suite of programs affords all structural optimisations and energy
7
8 calculations (∆≠Ho298, ∆rHo298, ∆≠Go298, and ∆rGo298). The theoretical methodology comprises
9
10 the gradient-corrected density functional theory (DFT) functionals B3LYP38,39 and the meta-
11 hybrid M0640 functional, both, supplemented with the polarised basis set of 6-311+G(d,p).41
12
13 Furthermore, we have performed intrinsic reaction coordinate (IRC) calculations, confirming
14
15 the identity of each transition structure.
16
17
18 While the B3LYP function retains a relatively cost-effective accuracy in predicting
19
20 thermochemistry for organic species, even in systems encompassing diradicals,42 it often
21 suffers fundamental shortcomings in predicting accurate kinetic parameters43. For this
22
23 reason, we have also adopted the M06 method to obtain barriers of enthalpy and Gibbs free
24
25 energy. We simulate reactions in the aqueous media by deploying the polarisable continuum
26 model (PCM).44 Simulating aqueous-phase reactions based on gas phase models proves to
27
28 represent an adequate and representative model for reactions occurring in aqueous media.42
29
30 Albeit, we expect the energetic profiles representing water-phase reactions to fall below those
31 associated with their gas-phase analogues.45
32
33
34
35 The approximate spin-projection (AP) method pioneered by the Yamaguchi group 46,47 serves
36 to refine the enthalpies and Gibbs free energies of activation and reaction (∆≠Ho298, ∆rHo298,
37
38 ∆≠Go298 and ∆rGo298) of all singlet species that suffer from contamination by their lower-
39
40 energy triplet state. As expected, a significant difference between energies on G and H scales
41 for some reactions (especially for bimolecular channels) indicates entropy effects. We
42
43 eliminate the spin contamination error of singlet delta oxygen based on the broken symmetry
44
45 (BS) solution by performing structural optimisation with mixing the frontier molecular
46 orbitals (HOMO and LUMO). Adapting this approach yields satisfactory performance in
47
48 comparison with the more computationally expensive multi-reference wave-function methods
49
50 such as CASSCF and CASSTP2.48,49
51
52
53 The KiSThelP package provides a means to estimate reaction rate constants fitted in the
54
55 temperature range of 300 to 600 K.50 The kinetic parameters are estimated based on the
56 conventional transition state theory with the inclusion of one dimensional Eckart barrier,
57
58 accounting for quantum tunnelling effect contribution. We treat the internal rotations in
59
60 5
ACS Paragon Plus Environment
Page 7 of 25 The Journal of Physical Chemistry

1
2
3 aniline and all derived species that involve the amine group as hindered rotors with the
4
5 overall enthalpy barrier of 23.2 kJ/mol. The rotation enthalpy of the NH2 group is derived
6 from a recent estimate of Zanganeh et al.51 Finally, the DMol3 code52 facilitates the
7
53
8 calculations of the Hirshfeld atomic charges and the Fukui (ƒ1) indices.54 The Fukui
9
10 indices indicate the tendency of a molecular site to undergo an electrophilic addition. The
11 choice of the Hirshfeld scheme stems from its superior performance in estimating atomic
12
13 charges in reference to other approaches, such as the commonly deployed Mulliken
14
15 population analysis55 that often displays severe sensitivity to the implemented basis set.
16
17
18
19
20 3. Results and Discussions
21
22
23 3.1 Preliminary Analysis of Results from M06 and B3LYP Approaches
24
25
26 We first assess the performance of several DFT methods at 298.15 K in calculating the
27
28 enthalpy gap between singlet and triplet states (∆GH298) for selected diradical species, namely
29
30 O, O2, NH, NF, C, and Si. Table S1 in the Supporting Information (S1) gives calculated
31 energy and spin contamination operator <S2> terms for the aforementioned diradicals at
32
33 several theoretical levels. Table S2 lists (∆GH298) and the corrected calculated gap based on
34
35 the broken symmetry solution (∆GH298S-T) values for the chosen diradicals. Table S2
36 indicates that ∆GH298S-T estimates obtained from the M06 calculations fall below those
37
38 provided by the B3LYP computations. In the case of the oxygen molecule, M06 significantly
39
40 overestimates the ∆GH298S-T value by 37.7 kJ/mol, while B3LYP shows an excellent
41 performance with only 8.5 kJ/mol offset from the experimental value (94.6 kJ/mol)56. The
42
43 noticeable difference in the ∆GH298S-T value between the two methods confirms the relative
44
45 accuracy of B3LYP in determining thermochemical values for oxygen. However, except for
46 O2, M06 yields ∆GH298S-T values that are closer to experimental measurements than those
47
48 computed with B3LYP. Ess and Cook57 reported similar results for diradicals, with a mean
49
50 unsigned error (MUS) of 12.6 and 41.4 kJ/mol for M06 and B3LYP, respectively. B3LYP
51 does, however, have some limitations in computing reaction barriers, with the M06 family of
52
53 functionals delivering more satisfactory performance, well documented in literature.40,58
54
55 Thus, in this contribution, we report reaction and activation energies for all considered
56 reactions using both the two methods.
57
58
59
60 6
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 8 of 25

1
2
3
4
5 3.2 Characterisation of Singlet Oxygen and Aniline
6
7
8 Singlet oxygen is an electrophilic agent. Thus, unlike triplet oxygen, introducing singlet
9
10 oxygen atoms into olefins, dienes, or aromatic structures, such as aniline, proceeds through
11 distinct routes of Diels-Alder [4+2] and [2+2] cycloadditions, and ene reactions with isolated
12
13 double bonds.59–61 Olefins and substances with inaccessible allylic hydrogens represent the
14
15 primary candidates to undergo [2+2]-cycloaddition with singlet oxygen, forming
16 dioxetanes.62 In contrast, allylic hydrogen in the more congested side of olefins experiences
17
18 the so-called ene reaction, forming allylic hydroperoxides.63 Singlet oxygen reacts with
19
20 structural entities of 1,3-dienes (cyclic, noncyclic or aromatic) to produce endoperoxides.64
21 Determination of the dominant reaction rests on different electronic and structural factors,
22
23 including the distance between C1 and C4 atoms in dienes, ionisation potential of the 1,3-
24
25 diene, spatial alignment of allylic hydrogens, and the type of the solvent deployed.62,65
26
27
28 Aniline reaction with singlet oxygen exemplifies a process controlled by frontier molecular
29
30 orbitals, with singlet delta oxygen LUMOs interacting with aniline HOMOs. When the
31 amine group is added to the benzene ring, it decreases the energy gap of the HOMO (-5.084
32
33 eV) and LUMO (-1.055 eV) orbitals in comparison with that of the HOMO (-5.388 eV) and
34
35 LUMO (-0.168 eV) of benzene.66,67 This change means that the ring becomes more reactive
36 with singlet oxygen in the presence of an amine substituent. Figure 1 indicates the Fukui
37
38 indices and Hirshfeld electronic charges of aniline with respect to benzene. Higher f-1 values
39
40 at the para, ortho, and meta carbon atoms of aniline with respect to the indistinguishable sites
41 in benzene signify the addition of singlet delta oxygen to aniline proceeding at a higher rate
42
43 when compared with the analogous process in benzene. The higher f-1 associated with the
44
45 para carbon further highlights the para position as a more preferred site for the electrophilic
46 O2 1∆g addition in reference to ortho and meta sites. This appears to agree with the HOMO
47
48 map of aniline that shows the HOMO orbitals localised mainly on the para and ortho carbon
49
50 atoms; see Figure S1 in SI.
51
52
53
54
55 3.3 Reaction Mechanism
56
57
58
59
60 7
ACS Paragon Plus Environment
Page 9 of 25 The Journal of Physical Chemistry

1
2
3 Figure 2 depicts the initial corridors in the aniline + O2 1∆g mechanism which branches into
4
5 five channels via three routes, namely [4 + 2] cycloaddition reactions affording M1 and M2
6 1,4 peroxides, ene-type reactions producing M3 and M4 1,2/1,4 hydroperoxides, and direct
7
8 H-abstraction from the amine group generating the M5 anilino radical. Table S3 assembles
9
10 estimated reaction and activation enthalpies and Gibbs free energies based on the gas-phase
11 (B3LYP and M06) and solvation (M06) models. The [2+2] reaction was not observed, as
12
13 expected in aromatic rings holding only one electron-donating group such as an amine
14
15 substitute. Figures S2 and S3 compare geometries of the located transition structures for the
16 three theoretical approaches considered in this article.
17
18
19
20 The attack of the singlet oxygen at para-ipso and ortho-meta positions yields the two
21 peroxides structures M1 and M2, respectively. The enthalpy barrier in M1, characterised by
22
23 the transition state TS1, is considerably lower than that embedded in TS2 to form M2; i.e., 79
24
25 kJ/mol versus 101 kJ/mol, from the B3LYP calculations. By applying the Curtin-Hammond
26 principle33, the para-meta product ratio ([Ppara]/[Pmeta]) is dictated by the free energy gap
27
28 (∆Go298) between the para and meta transition states (TS1 and TS2) rather than the stabilities
29
30 of the two intermediates (M1 and M2); i.e., ∆GGo298P-M= 22 kJ/mol, B3LYP, see Figure 2 and
31 Table S3. Our ratios of [Ppara]/[Pmeta] at 298.15 K correspond to 9160:1 and 215:1, for the
32
33 B3LYP and M06 functionals, respectively. This finding indicates an expected dominance of
34
35 the para isomer and a negligible abundance of the meta isomer. This excludes the meta-
36 channel from further mechanistic analysis.
37
38
39
40 Hydroperoxides M3 and M4 originate from the ene-type reactions involving abstraction of
41 one of the amine hydrogens and simultaneous addition of the singlet oxygen at ortho and
42
43 para sites, respectively. The free energy of TS3 leading to the formation of the M3 adduct
44
45 holds an intermediate barrier (4 kJ/mol, M06) comparing with the para and meta initial
46 reaction channels. Unexpectedly, B3LYP ortho-TS3 assumes a slightly lower enthalpy
47
48 barrier than its para analogue (TS1), see Table S3. To examine the performance of the
49
50 B3LYP further, we map out the HOMO and LUMO of reactants and transition states in
51 Figures S1 and S4 in SI; respectively. Interestingly, the amine group visibly interacts with
52
53 HOMO and LUMO of TS3, lowering its enthalpy barrier. Conversely, the HOMO and
54
55 LUMO of para (TS1) are mainly located between the benzene ring and singlet delta oxygen,
56 with no apparent contribution from the amine substituent. TS4 requires a sizable activation
57
58 barrier of 147 kJ/mol (M06) rendering this route inaccessible. Direct abstraction of an H
59
60 8
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 10 of 25

1
2
3 atom from the amine group to form an anilino radical occurs through a barrier of 147 kJ/mol
4
5 (M06), via TS5. In general, both of the methods M06 and B3LYP provide the same reaction
6 pathways. However, the reaction enthalpies and free energies of reaction for the formation of
7
8 the M1-M5 intermediates in the gas phase, from B3LYP and M06 computations, differ by
9
10 17-80 kJ/mol (∆rHo298, Table S3 and Figure 2). This difference accounts for the lower
11 enthalpy barrier estimated for singlet oxygen by the M06 functional. Thus, even though both
12
13 of the functionals yield the same trends, M06 enthalpy barriers fall below those calculated
14
15 with the B3LYP functional.
16
17
18 Table 1 lists fitted modified Arrhenius parameters for the five entrance channels, and Figure
19
20 3 depicts branching ratios of these reactions. For the M06 functional, the formation of adduct
21 M1 dominates the fate of the reactants (aniline and singlet delta oxygen) in both gaseous
22
23 (Figure 3a) and aqueous (Figure 3b) media. The trend based on the M06 calculations accords
24 29
25 with the findings of Briviba et al. in which the para–channel assumes chief importance;
26 i.e., singlet delta oxygen selectively oxidises aniline to form para-iminobenzoquinone.
27
28 Despite the discrepancy in pinning down the major initial channel from the two media, we are
29
30 now in the position to limit the most critical operating pathways to the formation of M1 and
31 M3. Consequently, we mapped out the free energy surfaces for the further decompositions of
32
33 M1 and M3 adducts in Figures 4 and 5, respectively, with their fitted Arrhenius parameters
34
35 listed in Table 1. Moreover, in order to confirm pathways leading to formation of M1 and
36 M2, we mapped out the intrinsic reaction coordinates (IRC) leading to the formation of M1
37
38 and M2 in Figure S5 and Figure S6 (SI); correspondingly.
39
40
41
42
43 As Figure 4 illustrates, the unimolecular isomerisation of the M1 moiety proceeds via a
44
45 closed-shell mechanism to generate the 4-iminobenzoquinone molecule. The final products
46 along this route (M06) fall 471 kJ/mol below the entrance channel, and the overall enthalpy
47
48 barrier of 10 kJ/mol (M06) corresponds to the initial 1,4-cycloaddition step. The elementary
49
50 reactions comprise the formation of two epoxy intermediates (M6 and M7) and water
51 elimination from the M8 adduct. M1 stabilises into M6 in one step without encountering an
52
53 intrinsic reaction barrier. Inspection of the mechanism in Figure 4 reveals that, the water
54
55 elimination step demands a significant enthalpy barrier of 219 kJ/mol (M06) through TS8
56 over the highly stabilised adduct M8. Nonetheless, TS8 remains submerged by 191 kJ/mol
57
58 (M06) with respect to the initial reactants.
59
60 9
ACS Paragon Plus Environment
Page 11 of 25 The Journal of Physical Chemistry

1
2
3 The formation of the ortho-iminobenzoquinone follows the fission of the hydroperoxyl bond
4
5 in the M3 intermediate (Figure 5). Endothermicity of this process amounts to 120 kJ/mol
6 (solvation) and produces the M10 radical. Direct elimination of the out-of-plane H atom
7
8 from M10 forms the ortho-iminobenzoquinone through a considerable exothermic reaction of
9
10 403 kJ/mol (solvation). In microorganisms, should the metabolism of aniline lead to ortho-
11 aminophenol in the detoxification process, it would cause more harm than good, owing to the
12
13 fact that ortho-aminophenol incurs more toxicity than the parent aniline.35 Dennis et al.68
14
15 attempted to explain the ortho-product formation from a biological point of view based on the
16 presence of hydroxylation enzyme, saturation of the detoxification pathways, or DNA repair
17
18 system.
19
20
21 In Figure S7, we compared the enthalpies with Gibbs free energies of activation and reaction
22
23 for the formation of para-iminobenzoquinone in solution. A negative value of ∆rGo298
24
25 associated with a reaction reveals that this reaction may occur spontaneously at 298 K.
26 Higher values of barriers of Gibbs free energy in comparison to enthalpies for TS6 and TS7
27
28 point to positive changes in entropy, and hence to tight transition structures. The remaining
29
30 transition structures display loose transition structures.
31
32
33 Applying the values of the kinetic parameters for solution from the first row in Table 2, and
34
35 converting the units to L and mol, our calculated rate constant of photoxidation of aniline
36 with singlet delta oxygen, at 313 K, of k = 1.85 × 109 (L/(mol s)) concurs with the
37
38 experimental measurement of k = 2.0 × 109 (L/ (mol s)) 1, in methanol solution.69 With the
39
40 aid of PolyMath70 software, we solved a system of ordinary differential equations (ODEs) for
41 our solvation model, assuming a batch reactor with a steady state approximation for singlet
42
43 oxygen formation and a sensitiser quantum yield of 0.52, see Table S4 and Figure S8 in the
44
45 SI. The initial concentrations corresponded to 35 µmol/L and 1 mmol/L for methylene blue
46 sensitiser and aniline, respectively. In Figure 6, we compared the results of our solvation
47
48 model with the experimental measurements of Briviba et al.29, demonstrating close agreement
49
50 with an overall reaction rate of para-iminobenzoquinone formation amount to 3.54 (mol /(L
51 s)).
52
53
54
55
56
57
58
59
60 10
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 12 of 25

1
2
3 4. Conclusion
4
5
6 This contribution addressed the mechanistic and kinetic aspects pertinent of the photo-
7
8 induced reaction of aniline with singlet oxygen. The gaseous reaction pathways obtained
9
10 from the application of the B3LYP and M06 functionals predict the formation of para and
11 ortho-iminobenzoquinones. A water solvation model reveals the formation of para–
12
13 iminobenzoquinone as the predominant product. Moreover, our time-concentration history of
14
15 para–iminobenzoquinone matches the corresponding experimental profile of Briviba et al.29
16 well. Our findings resolved the lack of consensus on literature on the expected products, their
17
18 selectivity ratios and reaction mechanism of aniline reaction with O2 1∆g.
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
58
59
60 11
ACS Paragon Plus Environment
Page 13 of 25 The Journal of Physical Chemistry

1
2
3 Supporting Information Available
4
5
6 Table S1 (Singlet energy HSinglet, triplet energy HTriplet and spin unrestricted H(Singlet,Guess=Mix)
7
8 energy calculated by B3LYP, HSEH1PBE, B3PW91, CAM-B3LYP, and M06 for O, O2,
9
10 NH, NF, C, and Si diradicals in addition to the spin contaminations S2 Singlet and S2 Triplet for
11 singlet and triplet states, respectively), Table S2 Spin-projected (singlet-triplet) gap energy
12
13 (∆GH298S-T) for (O, O2, NH, NF, C, and Si) diradicals and the (singlet-triplet) gap energy
14
15 ∆GH298 for spin unrestricted B3LYP, HSEH1PBE, B3PW91, CAM-B3LYP, and M06
16 functionals), Spin-projected (singlet-triplet) gap enthalpy (∆GH298S-T) for O, O2, NH, NF, C,
17
18 and Si diradicals and the normal (singlet-triplet) gap energy ∆GH298 for spin unrestricted
19
20 B3LYP, HSEH1PBE, B3PW91, CAM-B3LYP, and M06 functionals. All energy values are
21 in kJ/mol. Table S3 Enthalpies and Gibbs free energies of activation and reaction (∆≠Ho298,
22
23 ∆rHo298, ∆≠Go298 and ∆rGo298) calculated based on gas-phase (B3LYP in italic and M06 in in
24
25 brackets) and solvation (M06) models. Table S4 PolyMath70 ordinary differential equations
26 (ODEs) and parameters for the solvation model, Figure S1 HOMO of transition states,
27
28 Figures S2 and S3 Optimised structures of transition states of aniline oxidation by singlet
29
30 oxygen, and Figure S4 LUMO of transition states. Figure S5 IRC calculation (M06/6-
31 311+G(d,p)) for the reaction pathway of aniline with singlet oxygen via TS1, Figure S6 IRC
32
33 calculation (M06/6-311+G(d,p)) for the reaction pathway of aniline with singlet oxygen via
34
35 TS3, Figure S7 Enthalpy map (blue) and Gibbs free energy map (red) for the formation of
36 para-iminobenzoquinone from the reaction of singlet oxygen with aniline using M06
37
38 (solvation) model and Figure S8 Concentration profile for aniline photoreaction with singlet
39
40 oxygen via the water-solvation model-M06 at 313 K. This material is available free of
41 charge via the Internet at http://pubs.acs.org.
42
43
44
45
46
Conflict of Interest: The authors declare no conflict of interest.
47
48
49
50 Acknowledgement
51
52
53
54 We gratefully acknowledge grants of computing time from the National Computational
55 Infrastructure (NCI) and from the Pawsey Supercomputing Centre, Australia as well as the
56
57
58
59
60 12
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 14 of 25

1
2
3 financial supports from the Australian Research Council (ARC). J.N. thanks Murdoch
4
5 University for the award of a MUSS postgraduate scholarship.
6 References
7
8
9
10 (1) Pinheiro, H. M.; Touraud, E.; Thomas, O. Aromatic Amines from Azo Dye Reduction:
11 Status Review with Emphasis on Direct UV Spectrophotometric Detection in Textile
12 Industry Wastewaters. Dyes Pigments 2004, 61 (2), 121–139.
13 (2) Gautam, S.; Kamble, S. P.; Sawant, S. B.; Pangarkar, V. G. Photocatalytic Degradation
14 of 4-Nitroaniline Using Solar and Artificial UV Radiation. Chem. Eng. J. 2005, 110
15 (1–3), 129–137.
16 (3) Buikema, A. L.; McGinniss, M. J.; Cairns, J. Phenolics in Aquatic Ecosystems: A
17
Selected Review of Recent Literature. Mar. Environ. Res. 1979, 2 (2), 87–181.
18
19 (4) Shackelford, W. M.; Cline, D. M.; Faas, L.; Kurth, G. An Evaluation of Automated
20 Spectrum Matching for Survey Identification of Wastewater Components by Gas
21 Chromatography—mass Spectrometry. Anal. Chim. Acta 1983, 146, 15–27.
22 (5) Garcia, N. A. New Trends in Photobiology: Singlet-Molecular-Oxygen-Mediated
23 Photodegradation of Aquatic Phenolic Pollutants. A Kinetic and Mechanistic
24 Overview. J. Photochem. Photobiol. B 1994, 22 (3), 185–196.
25 (6) Snyderwine, E. G.; Sinha, R.; Felton, J. S.; Ferguson, L. R. Highlights of the Eighth
26 International Conference on Carcinogenic/mutagenic N-Substituted Aryl Compounds.
27
Mutat. Res. 2002, 506–507, 1–8.
28
29 (7) Stanley, T. W.; Sawicki, E.; Johnson, H.; Pfaff, J. D. Comparison of Methods for the
30 Spectrophotometric Determination of Phenols in Automobile Exhaust Fumes.
31 Microchim. Acta 1965, 53 (1), 48–59.
32 (8) US EPA, O. Aniline | Technology Transfer Network Air Toxics Web site | US EPA
33 https://www3.epa.gov/ttn/atw/hlthef/aniline.html≠ref1 (accessed Aug 31, 2016).
34 (9) Shahrezaei, F.; Mansouri, Y.; Zinatizadeh, A. a. L.; Akhbari, A.; Shahrezaei, F.;
35 Mansouri, Y.; Zinatizadeh, A. a. L.; Akhbari, A. Photocatalytic Degradation of Aniline
36 Using TiO2 Nanoparticles in a Vertical Circulating Photocatalytic Reactor,
37
Photocatalytic Degradation of Aniline Using TiO2 Nanoparticles in a Vertical
38
39 Circulating Photocatalytic Reactor. Int. J. Photoenergy Int. J. Photoenergy 2012.
40 (10) Spain, J. C. Biodegradation of Nitroaromatic Compounds. Annu. Rev. Microbiol. 1995,
41 49 (1), 523–555.
42 (11) Lopez, T.; Gomez, R.; Sanchez, E.; Tzompantzi, F.; Vera, L. Photocatalytic Activity in
43 the 2,4-Dinitroaniline Decomposition Over TiO2 Sol-Gel Derived Catalysts. J. Sol-Gel
44 Sci. Technol. 2001, 22 (1–2), 99–107.
45 (12) Zhao, G.; Lu, X.; Zhou, Y. Aniline Degradation in Aqueous Solution by UV-Aeration
46 and UV-microO3 Processes: Efficiency, Contribution of Radicals and Byproducts.
47
Chem. Eng. J. 2013, 229, 436–443.
48
49 (13) Kamble, S. P.; Sawant, S. B.; Schouten, J. C.; Pangarkar, V. G. Photocatalytic and
50 Photochemical Degradation of Aniline Using Concentrated Solar Radiation. J. Chem.
51 Technol. Biotechnol. 2003, 78 (8), 865–872.
52 (14) Talebian, N.; Jafarinezhad, F. Morphology-Controlled Synthesis of SnO2
53 Nanostructures Using Hydrothermal Method and Their Photocatalytic Applications.
54 Ceram. Int. 2013, 39 (7), 8311–8317.
55 (15) Canonica, S.; Hellrung, B.; Müller, P.; Wirz, J. Aqueous Oxidation of Phenylurea
56 Herbicides by Triplet Aromatic Ketones. Environ. Sci. Technol. 2006, 40 (21), 6636–
57
6641.
58
59
60 13
ACS Paragon Plus Environment
Page 15 of 25 The Journal of Physical Chemistry

1
2
3 (16) Canonica, S. Oxidation of Aquatic Organic Contaminants Induced by Excited Triplet
4 States. Chim. Int. J. Chem. 2007, 61 (10), 641–644.
5 (17) Canonica, S.; Hellrung, B.; Wirz, J. Oxidation of Phenols by Triplet Aromatic Ketones
6 in Aqueous Solution. J. Phys. Chem. A 2000, 104 (6), 1226–1232.
7
(18) Canonica, S.; Jans, U. R. S.; Stemmler, K.; Hoigne, J. Transformation Kinetics of
8
9 Phenols in Water: Photosensitization by Dissolved Natural Organic Material and
10 Aromatic Ketones. Environ. Sci. Technol. 1995, 29 (7), 1822–1831.
11 (19) Erickson, P. R.; Walpen, N.; Guerard, J. J.; Eustis, S. N.; Arey, J. S.; McNeill, K.
12 Controlling Factors in the Rates of Oxidation of Anilines and Phenols by Triplet
13 Methylene Blue in Aqueous Solution. J. Phys. Chem. A 2015, 119 (13), 3233–3243.
14 (20) Weber, E. J.; Spidle, D. L.; Thorn, K. A. Covalent Binding of Aniline to Humic
15 Substances. 1. Kinetic Studies. Environ. Sci. Technol. 1996, 30 (9), 2755–2763.
16 (21) Eriksson, J.; Frankki, S.; Shchukarev, A.; Skyllberg, U. Binding of 2, 4, 6-
17
Trinitrotoluene, Aniline, and Nitrobenzene to Dissolved and Particulate Soil Organic
18
19 Matter. Environ. Sci. Technol. 2004, 38 (11), 3074–3080.
20 (22) Fukushima, M.; Tatsumi, K.; Morimoto, K. The Fate of Aniline after a Photo-Fenton
21 Reaction in an Aqueous System Containing Iron (III), Humic Acid, and Hydrogen
22 Peroxide. Environ. Sci. Technol. 2000, 34 (10), 2006–2013.
23 (23) Appiani, E.; McNeill, K. Photochemical Production of Singlet Oxygen from
24 Particulate Organic Matter. Environ. Sci. Technol. 2015, 49 (6), 3514–3522.
25 (24) Sandvik, S. L. H.; Bilski, P.; Pakulski, J. D.; Chignell, C. F.; Coffin, R. B.
26 Photogeneration of Singlet Oxygen and Free Radicals in Dissolved Organic Matter
27
Isolated from the Mississippi and Atchafalaya River Plumes. Mar. Chem. 2000, 69 (1),
28
29 139–152.
30 (25) Ray, P. Z.; Tarr, M. A. Solar Production of Singlet Oxygen from Crude Oil Films on
31 Water. J. Photochem. Photobiol. Chem. 2014, 286, 22–28.
32 (26) Zepp, R. G.; Wolfe, N. L.; Baughman, G. L.; Hollis, R. C. Singlet Oxygen in Natural
33 Waters. 1977.
34 (27) Haag, W. R.; Hoigne, J. Singlet Oxygen in Surface Waters. 3. Photochemical
35 Formation and Steady-State Concentrations in Various Types of Waters. Environ. Sci.
36 Technol. 1986, 20 (4), 341–348.
37
(28) Frimmel, F. H.; Bauer, H.; Putzien, J.; Murasecco, P.; Braun, A. M. Laser Flash
38
39 Photolysis of Dissolved Aquatic Humic Material and the Sensitized Production of
40 Singlet Oxygen. Environ. Sci. Technol. 1987, 21 (6), 541–545.
41 (29) Briviba, K.; Devasagayam, T. P. A.; Sies, H.; Steenken, S. Selective Para-
42 Hydroxylation of Phenol and Aniline by Singlet Molecular Oxygen. Chem. Res.
43 Toxicol. 1993, 6 (4), 548–553.
44 (30) Ratti, M.; Canonica, S.; McNeill, K.; Bolotin, J.; Hofstetter, T. B. Isotope
45 Fractionation Associated with the Indirect Photolysis of Substituted Anilines in
46 Aqueous Solution. Environ. Sci. Technol. 2015, 49 (21), 12766–12773.
47
(31) Nilsson, R.; Merkel, P. B.; Kearns, D. R. Unambiguous Evidence For The Participation
48
49 of Singlet Oxygen (1δ) in Photodynamic Oxidation of Amino Acids. Photochem.
50 Photobiol. 1972, 16 (2), 117–124.
51 (32) Booth, J.; Boyland, E. The Biochemistry of Aromatic Amines. 3. Enzymic
52 Hydroxylation by Rat-Liver Microsomes. Biochem. J. 1957, 66 (1), 73–78.
53 (33) Seeman, J. I. Effect of Conformational Change on Reactivity in Organic Chemistry.
54 Evaluations, Applications, and Extensions of Curtin-Hammett Winstein-Holness
55 Kinetics. Chem. Rev. 1983, 83 (2), 83–134.
56
57
58
59
60 14
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 16 of 25

1
2
3 (34) Karunakaran, C.; Karuthapandian, S. Enhancement of TiO2-Photocatalyzed Organic
4 Transformation by ZnO and ZnS. Oxidation of Diphenylamine. Egypt. J. Basic Appl.
5 Sci. 2015, 2 (1), 32–38.
6 (35) Parke, D. V. Studies in Detoxication. 84. The Metabolism of [14C] Aniline in the
7
Rabbit and Other Animals. Biochem. J. 1960, 77 (3), 493–503.
8
9 (36) Clennan, E. L.; Pace, A. Advances in Singlet Oxygen Chemistry. Tetrahedron 2005,
10 61 (28), 6665–6691.
11 (37) Frisch, M.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman,
12 J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; et al. Gaussian 09,
13 Revision A. 02, Gaussian. Inc Wallingford CT 2009, 200.
14 (38) Becke, A. D. Density-Functional Exchange-Energy Approximation with Correct
15 Asymptotic Behavior. Phys. Rev. A 1988, 38 (6), 3098.
16 (39) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti Correlation-Energy
17
Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37 (2), 785.
18
19 (40) Zhao, Y.; Truhlar, D. G. The M06 Suite of Density Functionals for Main Group
20 Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States,
21 and Transition Elements: Two New Functionals and Systematic Testing of Four M06-
22 Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120 (1–3), 215–
23 241.
24 (41) Montgomery Jr, J. A.; Ochterski, J. W.; Petersson, G. A. A Complete Basis Set Model
25 Chemistry. IV. An Improved Atomic Pair Natural Orbital Method. J. Chem. Phys.
26 1994, 101 (7), 5900–5909.
27
(42) Morales, J.; Günther, G.; Zanocco, A. L.; Lemp, E. Singlet Oxygen Reactions with
28
29 Flavonoids. A Theoretical – Experimental Study. PLOS ONE 2012, 7 (7), e40548.
30 (42) Westmoreland, P.; Kollman, P. A.; Chaka, A. M.; Cummings, P. T.; Morokuma, K.;
31 Neurock, M.; Stechel, E. B.; Vashishta, P. Applying Molecular and Materials
32 Modeling; Springer Science & Business Media, 2013.
33 (44) Mennucci, B. Polarizable Continuum Model. Wiley Interdiscip. Rev. Comput. Mol. Sci.
34 2012, 2 (3), 386–404.
35 (45) Rayson, M. S.; Altarawneh, M.; Mackie, J. C.; Kennedy, E. M.; Dlugogorski, B. Z.
36 Theoretical Study of the Ammonia- Hypochlorous Acid Reaction Mechanism. J. Phys.
37
Chem. A 2010, 114 (7), 2597–2606.
38
39 (46) Saito, T.; Nishihara, S.; Kataoka, Y.; Nakanishi, Y.; Matsui, T.; Kitagawa, Y.;
40 Kawakami, T.; Okumura, M.; Yamaguchi, K. Transition State Optimization Based on
41 Approximate Spin-Projection (AP) Method. Chem. Phys. Lett. 2009, 483 (1), 168–171.
42 (47) Yamanaka, S.; Kawakami, T.; Nagao, H.; Yamaguchi, K. Effective Exchange Integrals
43 for Open-Shell Species by Density Functional Methods. Chem. Phys. Lett. 1994, 231
44 (1), 25–33.
45 (48) Bobrowski, M.; Liwo, A.; Oldziej, S.; Jeziorek, D.; Ossowski, T. CAS MCSCF/CAS
46 MCQDPT2 Study of the Mechanism of Singlet Oxygen Addition to 1, 3-Butadiene and
47
Benzene. J. Am. Chem. Soc. 2000, 122 (34), 8112–8119.
48
49 (49) Abe, M. Diradicals. Chem. Rev. 2013, 113 (9), 7011–7088.
50 (50) Canneaux, S.; Bohr, F.; Henon, E. KiSThelP: A Program to Predict Thermodynamic
51 Properties and Rate Constants from Quantum Chemistry Results†. J. Comput. Chem.
52 2014, 35 (1), 82–93.
53 (51) Zanganeh, J.; Altarawneh, M.; Saraireh, I.; Namazi, S.; Zanganeh, J. Theoretical Study
54 on Thermochemical Parameters and pKa Values for Fluorinated Isomers of Toluene.
55 Comput. Theor. Chem. 2013, 1011, 21–29.
56 (52) Delley, B. From Molecules to Solids with the DMol3 Approach. J. Chem. Phys. 2000,
57
113 (18), 7756–7764.
58
59
60 15
ACS Paragon Plus Environment
Page 17 of 25 The Journal of Physical Chemistry

1
2
3 (53) Hirshfeld, F. L. Bonded-Atom Fragments for Describing Molecular Charge Densities.
4 Theor. Chim. Acta 1977, 44 (2), 129–138.
5 (54) Fukui, K. The Role of Frontier Orbitals in Chemical Reactions (Nobel Lecture).
6 Angew. Chem. Int. Ed. Engl. 1982, 21 (11), 801–809.
7
(55) Fonseca Guerra, C.; Handgraaf, J.-W.; Baerends, E. J.; Bickelhaupt, F. M. Voronoi
8
9 Deformation Density (VDD) Charges: Assessment of the Mulliken, Bader, Hirshfeld,
10 Weinhold, and VDD Methods for Charge Analysis. J. Comput. Chem. 2004, 25 (2),
11 189–210.
12 (56) Slipchenko, L. V.; Krylov, A. I. Singlet-Triplet Gaps in Diradicals by the Spin-Flip
13 Approach: A Benchmark Study. J. Chem. Phys. 2002, 117 (10), 4694–4708.
14 (57) Ess, D. H.; Cook, T. C. Unrestricted Prescriptions for Open-Shell Singlet Diradicals:
15 Using Economical Ab Initio and Density Functional Theory to Calculate Singlet–
16 Triplet Gaps and Bond Dissociation Curves. J. Phys. Chem. A 2012, 116 (20), 4922–
17
4929.
18
19 (58) Zhao, Y.; Truhlar, D. G. Exploring the Limit of Accuracy of the Global Hybrid Meta
20 Density Functional for Main-Group Thermochemistry, Kinetics, and Noncovalent
21 Interactions. J. Chem. Theory Comput. 2008, 4 (11), 1849–1868.
22 (58) Albini, A.; Fagnoni, M. Handbook of Synthetic Photochemistry; John Wiley & Sons,
23 2009.
24 (60) Wasserman, E.; Kuck, V. J.; Delavan, W. M.; Yager, W. A. Electron Paramagnetic
25 Resonance of 1. DELTA. Oxygen Produced by Gas-Phase Photosensitization with
26 Naphthalene Derivatives. J. Am. Chem. Soc. 1969, 91 (4), 1040–1041.
27
(60) Foote, C. S. Active Oxygen in Chemistry; Springer Science & Business Media, 1995.
28
29 (62) Clennan, E. L. Synthetic and Mechanistic Aspects of 1, 3-Diene Photooxidation.
30 Tetrahedron 1991, 47 (8), 1343–1382.
31 (63) Bartlett, P. D.; Mendenhall, G. D.; Schaap, A. P. Competitive Modes of Reaction of
32 Singlet Oxygen*. Ann. N. Y. Acad. Sci. 1970, 171 (1), 79–88.
33 (64) Gollnick, K.; Griesbeck, A. Interactions of Singlet Oxygen with 2, 5-Dimethyl-2, 4-
34 Hexadiene in Polar Ano Non-Polar Solvents Evidence for a Vinylog Ene-Reaction.
35 Tetrahedron 1984, 40 (17), 3235–3250.
36 (65) Griesbeck, A. G.; Oelgemöller, M.; Ghetti, F. CRC Handbook of Organic
37
Photochemistry and Photobiology; CRC Press, 2012; Vol. 1.
38
39 (66) Klopman, G.; others. Chemical Reactivity and Reaction Paths. 1974.
40 (67) Van den Heuvel, C. J. M.; Hofland, A.; Steinberg, H.; de Boer, T. J. The Photo-
41 Oxidation of Hexamethylbenzene and Pentamethylbenzene by Singlet Oxygen. Recl.
42 Trav. Chim. Pays-Bas 1980, 99 (9), 275–278.
43 (68) McCarthy, D. J.; Waud, W. R.; Struck, R. F.; Hill, D. L. Disposition and Metabolism
44 of Aniline in Fischer 344 Rats and C57BL/6 X C3H F1 mice. Cancer Res. 1985, 45
45 (1), 174–180.
46 (69) Miyoshi, N.; Tomita, G. Z. Naturforsch. B 1980, 35 (1), 107.
47
(70) POLYMATH is copyrighted by Shacham, M.; Cutlip, M. B; Elly M. POLYMATH,
48
49 http://www.polymath-software.com/, (accessed Feb 23, 2016)
50 (71) Schweitzer, C.; Schmidt, R. Physical Mechanisms of Generation and Deactivation of
51 Singlet Oxygen. Chem. Rev. 2003, 103 (5), 1685–1758.
52
53
54
55
56
57
58
59
60 16
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 18 of 25

1
2
3
4
5 Table 1: Arrhenius rate parameters fitted in the temperature range of 300-600 K at 1 M for
6 the initial five channels as well as the formation of 4-iminocyclohexa-2,5-dienone (M9); A
7
8 and Asolvation are cm3 molecule-1 s-1 and Ea and Ea-solvation in kJ/mol.
9
10
11
12 Reaction M06 B3LYP
13 A Asolvation Ea Ea-solvation A Ea
14
15 Aniline + 1O2 → M1 3.40 × 10-14 4.70 × 10-14 11.2 -11.0 1.97 × 10-14 89.8
16
17 Aniline + 1O2 → M2 1.56 × 10-14 1.10 × 10-14 31.7 8.59 1.46 × 10-14 111.7
18
19 Aniline + 1O2 → M3 2.06 × 10-14 5.07 × 10-14 16.7 2.09 1.76 × 10-14 84.3
20
Aniline + 1O2 → M4 4.98 × 10-14 1.41 × 10-14 142.3 116.1 6.70 × 10-20 235.6
21
22 Aniline + 1O2 → M5 1.75 × 10-12 3.75 × 10-12 107.3 99.2 1.05 × 10-12 160.4
23
24 M6 → M7 7.66 × 1014 1.83 × 1014 136.8 125.8 1.53 × 1014 115.6
25
26 M7 → M8 2.13 × 1013 4.10 × 1013 108.1 61.8 2.19 × 1013 92.8
27
28 M8 → M9 3.88 × 1013 1.73 × 1013 209.1 198.3 3.42 × 1010 197.1
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
58
59
60 17
ACS Paragon Plus Environment
Page 19 of 25 The Journal of Physical Chemistry

1
2
3
4
5
6
7
0.064
8 [0.045]
9
0.081
10 [-0.069]
0.11
11
[-0.041]
12 0.062
13 [-0.0568]
14
0.125
15 [-0.066]
16
17
18
19
20 Figure 1: Fukui f-1 indices (in bold) and Hirshfeld charges (blue in brackets) on benzene
21
22 (left) and aniline (right).
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
58
59
60 18
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 20 of 25

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
48
49
50
51
52
53
54
55 Figure 2: Initial channels in the reaction of O2 1∆g with aniline assigned with Gibbs free
56
57 energies of activation ∆≠Go298 and reaction ∆rGo298 (in brackets) calculated based on gas-
58
59
60 19
ACS Paragon Plus Environment
Page 21 of 25 The Journal of Physical Chemistry

1
2
3 phase B3LYP/6-311+G(d,p) represented by solid line, M06/6-311+G(d,p) (gas phase, in
4
5 dotted line), and M06/6-311+G(d,p) (solvation, in dashed line). All values are in kJ/mol at
6 298.15 K.
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
58
59
60 20
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 22 of 25

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24 Figure 3: Relative branching rates for the formation of M1 and M3; a: M06/6-311+G(d,p)
25
26 (gas phase), and b: M06/6-311+G(d,p) (Solvation model) represented in dashed line.
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 Figure 4: Gibes free energy potential map for the formation of 4-iminocyclohexa-2,5-
53
54 dienone (para-iminobenzoquinone) from the reaction of singlet oxygen with aniline. a:
55 B3LYP/6-311+G(d,p) (gas phase, in italic), b: M06/6-311+G(d,p) (gas phase, in brackets),
56
57 and c: M06/6-311+G(d,p) (solvation). All values are in kJ/mol at 298.15 K.
58
59
60 21
ACS Paragon Plus Environment
Page 23 of 25 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 Figure 5: Gibes free energy potential map for the formation of 6-iminocyclohexa-2,4-
33
34 dienone (ortho- iminobenzoquinone) product from the reaction of singlet oxygen with
35
36 aniline, a: B3LYP/6-311+G(d,p) (gas phase, in italic), b: M06/6-311+G(d,p) (gas phase, in
37 brackets), and c: M06/6-311+G(d,p) (solvation). All values in kJ/mol at 298.15 K.
38
39
40
41
42
43
44
45
46
47
48
49
50
51 2
52
53
54
55
56
57
58
59
60 22
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 24 of 25

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24 Figure 6: Concentration profiles for the formation of para-iminobenzoquinone via the water-
25
26 solvation model (red line) and the published experimental measurements of Briviba et al.29
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
58
59
60 23
ACS Paragon Plus Environment
Page 25 of 25 The Journal of Physical Chemistry

1
2
3
4
5 TOC graphic
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
58
59
60 24
ACS Paragon Plus Environment

You might also like