Rethinking Aromaticity in H-Bonded Systems. Caveats for Transition

Dec 22, 2014 - The graphical representations of the above functions give rise to two parabolas crossing at the point that determines the energy and po...
2 downloads 12 Views 599KB Size
Subscriber access provided by GAZI UNIV

Article

Rethinking Aromaticity in H-Bonded Systems. Caveats for Transition Structures Involving Hydrogen Transfer and #-Delocalization María Pilar Romero-Fernández, Martin Avalos, Reyes Babiano, Pedro Cintas, Jose L. Jimenez, and Juan C. Palacios J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/jp5113735 • Publication Date (Web): 22 Dec 2014 Downloaded from http://pubs.acs.org on December 29, 2014

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 27

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 56 57 58 59 60

The Journal of Physical Chemistry

Rethinking Aromaticity in H-Bonded Systems. Caveats for Transition Structures Involving Hydrogen Transfer and π-Delocalization† María P. Romero-Fernández,* Martín Ávalos, Reyes Babiano, Pedro Cintas,* José L. Jiménez, and Juan C. Palacios Departamento de Química Orgánica e Inorgánica, QUOREX Research Group, Facultad de Ciencias-UEX, Avda. de Elvas s/n E-06006 Badajoz, Spain

ABSTRACT: Monoaza- and diaza-derivatives of malondialdehydes, in short aminoacroleins and vinamidines, are prototypical examples of open-chain structures prone to π-electron delocalization, for which intramolecular hydrogen bonding enhances (or diminishes) their pseudo-aromaticity depending on the substitution pattern. This interplay is illustrated herein by DFT-based calculations of aromaticity indices in the gas phase and polar solvents. Elucidation of transition structures involved in tautomeric conversions helps to solve how the intramolecular hydrogen transfer occurs. While TSs exhibit a high degree of aromaticity, the dichotomy between forward and backward pathways point to a complex trajectory. Addition of thermal corrections to the electronic energy decreases both the enthalpy and free energy leading to negative ∆H‡ and ∆G‡ values. This variational effect accounts for the otherwise elusive distinction between transition structures and saddle points (usually overlooked for high electronic barriers). Also, this rationale fits well within the framework of Marcus’ theory.

INTRODUCTION

It is unnecessary to underline that the hydrogen bond constitutes a dominant weak interaction that plays a pivotal role in numerous chemical, biochemical, and physical processes, ranging from crystal packing to the stabilization of α-helices or β-sheets in proteins.1-5 Considerable research has also been focused on very strong hydrogen bonds involved in the transition structures of enzymatic processes.6-12 Among all the types of H-bonding interactions, the strongest ones are those assisted by additional effects, such as the so-called resonanceassisted or charge-assisted hydrogen bonds.13 In a nutshell, given a system H-X-(CH=CH)nCH=Y, the intramolecular resonance-assisted hydrogen-bond (RAHB),14,15 is identified as the relationship between the hydrogen bond strength (measured by the X⋅⋅⋅⋅Y distance, the X-H bond stretching frequency, or the X-H proton chemical shift) and the π-delocalization of the

ACS Paragon Plus Environment

1

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 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 27

conjugated chain connecting the hydrogen bond donor (X) and acceptor (Y) atoms (measured by delocalization parameters).16 Moreover, the RAHB can be interpreted in terms of the minimum difference between X and Y proton affinities [∆PA = PA(X–) – PA(Y)] or, alternatively the minimum difference between their pKa values [∆pKa = pKa (X-H) – pKa (H-Y+)] as measured in a specific solvent.17,18 In any case, the RAHB concept is controversial as some authors consider that stabilization is largely provided by the σ-skeleton, rather than π-effects, even if some correlation between the delocalized π-electron structure and RAHB parameters has also been found.19-26 This context brings us to mind the fact that H-bonding interactions that increase cyclic πelectron delocalization also enhances both the aromaticity of such H-bonded compounds and their association energy. Conversely, H-bonding leading to greater π-electron localization decreases the aromatic character and association energies.27 In a recent study, Wu et al. have also computed H-bonding interaction energies (at the PBE0/6-311++G(3df,3pd) level without zeropoint energy corrections) and dissected nucleus-independent chemical shifts (NICS) at the PW91/Def2-TZVPP level, both accounting for the enhancement (or reduction) of aromaticity.28 Along with inherently aromatic compounds, the concept of aromaticity can be extended to pseudo-aromatic rings (or quasi-aromatic as often denoted), for which π-electron delocalization mediated by hydrogen bonding can be evaluated through indices based on geometrical considerations.16 Recently, we have studied in detail the tautomeric and conformational equilibria of mono- [Ar1-NH-CH=CAr2-CH=O] and di-azaderivatives [Ar1-NHCH=CAr2-CH=N-Ar1] of malondialdehydes, namely acrolein and vinamidines structures with extended conjugation.29 The tautomeric preferences for these substances could be rationalized by dissecting and quantifying contributions arising from structural fragments and intramolecular interactions (at the M06-2X/6-311++G(d,p) level and including solvent effects via the SMD model). The major stabilizing effect responsible for a favored tautomer or conformer is provided by intramolecular hydrogen bonding (O-H····N and N-H····N bonds), which contributes to further electron delocalization. Our theoretical modeling showed that H-bonded structures were largely favored in CDCl3, while fully extended (E-configured, s-trans conformers) enamine derivatives became prevalent as solvent polarity increased (DMSO or ethanol). The latter fully agreed with experimental observations of the corresponding tautomeric and/or conformational equilibria as inferred from NMR monitoring in DMSO-d6.29

ACS Paragon Plus Environment

2

Page 3 of 27

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 56 57 58 59 60

The Journal of Physical Chemistry

We now present a further computational estimation of the intramolecular hydrogen bond that characterizes the pseudo-heterocyclic structure of 2-arylmalondialdehydes and their monoand di-azaderivatives. The delocalization degree in both the gas phase and standard polar solvents (chloroform, dimethyl sulfoxide, and ethanol) for tautomers a and b of 2-aryl-3methyl(aryl)aminoacroleins (1-9), as well as the chelated forms of 2-arylmalondialdehydes 1012, and 2-arylvinamidines 13-21 (Figure 1) have been thoroughly quantified by geometry-based aromaticity indices.

O

H

N

R2

d1

d4

H

H

d2

1 2 3 4 5 6 7 8 9

d3

O

H

H

N

R2 H

R1

R1

a

b

O

H

H

R2

O H

10 R1 = H 11 R1 = OCH3 12 R1 = NO2

H

H

R1

R1 = H; R2 = CH3 R1 = CH3O; R2 = CH3 R1 = NO2; R2 = CH3 R1 = CH3O; R2 = C6H5 R1 = NO2; R2 = C6H5 R1 = CH3O; R2 = 4-CH3OC6H4 R1 = CH3O; R2 = 4-NO2C6H4 R1 = NO2; R2 = 4-CH3OC6H4 R1 = NO2; R2 = 4-NO2C6H4

N

N

R2 H

R1

13 14 15 16 17 18 19 20 21

R1 = H; R2 = CH3 R1 = OCH3; R2 = CH3 R1 = NO2; R2 = CH3 R1 = CH3O; R2 = C6H5 R1 = NO2; R2 = C6H5 R1 = CH3O; R2 = 4-CH3OC6H4 R1 = CH3O; R2 = 4-NO2C6H4 R1 = NO2; R2 = 4-CH3OC6H4 R1 = NO2; R2 = 4-NO2C6H4

Figure 1

Moreover, this study focuses on the location of transition structures accounting for the hydrogen transfer process. In doing so, some pitfalls and inconsistencies emerge, as documented in previous literature. The apparent dead end has now been satisfactorily settled by detours involving an analysis of variational effects with inclusion of zero-point energy (ZPE) corrections.

ACS Paragon Plus Environment

3

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 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 27

RESULTS AND DISCUSSION

Computational Details. The computational DFT study has been performed using the M06-2X30 hybrid density functional in conjunction with the 6-311++G(d,p) basis set31-35 as implemented in the Gaussian09 package.36 The M06-2X method was chosen on the basis of previous studies showing its accuracy in estimating conformational energies related to noncovalent interactions;37 while energy re-assessments using other functionals and basis sets lie beyond the scope of this work. In all cases, frequency calculations were also carried out to confirm the existence of true stationary points on the potential energy surface. All thermal corrections were calculated at the standard values of 1 atm at 298.15K. Solvent effects were modeled through the method of density-based, self-consistent reaction field (SCRF) theory of bulk electrostatics, namely, the solvation model density (SMD) method38-42 as implemented in the Gaussian09 suite of programs. This solvation method accounts for long-range electrostatic polarization (bulk solvent)43 as well as for short-range effects associated with cavitation, dispersion, and solvent structural effects.44 Pseudo-aromaticity and Molecular Geometry Indices. As mentioned above, the aromatic character of planar pseudo-heterocycles can adequately be interrogated through a series of indices developed since the early 1970s, all possessing pros and cons, although as a whole they provide a valuable description of cyclic π-electron delocalization. For structures 1-21, Table 1 displays the values of the sum of the bond orders differences, Σ∆N,45 the Bird’s aromaticity index, I,46-48 the Kotelevskii and Prezhdo modification, APoz,KP,49 of the Pozharskii structural index, APoz,50,51 the Gilli parameters, Q and λ,52,53 the harmonic oscillator model of aromaticity, HOMA,54,55 the harmonic oscillator model of electron delocalization, HOMED,56-58 and the new HOMA index parametrization for π-electron heterocycles, HOMHED.59 A complete description of the above-mentioned indices together with specific parameters employed for the calculation of HOMA, HOMED, and HOMHED indices have been included with the SI (Appendix 1).

ACS Paragon Plus Environment

4

Page 5 of 27

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 56 57 58 59 60

The Journal of Physical Chemistry

Table 1. Aromaticity indices: Σ∆N, I, APoz,KP, Q, λ, HOMA, HOMED, and HOMHED, calculated at the M06-2X/6-311++G(d,p) level for structures 1-21 in the gas-phase. Entry 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

Structure 1a 2a 3a 4a 5a 6a 7a 8a 9a 1b 2b 3b 4b 5b 6b 7b 8b 9b 10 11 12 13 14 15 16 17 18 19 20 21

Σ∆N 55.77 54.94 57.84 44.27 48.64 48.68 35.59 52.31 39.40 23.40 21.62 25.33 31.05 33.20 29.49 35.10 31.27 35.69 26.01 25.16 26.60 29.41 27.05 33.84 29.34 33.71 31.64 25.71 37.02 29.73

I 67.08 66.96 66.05 61.99 62.88 64.30 57.27 64.31 58.45 48.03 46.84 48.70 53.30 54.29 52.32 56.03 53.16 56.11 52.51 52.08 52.55 44.50 43.87 44.57 47.66 48.10 49.09 46.30 49.60 47.29

APoz,KP 64.64 63.98 66.34 55.11 58.74 58.68 47.82 61.70 51.07 37.90 36.45 39.64 43.84 45.78 42.59 47.16 44.10 47.85 37.16 36.45 37.74 48.15 46.37 51.46 47.64 51.02 49.37 44.77 53.49 47.98

Q 0.174 0.176 0.169 0.204 0.193 0.193 0.226 0.184 0.216 -0.134 -0.139 -0.128 -0.115 -0.108 -0.119 -0.105 -0.113 -0.102 0.180 0.182 0.178 0.139 0.144 0.130 0.140 0.131 0.135 0.148 0.124 0.139

λ 47.27 46.67 48.79 38.18 41.52 41.52 31.52 44.24 34.55 44.17 42.08 46.67 52.08 55.00 50.42 56.25 52.92 57.50 43.75 43.13 44.38 44.40 42.40 48.00 44.00 47.60 46.00 40.80 50.40 44.40

HOMA 0.93 0.93 0.92 0.89 0.90 0.91 0.85 0.91 0.86 0.57 0.55 0.60 0.62 0.66 0.60 0.64 0.64 0.67 0.65 0.64 0.67 0.85 0.84 0.85 0.86 0.87 0.87 0.86 0.88 0.86

HOMED 0.89 0.89 0.88 0.85 0.86 0.87 0.81 0.87 0.82 0.80 0.79 0.80 0.84 0.85 0.83 0.86 0.84 0.86 0.82 0.82 0.82 0.86 0.85 0.86 0.87 0.88 0.88 0.87 0.88 0.87

HOMHED 0.88 0.88 0.87 0.85 0.85 0.86 0.82 0.86 0.82 0.79 0.78 0.80 0.83 0.84 0.83 0.85 0.83 0.85 0.82 0.82 0.82 0.84 0.84 0.84 0.87 0.86 0.87 0.86 0.87 0.86

Some conclusions can be extracted from data gathered in Table 1. Thus, a) the HOMED and HOMHED indices underestimate the differences in pseudo-cyclic electron delocalization showing in all cases values between 0.80 and 0.89, which shadow the tautomeric structural features. The substituent effect is also underestimated by the HOMA index within each structural family. b) Parameters based on bond orders (Σ∆N, I, and APoz,KP,) as well as the HOMA index do not afford clear-cut aromaticity relationships. While 3-aminoacroleins (structures 1a-9a) exhibit a greater pseudo-aromaticity than other series, small variations and divergences are observed for vinamidines (structures 13-21), malondialdehydes (structures 10-11), or iminoenols (structures 1b-9b). However, parameters based on bond lengths, such as Q and, as a result λ, lead to other

ACS Paragon Plus Environment

5

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 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 27

conclusive statement. Thus, c) within each of these structural families, the π-electron delocalization is slightly enhanced as the electron-withdrawing character of R1 increases (see for example, entries 4 and 5, 6 and 8, and 7 and 9). d) Disparate results are instead obtained for the effect caused by R2 substituents, which varies in each group of structures. While for 3aminoacroleins (1a-9a) electron delocalization decreases as the electron-withdrawing character of R2 increases (compare entries 2, 4, 6, and 7 or entries 3, 5, 8, and 9), for iminoenols 1b-9b the π-delocalization increases as the electron-withdrawing character of R2 increases too (compare entries 11, 13, 15, and 16 or entries 12, 14, 17, and 18). For vinamidines 13-21, electron delocalization decreases as the electron-withdrawing character of R2 increases, regardless of the electronic effect of R1. The scarce influence observed for substituents on the pseudo-aromaticity of aminoacroleins and vinamidines can reasonably be ascribed to the lack of coplanarity between the conjugated unit and its aromatic substituents (Figure 2). Thus, for aminoacroleins (1a-9a) the dihedral angle ϕ varies from 34º to 48º while the dihedral angle ψ ranges from 3º to 31º, with lower values measured in structures where R2 is an electron-withdrawing group (Table S1). On the contrary, enolimine tautomers 1b-9b show similar ϕ and ψ angles (35º-46º), irrespective of the electronic nature of R1 and R2 groups. For vinamidines 13-21, the calculated dihedral angles ϕ, ψ, and ω show the following variations: 30º < ϕ < 47º, 1º < ψ < 29º, and 35º < ϕ < 43º (Tabla S2). Once again, the presence of electron-withdrawing groups at R2 gives rise to lower ψ angles. Figure 2 depicts optimized structures for aminoacrolein 9a and vinamidine 21, showing for the sake of clarity the above-mentioned dihedral angles.

O

H

N

H

R2

R2

ψ

ω

H

N

H

H

H

φ

R1

R2 N ψ

φ

R1

Figure 2. Optimized geometries at the M06-2X/6-311++G(d,p) level for aminoacrolein 9a (ϕ = 36.7º y ψ = 16.5º) and vinamidine 21 (ϕ = 36.4º, ψ = 1.6º, y ω = 37.4º) in ethanol.

ACS Paragon Plus Environment

6

Page 7 of 27

It has been demonstrated that on decreasing the X····X(Y) distance, the intramolecular hydrogen bonding becomes stronger.52 Table S3 collects all distances calculated for the fragment X····H-X(Y) in compounds 1-21, both in the gas-phase and polar solvents: chloroform, dimethyl sulfoxide, and ethanol. Such data show that regardless of the medium, the distances X····H and X····X(Y) decrease as the electron-withdrawing character of R1 increases. As expected, the effect of R2 is noteworthy in H-bonded structures 1b-9b, for which the intramolecular hydrogen bonds weaken in all media when R2 = 4-NO2C6H4. The influence of these three solvents on the π-delocalization has further been assessed by optimizing structures 1-21 at the M06-2X/6-311++G(d,p) level with the SMD model,38-42 and then re-calculating the parameters collected in Table 1 from the new bond lengths d1-d4. These results recorded in Tables S4-S6 (Supporting Information) show that solvent replacement does not result in a significant alteration of the delocalization indices, a fact particularly noticeable when comparing the results obtained in liquids of different solvation abilities, i.e. chloroform (ε = 4.71) and dimethyl sulfoxide (ε = 46.83). For instance, Figure 3 shows a comparative picture of the sum of the bond order differences (Σ∆N) for structures 1a-9a in chloroform, dimethyl sulfoxide, and ethanol. Data collected in Table S7 enable an additional inspection of the pseudoaromatic stabilization linked to solvent effects. 75.00 70.00 65.00 Σ∆ Σ∆N

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 56 57 58 59 60

The Journal of Physical Chemistry

60.00 chloroform 55.00

dmso

50.00

ethanol

45.00 40.00 1a

2a

3a

4a

5a

6a

7a

8a

9a

Structure

Figure 3. Sum of the bond order differences (Σ∆N) for structures 1a-9a in chloroform, dimethyl sulfoxide, and ethanol.

ACS Paragon Plus Environment

7

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 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 27

Transition Structures. The location and identification of transition structures (TSs) corresponding to the intramolecular hydrogen transfer provide further insights into the πdelocalization and pseudo-aromaticity of these chelated open-chain molecules. Tables 2-4 show the relative electronic energies, enthalpies, and free energies in the gas phase and polar solvents (chloroform, dimethyl sulfoxide, and ethanol as above) for transition structures accounting for the interconversion of tautomers a and b in 2-aryl-3-methylaminoacroleins 1-9, as well as the interconversion of two equivalent forms in malondialdehydes 10-11 and vinamidines 13-21 (Figure 4). Furthermore, Table S8 lists X····H-X(Y) distances involved in the hydrogen transfer, whilst Table 5 displays the values of Σ∆N, I, APoz,KP, Q, λ, HOMA, HOMED, and HOMHED, calculated at the M06-2X/6-311++G(d,p) level for transition structures TS1a/1b-TS9a/9b and TS1021

in the gas phase, all serving to characterize the delocalization degree. Data obtained for the

same transition structures in chloroform, dimethyl sulfoxide and ethanol have been included in the SI (see Tables S9-S11).

O

H

H

R2 H

R1 TS1a/1b TS2a/2b TS3a/3b TS4a/4b TS5a/5b TS6a/6b TS7a/7b TS8a/8b TS9a/9b

N

O

H

H

R2 O

H

H

H

N

R2 H

R1

R1

TS10 R1 = H R1 = H; R2 = CH3 1 2 TS11 R1 = OCH3 R = CH3O; R = CH3 TS12 R1 = NO2 R1 = NO2; R2 = CH3 1 2 R = CH3O; R = C6H5 R1 = NO2; R2 = C6H5 R1 = CH3O; R2 = 4-CH3OC6H4 R1 = CH3O; R2 = 4-NO2C6H4 R1 = NO2; R2 = 4-CH3OC6H4 R1 = NO2; R2 = 4-NO2C6H4

N

TS13 TS14 TS15 TS16 TS17 TS18 TS19 TS20 TS21

R1 = H; R2 = CH3 R1 = OCH3; R2 = CH3 R1 = NO2; R2 = CH3 R1 = CH3O; R2 = C6H5 R1 = NO2; R2 = C6H5 R1 = CH3O; R2 = 4-CH3OC6H4 R1 = CH3O; R2 = 4-NO2C6H4 R1 = NO2; R2 = 4-CH3OC6H4 R1 = NO2; R2 = 4-NO2C6H4

Figure 4

ACS Paragon Plus Environment

8

Page 9 of 27

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 56 57 58 59 60

The Journal of Physical Chemistry

Table 2. Relative electronic energies (∆E), entalphies (∆H), and free energies (∆G) (kcal/mol) for transition structures TS1a/1b-TS9a/9b relative to the most stable isomer (1a-9a). Gas phase Struct. TS1a/1b TS2a/2b TS3a/3b TS4a/4b TS5a/5b TS6a/6b TS7a/7b TS8a/8b TS9a/9b

∆E 7.48 7.52 7.60 7.70 7.46 7.26 8.11 7.12 7.56





Chloroform ‡

∆H 4.43 4.35 4.45 4.78 4.64 4.20 5.14 3.95 4.48

∆G 5.69 5.10 5.37 5.77 5.84 4.82 6.10 4.59 5.54

∆E 8.73 8.70 9.03 8.58 8.45 7.97 9.10 7.99 8.53





Dimethyl sulfoxide ‡

∆H 5.68 5.47 5.79 5.60 5.58 4.82 6.03 4.75 5.14

∆G 6.86 6.06 6.35 6.37 7.06 4.90 6.74 5.55 6.00

∆E 9.14 9.12 9.56 8.87 8.80 8.33 9.55 8.38 9.19





Ethanol ‡

∆H 5.93 5.89 6.24 5.60 5.75 5.43 6.30 5.18 5.96

∆G 7.19 6.56 6.67 6.37 7.69 6.33 6.85 5.99 6.66

∆E 8.98 9.71 10.17 9.29 9.19 8.72 9.88 8.84 9.40







∆H 6.03 6.43 6.82 6.07 5.87 5.70 6.74 5.71 6.08

∆G 7.32 7.04 7.12 6.67 7.18 6.57 7.47 7.53 7.23

Table 3. Relative electronic energies (∆E), entalphies (∆H), and free energies (∆G) (kcal/mol) for transition structures TS1a/1b-TS9a/9b relative to the less stable isomer (1b-9b). Gas phase Struct. TS1a/1b TS2a/2b TS3a/3b TS4a/4b TS5a/5b TS6a/6b TS7a/7b TS8a/8b TS9a/9b

∆E 2.37 2.54 1.95 2.50 2.01 2.57 2.71 2.00 2.34





Chloroform ‡

∆H -0.31 -0.23 -0.84 -0.26 -0.70 -0.14 -0.01 -0.54 -0.47

∆G 0.06 -0.12 -0.51 0.33 -0.06 0.53 0.47 -0.05 0.13





∆E 1.51 1.70 1.08 2.05 1.51 1.97 2.01 1.39 1.62

∆H -0.91 -1.02 -1.55 -0.77 -1.25 -0.50 -0.54 -1.00 -0.98

Dimethyl sulfoxide ‡

∆G 0.17 -0.95 -1.32 -0.56 -0.50 -0.33 -0.37 -0.82 -0.84





∆E 1.22 1.46 0.85 1.84 1.28 1.86 1.71 1.18 1.27

Ethanol ‡

∆H -1.36 -1.27 -1.79 -1.19 -1.69 -0.50 -0.85 -1.15 -1.21

∆G -0.46 -1.07 -1.71 -1.14 -1.23 0.28 -0.28 -0.39 -0.55





∆E 0.36 1.30 0.78 1.79 1.28 1.68 1.78 1.16 1.37



∆H -1.90 -1.33 -1.83 -1.14 -1.69 -0.62 -0.74 -1.17 -1.19

∆G -0.87 -1.10 -1.95 -0.96 -1.56 0.25 -0.38 -0.56 -0.56

Table 4. Relative electronic energies (∆E), entalphies (∆H), and free energies (∆G) (kcal/mol) for transition structures TS10-21 relative to their corresponding tautomers (10-21). Gas phase Structure TS10 TS11 TS12 TS13 TS14 TS15 TS16 TS17 TS18 TS19 TS20 TS21

∆E 3.14 3.21 3.00 6.98 7.11 6.54 6.89 6.41 6.60 7.20 6.24 6.73





∆H 0.37 0.40 0.26 3.68 4.14 3.08 3.63 2.92 3.48 4.27 2.79 3.05

Chloroform ‡

∆G 0.97 0.87 0.88 3.94 4.98 3.61 4.70 3.28 4.72 5.45 3.03 3.81

∆E 2.93 3.00 2.80 7.47 7.60 6.86 7.24 6.71 6.90 7.45 6.35 7.01





∆H 0.21 0.26 0.00 3.98 4.44 3.43 3.74 3.17 3.64 3.90 3.10 3.92

Dimethyl sulfoxide ‡

∆G 0.88 1.06 0.42 3.94 4.82 4.05 3.95 2.80 4.35 4.09 3.20 4.51

∆E 2.75 2.79 2.57 7.65 7.79 6.99 7.46 6.86 7.13 7.30 6.62 7.01





∆H 0.05 0.15 -0.20 4.00 4.54 3.70 3.93 3.53 3.59 3.94 3.20 3.90

Ethanol ‡

∆G 0.65 1.09 0.52 3.89 4.91 4.01 4.80 3.99 4.08 5.84 3.24 5.68

∆E 3.19 3.28 3.02 7.81 7.96 7.04 7.29 6.90 7.21 7.71 6.59 7.26





∆H 0.43 0.45 0.20 4.25 4.67 3.78 3.87 3.44 3.95 4.41 3.16 4.11



∆G 1.10 0.94 1.23 4.55 4.91 4.23 4.49 3.46 5.19 5.92 3.09 5.78

ACS Paragon Plus Environment

9

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 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 27

The computational results gathered in Table 4 suggest that intramolecular hydrogen transfer in 2-arylmalondialdehydes 10-12 is faster (∆G‡ < 1.25 kcal/mol) than for 2arylvinamidines 13-21 (3.0 kcal/mol < ∆G‡ < 6.0 kcal/mol), whose barriers are somewhat lower than those found for the conversion of 3-alkyl(aryl)-2-arylaminoacroleins 1a-9a into their tautomers 1b-9b (4.5 kcal/mol < ∆G‡ < 7.7 kcal/mol) (Table 2). However, like in other intramolecular hydrogen transfers occurring on unsymmetrical systems,60 the energy barriers (∆H‡ and ∆G‡) found for the opposite process, i.e., for the conversion of 1b-9b into 1a-9a, are < 0 kcal/mol (Table 3), which deserves an explanation (see below). The computed distances between all atoms involved in the intramolecular hydrogen transfer for compounds 1-21 (Table S8) not only unravel the transfer evolution, which takes place primarily through the N···H distance leading to an increase in symmetry for transition stuctures TS1a/1b-TS9a/9b and TS10-TS21, but also a shortening of the distance X···X(Y) by 0.20.3 Å with respect to the corresponding tautomers. The latter results in an increase of the πdelocalization along the unsaturated spacer. In order to confirm these conclusions, Tables 5 and S9-S11 (Supporting Information) show the Σ∆N, I, APoz,KP, Q, λ, HOMA, HOMED, and HOMHED indices, calculated at the aforementioned M06-2X/6-311++G(d,p) level for transition structures TS1a/1b-TS9a/9b and TS10-21 in the gas phase and polar solvents. An analysis of such data evidence that all parameters show an increase of π-electron delocalization in the transition structures relative to the parent tautomers (see Tables 1 and S4-S6).

ACS Paragon Plus Environment

10

Page 11 of 27

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 56 57 58 59 60

The Journal of Physical Chemistry

Table 5. Aromaticity indices: Σ∆N, I, APoz,KP, Q, λ, HOMA, HOMED, and HOMHED, calculated at the M06-2X/6-311++G(d,p) level for structures TS1a/1b-TS9a/9b and TS10-21 in the gas phase Entry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21

Structure TS1a/1b TS2a/2b TS3a/3b TS4a/4b TS5a/5b TS6a/6b TS7a/7b TS8a/8b TS9a/9b TS10 TS11 TS12 TS13 TS14 TS15 TS16 TS17 TS18 TS19 TS20 TS21

Σ∆N 67.76 68.31 66.44 79.10 75.90 76.26 87.96 73.36 84.48 93.19 94.02 90.45 70.55 71.89 65.15 78.18 72.82 78.18 81.92 71.49 76.84

I 73.00 73.59 70.72 80.84 77.17 79.01 85.52 75.78 81.69 93.51 94.31 90.92 70.00 71.36 64.55 77.56 72.08 77.56 81.67 70.71 76.19

APoz,KP 74.06 74.48 72.99 83.10 80.54 80.79 89.86 78.46 85.84 94.23 94.94 91.92 78.40 79.38 74.47 83.84 79.89 83.84 86.58 78.91 82.85

Q -0.019 -0.018 -0.022 0.009 0.002 0.002 0.031 -0.004 0.023 0.001 0.000 0.001 0.000 0.000 0.000 0.000 0.000 0.000 -0.002 0.000 0.000

λ 94.24 94.55 93.33 97.27 99.39 99.39 90.61 98.79 93.03 99.69 100 99.69 100 100 100 100 100 100 99.20 100 100

HOMA 0.88 0.88 0.88 0.92 0.91 0.91 0.94 0.91 0.94 0.93 0.93 0.94 0.97 0.97 0.95 0.98 0.97 0.98 0.99 0.97 0.98

HOMED 0.91 0.91 0.90 0.96 0.95 0.96 0.98 0.95 0.97 0.99 0.99 0.99 0.96 0.96 0.93 0.98 0.97 0.98 0.99 0.96 0.98

HOMHED 0.90 0.90 0.88 0.95 0.94 0.94 0.97 0.93 0.96 0.99 0.99 0.98 0.93 0.94 0.90 0.97 0.94 0.96 0.97 0.94 0.96

A plot of the electronic energy profiles extracted from Tables 2 and 3, as shown in Figure 5, reveals the lack of symmetry in the tautomerization of 3-alkylaminoacroleins; the energy barrier that separates the transition structure from tautomers a and b is greater than 7 kcal/mol and less than 3 kcal/mol, respectively. This accounts for the distinctive bond lengths calculated at the saddle point for the O-H (< 1.20 Å) and N-H (> 1.27 Å) bonds. Figure 6 shows a similar picture for the intramolecular hydrogen transfer between the tautomers a and b of 3-alkyl(aryl)-2-arylaminoacroleins 1-9 based on the ∆G‡ values calculated (Tables 2 and 3), which might suggest a more complex reaction path involving at least another relative local minimum and an additional saddle point sandwiched between the transition structure drawn and tautomer b.

ACS Paragon Plus Environment

11

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 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 27

Figure 5. Schematic diagram showing the different stability of tautomers a and b for 2-aryl-3methylaminoacroleins 1-9.

Figure 6. Schematic diagram showing the relative free energy of tautomers a and b for 2-aryl-3methylaminoacroleins 1-9.

ACS Paragon Plus Environment

12

Page 13 of 27

In the search for a plausible rationale, we computed the internal reaction coordinate (IRC) for TS3a/3b in ethanol (∆E‡ = 0.78, ∆H‡ = –1.83, ∆G‡ = –1.95 kcal/mol), the transition structure corresponding to the tautomerization of one 3-alkyl-2-arylaminoacrolein derivative. Figure 7 shows the electronic energy variation along the reaction pathway as plotted for the equilibrium between 3a and 3b. The IRC calculation indicated that TS3a/3b is actually the true saddle point for that prototropy characterized by both a low energy barrier (0.78 kcal/mol) and a high imaginary frecuency (–933.62 cm-1), which cause a sharp drop in the zero point energy (ZPE) as shown in Figure 8, constructed from thermochemical data of each structure in the reaction path.

0.3 0.2 Energy (kcal/mol)

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 56 57 58 59 60

The Journal of Physical Chemistry

0.0 -0.2 -0.3 -0.5 -0.6 -0.3

-0.2

-0.1

0.0

0.1

0.2

0.3

Reaction coordinate (Bohr)

Figure 7. Tautomeric equilibrium between 3a and 3b in ethanol. The curve shows the variation of the electronic energy (kcal/mol) along the reaction pathway.

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry

0.5 0.0 Energy (kcal/mol)

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 56 57 58 59 60

Page 14 of 27

-0.5 -1.0 E

-1.5

ZPE -2.0 H -2.5 -3.0 -0.15

G -0.10

-0.05

0.00

0.05

0.10

0.15

Reaction coordinate (Bohr)

Figure 8. Variations of electronic energy (E), zero point energy (ZPE), as well as enthalpy (H298.15) and free energy (G298.15) calculated at 298.15 K (in kcal/mol) near the saddle point.

Addition of thermal corrections, including the ZPE correction, to the electronic energy does justify a marked decrease in the internal energy and hence, of the enthalpy and free energy at the saddle point, thus leading to negative values for ∆H‡ and ∆G‡. This fact shifts the maximum values of ∆H‡ and ∆G‡ (variational effect) with respect to the saddle point, which is not particularly evident in the presence of high electronic energy barriers. Similar results are obtained when this methodology is applied to the tautomeric interconversion between forms a and b of another 3-arylaminoacrolein such as 4. The internal reaction coordinate (IRC) for TS4a/4b in chloroform (∆E‡ = 2.05, ∆H‡ = –0.77, ∆G‡ = –0.56 kcal/mol) shows that TS4a/4b is the real saddle point for the tautomerization reaction and exhibits a low energy barrier (2.05 kcal/mol) plus an imaginary frecuency (–466.83 cm-1), which cause a drop in the zero point energy (ZPE) (Figures 9 and 10). A variational effect can newly be invoked as the inclusion of thermal corrections, including the ZPE correction, to the electronic energy decreases the internal energy and accordingly both the enthalpy and free energy at the saddle point.

ACS Paragon Plus Environment

14

Page 15 of 27

3.0 2.0 Energy (kcal/mol)

1.0 0.0 -1.0 -2.0 -3.0 -4.0 -5.0 -6.0 -2.5

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

2.5

Reaction coordinate (Bohr)

Figure 9. Variation of the electronic energy (kcal/mol) along the reaction coordinate for the tautomeric equilibrium between 4a and 4b in chloroform. 3.0 1.5 Energy (kcal/mol)

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 56 57 58 59 60

The Journal of Physical Chemistry

0.0

E ZPE

-1.5

H -3.0

G

-4.5 -6.0 -1.5

-1

-0.5

0

0.5

1

1.5

Reaction coordinate (Bohr)

Figure 10. Variations of electronic energy (E), zero point energy (ZPE), and enthalpy (H298.15) and free energy (G298.15) near the saddle point.

Finally, it should be emphasized that if the intramolecular hydrogen transfer occurs on a symmetrical substrate, i.e., malondialdehydes like 10-12 or vinamidines like 13-21, the calculate barriers (∆E‡, ∆H‡ and ∆G‡) are low, yet positive (see Table 4), whereas for substrates lacking symmetry such as aminoacroleins 1a-9a, their transformation into the corresponding tautomers 1b-9b, requires more energy and tends to make equal the reaction energy (∆Eº, ∆Hº, and ∆Gº ) to

ACS Paragon Plus Environment

15

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 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 27

the activation energy (∆E‡, ∆H‡ and ∆G‡). This thermodynamic contribution to the reaction barrier, also responsible for the negative values of ∆H‡ and ∆G‡ as found for the inverse transformation (1b-9b → 1a-9a), does not exist when ∆Eº, ∆Hº and ∆Gº are equal to 0. The necessary correction could then be formulated in terms of the Marcus theory,61 by expressing the reaction barrier for proton transfers, ∆X‡, as function of ∆Xº and ∆‡ , the latter being the value of ∆X‡ when ∆Xº = 0, or “intrinsic kinetic barrier” (X = E, H or G).



∆ =

∆‡

∆  ∆  + + 2 16∆‡

Given the different nature of the atoms involved in proton transfer (N and O), the set of ‡ ) were calculated as the arithmetic mean of the ∆‡ data for aminoacroleins (∆, ‡ ‡ corresponding ∆‡ values for malondialdehydes (∆, ) and vinamidines (∆, ).

Accordingly, the preceding expression should be written as follows: ∆  1 ∆  ‡ ‡ ∆ = ∆, + ∆, + + ‡ ‡ 2 2 8 ∆, + ∆,



Tables 6 and 7 depict the new values for ∆E‡, ∆H‡, and ∆G‡ calculated for the transformation 1a-9a ⇄ 1b-9b by applying the Marcus theory, and Figures 11 and 12 show, as representative examples, the variations in electronic energy and free energy along the reaction pathway for the intramolecular proton transfer from the iminoenol donor 3b to the acceptor molecule, aminoacrolein 3a, in ethanol. In these plots, the reaction coordinate (RC) is defined as the difference between the distances of the O-H and N-H bonds (RC = dO-H – dN-H). Then, the total reaction pathway length (d) can be estimated as d = RC3a – RC3b, while the introduction of the relative reaction coordinate, r = RC – RC3b, allow us to obtain r values between 0 and d.62 In the framework of Marcus’ theory, the reactant and the product of the hydrogen transfer process are regarded as simple harmonic oscillators with identical force constant, k, though shifted by the pathway length (d), and whose energies can be calculated according to:

ACS Paragon Plus Environment

16

Page 17 of 27

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 56 57 58 59 60

The Journal of Physical Chemistry

1  2 1 = ∆  +     2  =



being ∆  the reaction energy and the force constant of the oscillator (k) calculated as: 8∆‡ =  The graphical representations of the above functions give rise to two parabolas crossing at the point that determines the energy and position of the transition structure along the reaction coordinate (Figures 11 and 12).

Figure 11. Marcus plot of the electronic energy data for the N---H-O → N-H---O proton transfer for 3b ⇄ 3a calculated at the M06-2X/6-311++G(d,p) level versus the relative reaction coordinate (r). [k = 17.07 kcal mol-1 Å-2; d = 1.536 Å; ∆Er = -9.39 kcal mol-1; ∆E‡N---H-O → N--H--O = 1.43 kcal mol-1; ∆E‡N-H---O → N--H--O = 10.83 kcal mol-1]

ACS Paragon Plus Environment

17

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 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 27

Figure 12. Marcus plot of the free energy data for the N---H-O → N-H---O proton transfer for 3b ⇄ 3a calculated at M06-2X/6-311++G(d,p) level versus the relative reaction coordinate (r). [k = 9.25 kcal mol-1 Å-2; d = 1.536 Å; ∆Gr = -9.08 kcal mol-1; ∆G‡N---H-O → N--H--O = 0.08 kcal mol-1; ∆G‡N-H---O → N--H--O = 9.15 kcal mol-1]

Data shown in Tables 6 and 7 indicate that the hypothetical equilibrium is completely shifted to the left, being 3-alkyl(aryl)-2-arylaminoacroleins (1a-9a) the only species present in solution, particularly in polar solvents, as it has been experimentally confirmed through 1H and 13

C NMR experiments.29 It should finally be pointed out in context that correlations between

aromatic stabilization and reactivity in some pericyclic transformations have been discussed in the light of Marcus theory.63,64 Table 6. Relative electronic energies (∆E‡), entalphies (∆H‡), and free energies (∆G‡) (kcal/mol) for transition structures TS1a/1b-TS9a/9b relative to the most stable isomer (1a-9a) obtained by application of the Marcus equation. Gas phase Struct. TS1a/1b TS2a/2b TS3a/3b TS4a/4b TS5a/5b TS6a/6b TS7a/7b TS8a/8b TS9a/9b

∆E 7.94 7.95 8.02 7.99 7.83 7.53 8.26 7.54 7.82





∆H 5.08 5.14 5.37 5.33 5.38 4.72 5.62 4.60 5.06

Chloroform ‡

∆G 6.07 6.11 6.14 6.17 6.08 5.35 6.60 4.97 5.83

∆E 9.44 9.37 9.63 8.91 8.86 8.41 9.37 8.46 8.97





∆H 6.69 7.28 7.34 6.46 6.84 5.52 6.66 5.75 6.22

Dimethyl sulfoxide ‡

∆G 6.92 7.38 7.71 7.17 7.60 5.95 7.35 6.40 7.07

∆E 9.91 9.81 10.13 9.17 9.22 8.72 9.73 8.90 9.57





∆H 7.92 7.29 8.06 6.96 7.46 6.01 7.18 6.34 7.17

Ethanol ‡

∆G 7.39 8.03 8.39 7.81 8.92 6.50 7.95 6.43 7.75

∆E 10.66 10.61 10.83 9.77 9.70 9.36 10.29 9.42 9.94





∆H 7.99 7.91 8.6 7.28 7.56 6.50 7.61 6.88 7.32



∆G 8.41 8.41 9.15 7.95 8.75 7.04 8.47 8.10 8.48

ACS Paragon Plus Environment

18

Page 19 of 27

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 56 57 58 59 60

The Journal of Physical Chemistry

Tables S12 and S13 show the dihedral angles ϕ, ψ, and ω computed for transition structures TS1a/1b-TS9a/9b and TS10-TS21. Like in their corresponding tautomers, the ψ angle decreases as the electron-withdrawing character of the R2 substituents increases, although a narrower variation could be observed (14º < ψ < 35º in TS1a/1b-TS9a/9b and 21º < ψ < 32º for TS10-TS21). Likewise, the values for ϕ oscillate between 32º and 45º for structures TS1a/1b-TS9a/9b and between 25º and 43º for TS10-TS21. Table 7. Relative electronic energies (∆E‡), entalphies (∆H‡), and free energies (∆G‡) (kcal/mol) for transition structures TS1a/1b-TS9a/9b relative to the less stable isomer (1b-9b) obtained by application of the Marcus equation. Gas phase Struct. TS1a/1b TS2a/2b TS3a/3b TS4a/4b TS5a/5b TS6a/6b TS7a/7b TS8a/8b TS9a/9b

∆E 2.83 2.97 2.37 2.78 2.38 2.84 2.86 2.42 2.61





∆H 0.35 0.56 0.07 0.28 0.04 0.38 0.47 0.11 0.11

Chloroform ‡

∆G 0.45 0.89 0.27 0.73 0.18 1.06 0.97 0.32 0.42

∆E 2.22 2.38 1.67 2.38 1.91 2.41 2.28 1.87 2.06





∆H 0.10 0.14 0.01 0.08 0.01 0.20 0.09 0.01 0.09

Dimethyl sulfoxide ‡

∆G 0.23 0.51 0.05 0.24 0.05 0.72 0.25 0.03 0.23

∆E 1.99 2.16 1.42 2.14 1.71 2.25 1.89 1.70 1.65





∆H 0.00 0.13 0.04 0.04 0.02 0.08 0.03 0.00 0.00

Ethanol ‡

∆G 0.09 0.40 0.01 0.30 0.00 0.44 0.82 0.04 0.54

∆E 2.04 2.20 1.43 2.27 1.79 2.32 2.19 1.73 1.91





∆H 0.06 0.15 0.02 0.07 0.00 0.17 0.13 0.00 0.05



∆G 0.21 0.27 0.08 0.32 0.01 0.72 0.63 0.01 0.69

CONCLUSIONS

A series of well-established aromaticity indices, namely Σ∆N, I, APoz,KP, Q, λ, HOMA, HOMED, and HOMHED, provide insightful correlations to evaluate the

electronic

delocalization of pseudo-heterocyclic tautomers of 2-aryl-substituted alkyl(aryl)aminoacroleins and 3-aryl-substituted 1,5-dialkyl(aryl)vinamidines, as well as their transition structures corresponding to the intramolecular proton transfer. Our DFT-based study (M06-2X/6311++G(d,p) conducted in both the gas phase and polar solvents allowed us to estimate the influence of substituents on the conjugated backbone. No appreciable solvent effect could be detected from this computation. The striking fact that negative ∆G‡ values were found for the proton transfer between iminoenol and aminoketone tautomers of 3-aminoacroleins, characterized by low energy barriers

ACS Paragon Plus Environment

19

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 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 27

and large imaginary frecuencies, can be attributed to variational effects, which are usually overlooked for processes involving higher energy barriers (as the proton transfer from aminoketone to iminoenol tautomers in the aforementioned 3-aminoacroleins). These negative ∆G‡ values can however be easily corrected by applying the Marcus theory.

ASSOCIATED CONTENT Supporting Information Mathematical description of aromaticity indices (appendix 1), additional data from theoretical calculations and Cartesian coordinates for all structures, as well as full Reference 19 (Gaussian package). This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author [email protected] Notes †

In memoriam Paul von Ragué Schleyer.

The authors declare no competing financial interest.

ACKNOWLEDGMETNS This work was supported b the Spanish Ministry of Science and Innovation (CTQ201018938/BQU) and the Autonomous Government of Extremadura (Ayudas a Grupos Consolidados, Grant GRU10049). We thank the Centro de Investigación, Innovación Tecnológica y Supercomputación de Extremadura (CenitS) for allowing us the use of supercomputer LUSITANIA.

ACS Paragon Plus Environment

20

Page 21 of 27

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 56 57 58 59 60

The Journal of Physical Chemistry

REFERENCES

(1) Jeffrey, G. A. in An Introduction to Hydrogen Bonding; Truhlar, D. G., Ed.; Oxford University Press: New York; 1997. (2) Desiraju, G. R.; Steiner, T. The Weak Hydrogen Bond in Structural Chemistry and Biology; Oxford Universty Press: Oxford, 1999. (3) Hydrogen Bonding-New Inshigts; Grabowski, S. J., Ed.; Springer: Dordrecht, The Netherlands, 2006. (4) Gilli, G.; Gilli, P. The Nature of the Hydrogen Bond: Outline of a Comprehensive Hydrogen Bond Theory. Oxford University Press: Oxford; 2009. (5) Desiraju, G. R.; Vittal, J. J.; Ramanan, A. Crystal Engineering. A Textbook; World Scientific Publishing: Singapore, 2011. (6) Cleland, W. W.; Kreevoy, M. M. Low-Barrier Hydrogen Bonds and Enzymic Catalysis. Science 1994, 264, 1887-1890. (7) Frey, P. A.; Whitt, S. A.; Tobin, J. B. A Low-Barrier Hydrogen Bond in the Catalytic Triad of Serine Proteases. Science 1994, 264, 1927-1930. (8) Warshel, A.; Papazyan, A.; Kollman, P. A. On Low-Barrier Hydrogen Bonds and Enzyme Catalysis. Science 1995, 269, 102-106. (9) Zheng, Y.-J.; Bruice, T. C. Is Strong Hydrogen Bonding in the Transition State Enough to Account for the Observed Rate Acceleration in a Mutant of Papain? Proc. Natl. Acad. Sci. USA 1997, 94, 4285-4288. (10) Schowen, K. B.; Limbach, H.-H.; Denisov, G. S.; Schowen, R. L. Hydrogen Bonds and Proton Transfer in General-Catalytic Transition-State Stabilization in Enzyme Catalysis. Biochim. Biophys. Acta 2000, 1458, 43-62. (11) Kim, K. S.; Oh, K. S.; Lee, J. Y. Catalytic Role of Enzymes: Short Strong H-BondInduced Partial Proton Shuttles and Charge Redistributions. Proc. Natl. Acad. Sci. USA 2000, 97, 6373-6378. (12) Vishveshwara, S.; Madhusudhan, M. S.; Maizel Jr., J. V. Short-Strong Hydrogen Bonds and a Low Barrier Transition State for the Proton Transfer Reaction in RNase A Catalysis: A Quantum Chemical Study. Biophys. Chem. 2001, 89, 105-117.

ACS Paragon Plus Environment

21

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 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 27

(13) Gilli, P.; Bertolasi, V.; Ferretti, V.; Gilli, G. Evidence for Resonance-Assisted Hydrogen Bonding. 4. Covalent Nature of the Strong Homonuclear Hydrogen Bond. Study of the O-H--O System by Crystal Structure Correlation Methods. J. Am. Chem. Soc. 1994, 116, 909-915. (14) Gilli, G.; Bellucci, F.; Ferretti, V.; Bertolasi, V. Evidence for Resonance-Assisted Hydrogen Bonding from Crystal-Structure Correlations on the Enol Form of the β-Diketone Fragment. J. Am. Chem. Soc. 1989, 111, 1023-1028. (15) Bertolasi, V.; Gilli, P.; Ferretti, V.; Gilli, G. Evidence for Resonance-Assisted Hydrogen Bonding. 2. Intercorrelation between Crystal Structure and Spectroscopic Parameters in Eight Intramolecularly Hydrogen Bonded 1,3-Diaryl-1,3-propanedione Enols. J. Am. Chem. Soc. 1991, 113, 4917-4925. (16) Krygowski, T. M.; Szatylowicz, H.; Stasyuk, O. A.; Dominikowska, J.; Palusiak, M. Aromaticity from the Viewpoint of Molecular Geometry: Application to Planar Systems. Chem. Rev. 2014, 114, 6383-6422. (17) Gilli, P.; Bertolasi, V.; Pretto, L.; Ferretti, V.; Gilli, G. Covalent versus Electrostatic Nature of the Strong Hydrogen Bond: Discrimination among Single, Double, and Asymmetric Single-Well Hydrogen Bonds by Variable-Temperature X-ray Crystallographic Methods in βDiketone Enol RAHB Systems. J. Am. Chem. Soc. 2004, 126, 3845-3855. (18) Palusiak, M.; Simon, S.; Solà, M. Interplay Between Intramolecular Resonance-Assisted Hydrogen Bonding and Aromaticity in o-Hydroxyaryl Aldehydes. J. Org. Chem. 2006, 71, 52415248. (19) Sanz, P.; Mó, O.; Yañez, M.; Elguero. J. Resonance-Assisted Hydrogen Bonds: A Critical Examination. Structure and Stability of the Enols of β-Diketones and β-Enaminones. J. Phys. Chem. A 2007, 111, 3585-3591. (20) Sanz, P.; Mó, O.; Yañez, M.; Elguero, J. Bonding in Tropolone, 2-Aminotropone, and Aminotroponimine: No Evidence of Resonance-Assisted Hydrogen-Bond Effects. Chem. Eur. J. 2008, 14, 4225-4232. (21) Palusiak, M.; Simon, S.; Solà, M. Interplay Between Intramolecular Resonance-Assisted Hydrogen Bonding and Local Aromaticity. II. 1,3-Dihydroxyaryl-2-aldehydes. J. Org. Chem. 2009, 74, 2059-2066.

ACS Paragon Plus Environment

22

Page 23 of 27

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 56 57 58 59 60

The Journal of Physical Chemistry

(22) Dobosz, R.; Gawinecki, R. DFT Studies on Tautomeric Preferences: Proton Transfer in 1,5-Bis(pyridin-2-yl)-

and

1,5-Bis(quinolin-2-yl)pentane-2,4-diones.

J.

Mol.

Struct.

(THEOCHEM) 2010, 940, 119-123. (23) Krygowski, T. M.; Zachara-Horeglad, J. E.; Palusiak, M. H-Bonding-Assisted Substituent Effect. J. Org. Chem. 2010, 75, 4944-4949. (24) Zarycz, N.; Aucar, G. A.; Della Védova, C. O. NMR Spectroscopic Parameters of Molecular Systems with Strong Hydrogen Bonds. J. Phys. Chem. A 2010, 114, 7162-7172. (25) Guillaumes, L.; Simon, S. Dealing with Quasi-Ring Formation by Two Hydrogen Bonds. Cooperativity Analysis with Delocalization Indices. J. Phys. Chem. A 2014, 118, 9727-9733. (26) Zarycz, M. N. C.; Provasi, P. F. Investigation of the Resonance-Assisted Hydrogen Bond in Model β-Diketones through Localized Molecular Orbital Analysis of the Spin-Spin Coupling Constants Related to the O-H···O Hydrogen Bond. Magn. Reson. Chem. 2014, DOI: 10.1002/mrc.4152. (27) Balaban, A. T.; Schleyer, P. v. R.; Rzepa, H. S. Crocker, Not Armit and Robinson, Begat the Six Aromatic Electrons. Chem. Rev. 2005, 105, 3436-3447. (28) Wu, J. I.; Jackson, J. E.; Schleyer, P. v. R. Reciprocal Hydrogen Bonding-Aromaticity Relationships. J. Am. Chem. Soc. 2014, 136, 13526-13529. (29) Romero-Fernández, M. P.; Ávalos, M.; Babiano, R.; Cintas, P.; Jiménez, J. L.; Light, M. E.; Palacios, J. C. Pseudo-Cyclic Structures of Mono- and Di-Azaderivatives of Malondialdehydes. Synthesis and Conformational Disentanglement by Computational Analyses. Org. Biomol. Chem. 2014, 12, 8997-9010. (30) Zhao, Y.; Truhlar, D. G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120, 215-241. (31) Ditchfield, R.; Hehre, W. J.; Pople, J. A. Self-Consistent Molecular-Orbital Methods. IX. An Extended Gaussian-Type Basis for Molecular-Orbital Studies of Organic Molecules. J. Chem. Phys. 1971, 54, 724-728. (32) Hehre, W. J.; Ditchfield, R.; Pople, J. A. Self-Consistent Molecular-Orbital Methods. XII. Further Extensions of Gaussian-Type Basis Sets for Use in Molecular-Orbital Studies of Organic Molecules. Mol. Phys. 1974, 27, 209-214.

ACS Paragon Plus Environment

23

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 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 27

(33) Gordon, M. S. The Isomers of Silacyclopropane. Chem. Phys. Lett. 1980, 76, 163-168. (34) Hariharan, P. C.; Pople, J. A. Influence of Polarization Functions on MO Hydrogenation Energies. Theor. Chim. Acta 1973, 28, 213-222. (35) Francl, M. M.; Pietro, W. J.; Hehre, W. J.; Binkley, J. S.; DeFrees, D. J.; Pople, J. A.; Gordon, M. S. Self-Consistent Molecular Orbital Methods. XXIII. A Polarization-Type Basis Set for Second-Row Elements. J. Chem. Phys. 1982, 77, 3654-3665. (36) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A. et al. Gaussian 09, revision A.1., Gaussian, Inc.: Wallingford CT, 2009. (37) Zhao, Y.; Truhlar, D. G. Density Functionals with Broad Applicability in Chemistry. Acc. Chem Res. 2008, 41, 157-167. (38) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Performance of SM6, SM8, and SMD on the SAMPL1 Test Set for the Prediction of Small-Molecule Solvation Free Energies. J. Phys. Chem. B 2009, 113, 4538-4543. (39) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378-6396. (40) Ribeiro, R. F.; Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Prediction of SAMPL2 Aqueous Solvation Free Energies and Tautomeric Ratios using the SM8, SM8AD, and SMD Solvation Models. J. Comput.-Aided Mater. 2010, 24, 317-333. (41) Halim, M. A.; Shaw, D. M.; Poirier, R. A. Medium Effect on the Equilibrium Geometries, Vibrational Frequencies and Solvation Energies of Sulfanilamide. J. Mol. Struct. (THEOCHEM) 2010, 960, 63-72. (42) Saielli, G. Differential Solvation Free Energies of Oxonium and Ammonium Ions: Insights from Quantum Chemical Calculations. J. Phys. Chem. A 2010, 114, 7261−7265. (43) Tomasi, J.; Mennucci, B.; Cancès, E. The IEF Version of the PCM Solvation Method: An Overview of a New Method Addressed to Study Molecular Solutes at the QM ab initio level. J. Mol. Struct. (THEOCHEM) 1999, 464, 211-226. (44) Liotard, D. A.; Hawkins, G. D.; Lynch, G. C.; Cramer, C. J.; Truhlar, D. G. Improved Methods for Semiempirical Solvation Models. J. Comp. Chem. 1995, 16, 422−440.

ACS Paragon Plus Environment

24

Page 25 of 27

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 56 57 58 59 60

The Journal of Physical Chemistry

(45) Fringuelli, F.; Marino, G.; Taticchi, A.; Grandolini, G. A Comparative Study of the Aromatic Character of Furan, Thiophen, Selenophen, and Tellurophen. J. Chem. Soc., Perkin Trans. 2 1974, 332-337. (46) Bird, C. W. A New Aromaticity Index and its Application to Five-Membered Ring Heterocycles. Tetrahedron 1985, 41, 1409-1414. (47) Bird, C. W. The Application of a New Aromaticity Index to Six-Membered Ring Heterocycles. Tetrahedron 1986, 42, 89-92. (48) Bird, C. W. Heteroaromaticity. 8. The Influence of N-Oxide Formation on Heterocyclic Aromaticity. Tetrahedron 1993, 49, 8441-8448. (49) Kotelevskii, S. I.; Prezhdo, O. V. Aromaticity Indices Revisited: Refinement and Application to Certain Five-Membered Ring Heterocycles. Tetrahedron 2001, 57, 5715-5729. (50) Pozharskii, A. F. Theoretical Principles of Heterocyclic Chemistry; Khimia: Moscow, 1985; p. 22. (51) Pozharskii, A. F. Heteroaromaticity. Chem. Heterocycl. Compd. 1985, 21, 717-749. (52) Gilli, G.; Bellucci, F.; Ferretti, V.; Bertolasi, V. Evidence for Resonance-Assisted Hydrogen Bonding from Crystal-Structure Correlations on the Enol Form of the β-Diketone Fragment. J. Am. Chem. Soc. 1989, 111, 1023-1028. (53) Mezei, M.; Beveridge, D. L. Free Energy Simulations. Anal. N. Y. Acad. Sci. 1986, 482, 123. (54) Kruszewski, J.; Krygowski, T. M. Definition of Aromaticity Basing on the Harmonic Oscillator Model. Tetrahedron Lett. 1972, 13, 3839-3842. (55) Krygowski, T. M. Crystallographic Studies of Inter- and Intramolecular Interactions Reflected in Aromatic Character of π-Electron Systems. J. Chem. Inf. Comput. Sci. 1993, 33, 7078. (56) Raczynska, E. D.; Duczmal, K.; Hallmann, M. Tautomeric Equilibria and π-Electron Delocalization for the Substrate (Pyruvate) and Inhibitor (Oxamate) of Lactate Dehydrogenase – DFT Studies. Polish J. Chem. 2007, 81, 1655-1666. (57) Raczynska, E. D.; Hallmann, M.; Duczmal, K. Consequences of Covalent and NonCovalent Interactions on Conformational Preferences and π-Electron Delocalization for the Substrate (Pyruvate) and Inhibitor (Oxamate) of Lactate Dehydrogenase – DFT Studies. Polish J. Chem. 2008, 82, 1077-1090.

ACS Paragon Plus Environment

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 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 27

(58) Raczynska, E. D.; Hallmann, M.; Kolczynska, K.; Stepniewski, T. On the Harmonic Oscillator Model of Electron Delocalization (HOMED) Index and Its Application to Heteroaromatic π-Electron Systems. Symmetry 2010, 2, 1485-1509. (59) Frizzo, C. P.; Martins, M. A. P. Aromaticity in Heterocycles: New HOMA Index Parametrization. Struct. Chem. 2012, 23, 375-380. (60) Gilli, P.; Bertolasi, V.; Pretto, L.; Lyčka, A.; Gilli, G. The Nature of Solid-State N-H--O/O-H---N Tautomeric Competition in Resonant Systems. Intramolecular Proton Transfer in Low-Barrier

Hydrogen

Bonds

Formed

by

the

--OC-CN-NH--



--HO-CC-NN-

Ketohydrazone.Azoenol System. A Variable-Temperature X-ray Crystallographic and DFT Computational Study. J. Am. Chem. Soc. 2002, 124, 13554-13567. (61) Marcus, R. A. Theoretical Relations among Rate Constants, Barriers, and Broensted Slopes of Chemical Reactions. J. Phys. Chem. 1968, 72, 891-899. (62) Gilli, P.; Bertolasi, V.; Pretto, L.; Antonov, L.; Gilli, G. Variable-Temperature X-ray Crystallographic and DFT Computational Study of the NH---O/N---HO Tautomeric Competition in 1-(Arylazo)-2-naphtols. Outline of a Transition-State Hydrogen-Bond Theory. J. Am. Chem. Soc. 2005, 127, 4943-4953. (63) Alabugin, I. V.; Manoharan, M.; Breiner, B.; Lewis, F. D. Control of Kinetics and Thermodynamics of [1,5]-Shifts by Aromaticity: A View through the Prism of Marcus Theory. J. Am. Chem. Soc. 2003, 125, 9329-9342. (64) Hayden, A. E.; Houk, K. N. Transition State Distortion Energies Correlate with Activation Energies of 1,4-Dihydrogenations and Diels-Alder Cycloadditions of Aromatic Molecules. J. Am. Chem. Soc. 2009, 131, 4084-4089.

ACS Paragon Plus Environment

26

Page 27 of 27

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 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment