Reinvestigation of Acetophenones Oxidation by Performic Acid in

J. Phys. Chem. A , Article ASAP. DOI: 10.1021/acs.jpca.8b11256. Publication Date (Web): February 21, 2019. Copyright © 2019 American Chemical Society...
1 downloads 0 Views 327KB Size
Subscriber access provided by Washington University | Libraries

A: Spectroscopy, Molecular Structure, and Quantum Chemistry

Reinvestigation of the Acetophenones Oxidation by Performic Acid in Formic Acid Lino Reyes, María Inés Nicolás-Vázquez, Cristina Iuga, and Juan Raul Alvarez-Idaboy J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.8b11256 • Publication Date (Web): 21 Feb 2019 Downloaded from http://pubs.acs.org on February 27, 2019

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

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 12 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

Reinvestigation of the Acetophenones Oxidation by Performic Acid in Formic Acid Lino Reyes,*† María Inés Nicolás-Vázquez,§ Cristina Iuga∥ and J. Raul AlvarezIdaboy*‡ Facultad de Química, †Departamento de Química Organica, and ‡Departamento de Física y Química Teorica, Universidad Nacional Autonoma de Mexico, C.P. 04510, Mexico City, México §Facultad

de Estudios Superiores Cuautitlán, Universidad Nacional Autónoma de México, Cuautitlan 54740, Estado México, Mexico ∥Universidad

Autónoma Metropolitana-Xochimilco, Calzada del Hueso 1100, Mexico D.F. 04960,

Mexico

ABSTRACT The Baeyer-Villiger (BV) reaction of acetophenones R-COCH3 (R = phenyl, 4methylphenyl, 3,4-dimethoxyphenyl) with performic acid (PFA) in formic acid (FA) as catalyst and solvent was reinvestigated, using the MPWB1K functional in conjunction with the 6-311G(d,p) and 6-311++G(d,p) basis sets. For the acid-catalyzed addition step, we used the 8-membered ring transition structure proposed in our previous work. The calculated and experimental results obtained for the BV reaction under the mentioned conditions lead to the conclusion that our mechanism is more reliable than the one reported by Liu and coworkers, in which the acid-catalyzed first step involves a transition state with a six-membered ring structure.

ACS Paragon Plus Environment

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 12

1. Introduction The Baeyer-Villiger (BV) reaction1 is a relevant synthetic procedure of organic chemistry which has been widely used in the oxidation of ketones and aldehydes to form esters and acids respectively. The essential mechanism of the BV rearrangement can be described as a process that includes the attack of a peroxyacid to the carbonyl group to form a tetrahedral adduct (Criegge intermediate), followed by the migration of an alkyl or aryl group to one of the peroxide oxygens and the concerted dissociation of the O-O bond (Figure 1).2 O R

O

+

CH3

H

O

H

HCO2H

C

O

OOH

H

H3C R

O

H3C R

O

R= phenyl, 4-methylphenyl, 3,4-dimethoxyphenyl

H

OH

O

O H

O

O

O

H

Criegee Intermediate

TS1 H3C

O

R H3C R

H O

O

OH

O

H

O

O

O

TS2 O

H3C

H O

H3C

H

O

R

H

OH

O

R

C H

O

O

O H

O

+

H O

TS2-cat

Figure 1. Baeyer-Villiger Reaction of Acetophenones with Performic acid. Although the migration is considered as the rate-determining step (RDS), several experimental data suggest that this step is strongly dependent on the structure of the substrate as well as on the reaction conditions.3-5 A number of theoretical studies have been performed to clarify its mechanism.6-28 2 ACS Paragon Plus Environment

Page 3 of 12 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

In 2006 we proposed a new mechanism for the acid-catalyzed addition including a concerted transition structure (TS) with the lowest free energy barrier to date.20. In this transition structure, the protonation of the ketone oxygen is concurrent with deprotonation of the peracid. In later work, we also have supplied theoretical proof that sustain a noncatalyzed TS for the migration step, as previously proposed by Cárdenas et al.14 Considering both steps of our suggested mechanism, we have calculated the rate constants for the oxidation of propanone and cyclohexanone with trifluoroperacetic acid (TFPAA), catalyzed by trifluoroacetic acid (TFAA) in dichloromethane. The results are in line with experimental data, supporting the validity of this mechanism.22 In a later publication, our proposed mechanism was used in a theoretical study of the oxidation of some p-substituted acetophenones with performic (PFA) and trifluoroperacetic (TFPAA) acids, catalyzed by formic (FA) and trifluoroacetic (TFAA) acids, using dichloromethane as solvent.26 The results showed that, for these systems, the two steps are concerted, and that the addition step is catalyzed, while the migration step is not. Substituent effects on the migration Gibbs free energy barrier are very pronounced, to the point that a shift in the RDS is possible. This fact is in agreement with the statement mentioned above, i. e. that the RDS is strongly affected by several factors, such as the structure of the reactants, the catalyst and the solvent polarity. The influence of solvent polarity has been explored by Hawthorne and Emmons, by performing a kinetic study of the BV reaction using several aliphatic and aromatic ketones reacting with TFPAA catalyzed by TFAA.27 In all cases, from ethylene chloride to acetonitrile-ethylene chloride (10:1 volume), migration was the RDS. In order to explore the BV reaction in polar solvents, we performed a theoretical study on the oxidation of propanone with performic acid and formic acid as catalyst.28 The results agree well with the experimental kinetic values, the rate constant of the reaction decreases as the solvent polarity increases. In addition, our calculations showed that migration continued to be the rate determining step as solvent polarity increased. Some years ago, Liu and colaborators29 performed a theoretical and experimental study of the BV reactions of acetophenone, 4-methyl acetophenone and 3,4-dimethoxy acetophenone, with performic acid (PFA) in formic acid (FA) acting both as catalyst and solvent. The authors claim that, according to their results, the addition step is the RDS for the three 3 ACS Paragon Plus Environment

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 12

aromatic ketones. This conclusion challenges our previous results, since in our work, Reference 23, on the reactions of acetophenone and 4-methyl acetophenone with performic acid catalyzed by formic acid employing dichloromethane as solvent23 we showed that migration is clearly the RDS. In our previous work (ref 23) we proposed a transition state with an 8-membered ring for the acid-catalyzed addition step of de BV reaction (Figure 2a). In ref 29 the authors found that in the acid-catalyzed addition step for the substituted acetophenones, the transition state with a six-membered ring structure (Figure 2b) is more stable than the uncatalyzed one. Using this structure they obtained ΔG≠ values that seems to be incorrect (41.63, 34.27 and 34.13 kcal/mol for H, 4-CH3 and 3,4-dimethoxi derivatives, respectively), since such large barriers correspond to reactions that do not occur significantly at usual reaction temperatures.

O

H

C

O

H

O

R

H

O

H

O

O

O H

H O

H3C

O

H3C R

H

O

O H

O

a) 8-membered ring TS

b) 6-membered ring TS

Figure 2. Transition structures for the acid-catalyzed addition step of de BV reaction. For the migration step, the authors of ref 29 obtained higher ΔG≠ values for the catalyzed transition state than for the corresponding uncatalyzed TS. They called to this observation “a complex situation”, assuming that the acid-catalyzed path is the preferred channel for the BV reaction. Thus, they selected the total electronic energies with thermal corrections (ΔE) to describe the energetic changes in the reaction mechanism of the BV reaction. In addition, they investigated experimentally the reaction kinetics of the oxidation with performic acid, using FA as both solvent and catalyst. From the kinetic rate constants obtained at several temperatures, they calculated the activation energy (Ea) for the reaction of 3,4-dimethoxy acetophenone and 4-methyl acetophenone employing the Arrhenius formula (7.5 and 8.2 kcal/mol respectively). These values were compared with the thermal corrections of electronic energies (ΔE) obtained for these aromatic ketones (8.7 and 9.1 kcal/mol 4 ACS Paragon Plus Environment

Page 5 of 12 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

respectively), and the conclusion was that there is a good agreement between the experimental and the theoretical values. In previous studies we showed that, in acidic media, the mechanism of the Baeyer-Villiger reaction appears to be generally acid catalyzed in the first step and not catalyzed in the second one.21-22 This fact can be explained considering that, while the acid catalyst decreases the enthalpy for the migration step, the entropy loss overtakes it. To prove the arguments above, we decided to model the mechanism proposed in ref 29, as well as ours, for both steps of the BV oxidation of acetophenones, R-COCH3 (R= phenyl, 4methylphenl, 3,4-dimethoxyphenyl), with PFA, using FA as both catalyst and solvent (Figure 1). 2. Computational Details All the calculations were carried out using Gaussian 09 package30 employing the MPWB1K functional31 and the 6-311G(d,p) basis set in the gas phase. The energy data were refined by single-point calculations with the same functional and the 6-311++G(d,p) basis set, with inclusion of the IEF-PCM continuum model using formic acid as solvent and the UFF radii. In our preceding studies, this functional has been demonstrated to give excellent results that agree very well with previous experimental values.22-31 The thermodynamic magnitudes were obtained by vibrational calculation. The Gibbs energies were obtained via thermodynamic cycle using corrections to Gibbs energies in gas phase at 298.15 K using standard state 1M and corrected electronic energies in solution. Furthermore, the solvent cage effects have been included according to the corrections proposed by Benson, considering the free volume theory.32 Intrinsic reaction coordinate (IRC) calculations were used to confirm that the located transition states properly connect the intended reactants and products. For further details, refer to References (20 and 33).

5 ACS Paragon Plus Environment

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 12

3. Results and Discussion The overall BV oxidation of the investigated ketones are presented in Figure 1. Considering that it has been observed that the noncatalyzed addition is infeasible, 19-21 in this work we have only studied the acid-catalyzed addition.20 To analyze the migration process, acidcatalyzed and not catalyzed options have been examined. For the addition process of the BV reaction, we have investigated two different acidcatalyzed transition structures: a) an eight membered ring TS proposed by us (TS1-A), and b) a six membered ring TS proposed in ref 29 (TS1-B). The transition structures TS1-A and TS1-B of the acid-catalyzed addition step for acetophenone are shown in Figure 3.

ΔG≠ = 21.03 kcal/mol

ΔG≠ = 29.59 kcal/mol

a) TS1-A

b) TS1-B

Figure 3. Transition structures of the acid-catalyzed addition step for acetophenone. The Gibbs free energies and enthalpies of the different stationary points calculated in the BV oxidation of the revised acetophenones, are shown in Table 1.

6 ACS Paragon Plus Environment

Page 7 of 12 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. Enthalpies and Gibbs free energies (in kcal/mol, at 298.15 K) for the BV reaction of substituted acetophenones with PFA using FA as catalyst and solvent.a -H

-CH3

-Di-OCH3

H

G

H

0

0

0

TS1-A



21.03



20.69



20.49

TS1-B

15.32

29.59

14.94

29.00

14.65

28.55

Criegee + FA

0.14

4.14

-3.94

4.43

-3.36

4.68

13.26

21.50 (20.79)b

System Reactants

G

H

G

0

0

0

TS2nc + FA

16.18

24.53

14.97

23.60 (22.00)b

TS2cat

11.61

25.14

10.10

24.01

8.99

23.50

Products

-72.37

-77.96

-72.39

-78.71

-72.23

-77.66

a Level b

of theory: MPWB1K/6-311G++(d,p)-IEF-PCM//MPWB1K/6-311G(d,p). Refers to values calculated from the corresponding reaction rate constants extrapolated to 298.15 K.

For all systems reviewed, the ΔG≠ is lower for the not catalyzed migration than for the catalyzed one, because of entropy effects (Table 1). We can see in Table 1 that the transition structure calculated in ref 29 presents a considerable ΔG≠ for the catalyzed addition step, even after the inclusion of the thermodynamic correction for the liquid phase22. As mentioned above, these authors selected the catalyzed TS1 structure for the addition step used by Okuno9 and Grein19 (see Figure 2b), which was shown to have a higher energy than ours20 (Figure 2a). Reaction profile of the BV reaction of acetophenone in terms of Gibbs free energy for our mechanism (-) and for the corresponding to Ref 29 (- - -) are depicted in Figure 4.

7 ACS Paragon Plus Environment

The Journal of Physical Chemistry

TS1-B

Relative Gibbs free 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 8 of 12

TS2

TS1-A

Ref 29 This work

-80

Reaction Coordinate

Figure 4. Reaction profile of the BV reaction of acetophenone in terms of Gibbs free energy for our mechanism (-) and for the one reported in Ref 29 (- - -)

Considering the substituent effects, as in the case of the BV reaction of the mentioned system in dichloromethane, the electron-donating ability of the substituent attached to the ketone group induces a low decrease of the energy barrier for the addition step.23 This observation can be explained considering that these substituents promote the protonation of the ketone, producing the stabilization of TS1.34 The Gibbs free energy barrier of the migration step for all the studied BV reactions accords with the observation that the energies decrease with the electron-releasing aptitude of the group (H > 4-methyl > 3,4-dimethoxi), making the TS2 more stable.4 As observed in our previous work for this system,26 the electronic effects of the groups in the cases analyzed are less marked in the addition step (ΔΔG=0.54 kcal/mol) than in the

8 ACS Paragon Plus Environment

Page 9 of 12 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

migration process (ΔΔG=1.64 kcal/mol), which remains rate-determining. This means that the formic acid as solvent has no effect on the RDS. The experimental results performed in ref 29 represents a good opportunity to prove that our mechanism is reliable in formic acid, as was shown previously in dichoromethane.22 Thus, employing the experimental rate constants obtained in the mentioned work at various temperatures, and using the classical transition state theory, we calculated the corresponding ΔG≠ values at 298.15 K. These values were 22.00 and 20.79 kcal/mol for the oxidation of 4methyl acetophenone and 3,4-dimethoxy acetophenone by performic acid in formic acid respectively. These results are in excellent correspondence with the theoretical calculation results associated to the uncatalyzed migration step (23.60 and 21.50 kcal/mol, respectively), see Table 1. It is important to note that the obtained experimental ΔG≠ values cannot be associated to the addition step for these ketones, because the electronic effects of the substituents in this stage are negligible, as was mentioned above23,26 (see row 1 in Table 1). 4. Conclusions In this work, a systematic investigation on the mechanism of BV oxidation of acetophenones R-COCH3 (R= phenyl, 4-methylphenyl, 3,4-dimethoxyphenyl) with PFA in FA as catalyst and solvent has been performed employing DFT calculations. We showed that the two steps of the studied reactions are concerted, and that the addition step is powerfully catalyzed by the formic acid, while the migration step is not, which is the RDS in all the analyzed cases. The ΔG≠ values for the BV oxidation obtained from the experimental reaction rate constants of 4-methyl acetophenone and 3,4-dimethoxy acetophenone are in good agreement with the theoretical calculation results associated to the uncatalyzed migration step.

9 ACS Paragon Plus Environment

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 12

Supporting Information IRCs, figures and Cartesian coordinates of the transition state structures for the reaction of 4-methylacetophenone with PFA using FA as catalyst and solvent at the MPWB1K/6311G(d,p) level of theory.

Author Information Corresponding Authors *E-mails: [email protected], [email protected]. ORCID Lino Reyes: 0000-0003-3421-1634 J. Raúl Alvarez-Idaboy: 0000-0002-2901-5412 Notes The authors declare no competing financial interest. Acknowledgment The authors acknowledge the Project LANCAD-UNAM-DGTIC-154 and LANCADUNAM-DGTIC-192 of DGTIC at UNAM, for access to computing resources.

4. References (1) Baeyer, A.; Villiger, V. Einwirkung des Caro’schen Reagens auf Ketone. Ber. Dtsch. Chem. Ges. 1899, 32, 3625−3633. (2) Criegee, R. Die Umlagerung der Dekalin‐peroxydester als Folge von kationischem Sauerstoff. Justus Liebigs Ann. Chem. 1948, 560, 127−135. (3) Krow, G. R. The Baeyer-Villiger Oxidation of Ketones and Aldehydes. Org. React. 1993, 43, 251−798. (4) Renz, M.; Meunier, B. 100 Years of Baeyer–Villiger Oxidations. Eur. J. Org. Chem. 1999, 4, 737−750. (5) Brinkten, G.-J.; Arends, I. W. C. E.; Sheldon, R. A. The Baeyer−Villiger Reaction: New Developments toward Greener Procedures. Chem. Rev. 2004, 104, 4105−4123. (6) Reyes, L.; Castro, M.; Cruz, J.; Rubio, M. Substituent Effects in the Migration Step of the Baeyer-Villiger Rearrangement. A Theoretical Study. J. Phys. Chem. A 2005, 109, 3383−3390. (7) Snowden, M.; Bermudez, A.; Kelly, D. R.; Radkiewicz-Poutsma, J. L. The Preference for Anti over Gauche Migration in the Baeyer-Villiger Reaction. J. Org. Chem. 2004, 69, 7148–7156. (8) Itoh, Y.; Yamakata, M.; Mikami, K. Complete Reversal in Regioselectivity in the Baeyer−Villiger Reaction of an α-CF3-Ketone and Theoretical Rationale for Axial Orientation of Sterically Demanding CF3 Group at the Transition State. Org. Lett. 2003, 5, 4803−4806. (9) Okuno, Y. Theoretical Investigation of the Mechanism of the Baeyer‐Villiger Reaction in Nonpolar Solvents. Chem. Eur. J. 1997, 3, 212−218. (10) Stoute, V. A.; Winnik, M. A.; Csizmadia, I. G. Theoretical Model for the Baeyer-Villiger Rearrangement J. Am. Chem. Soc. 1974, 96, 6388−6393. (11) Rubio, M.; Cetina, R.; Bejarano, A. Theoretical Study of Baeyer-Villiger reaction. Afinidad 1983, 40, 176−179. (12) Rubio, M.; Reyes, L.; Cetina, R.; Pozas, R. Effect of Distant Substituents in the Baeyer-Villiger Oxidation of C-4 Substituted Cyclohexanones with m-Choroperbenzoic Acid. Afinidad 1989, 46, 341−344. (13) Hannachi, H.; Anoune, N.; Arnaud, C.; Lantéri, P.; Longeray, R.; Chermette, H. PM3 Semi-Empirical Study of Stereoelectronic Effects in the Baeyer-Villiger Reaction. J. Mol. Struct. 1998, 434, 183−191.

10 ACS Paragon Plus Environment

Page 11 of 12 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

(14) Cárdenas, R.; Cetina, R.; Lagunez-Otero, J.; Reyes, L. J. Ab Initio and Semiempirical Studies on the Transition Structure of the Baeyer and Villiger Rearrangement. The Reaction of Acetone with Performic Acid. J. Phys. Chem. A 1997, 101, 192−200. (15) Cárdenas, R.; Reyes, L.; Lagunez-Otero, J.; Cetina, R. Semiempirical Studies on the Transition Structure of the Baeyer and Villiger Rearrangement. The Reaction of Acetone with Alkyl and Aryl Peracids. J. Mol. Struct. 2000, 497, 211−225. (16) Lehtinen, C.; Nevalainen, V.; Brunow, G. Experimental and Computational Studies on Substituent Effects in Reactions of Peracid-Aldehyde Adducts. Tetrahedron. 2000, 56, 9375−9382. (17) Carlqvist, P.; Eklund, R.; Brinck, T. A Theoretical Study of the Uncatalyzed and BF3-Assisted Baeyer−Villiger Reactions. J. Org. Chem. 2001, 66, 1193−1199. (18) Sever, R. R.; Root, T. W. Computational Study of Tin-Catalyzed Baeyer−Villiger Reaction Pathways Using Hydrogen Peroxide as Oxidant. J. Phys. Chem. B 2003, 107, 10848−10862. (19) Grein, F.; Chen, A. C.; Edwards, D.; Crudden, C. M. Theoretical and Experimental Studies on the Baeyer-Villiger Oxidation of Ketones and the Effect of -Halo Substituents. J. Org. Chem. 2006, 71, 861−872. (20) Alvarez-Idaboy, J. R.; Reyes, L.; Cruz, J. A New Specific Mechanism for the Acid Catalysis of the Addition Step in the Baeyer−Villiger Rearrangement. J. Org. Lett. 2006, 8, 1763−1765. (21) Alvarez-Idaboy, J. R.; Reyes, L. Reinvestigating the Role of Multiple Hydrogen Transfers in BaeyerVilliger Reactions. J. Org. Chem. 2007, 72, 6580−6583. (22) Alvarez-Idaboy, J. R.; Reyes, L.; Mora-Diez, N. The Mechanism of the Baeyer–Villiger Rearrangement: Quantum Chemistry and TST Study Supported by Experimental Kinetic Data. Org. Biomol. Chem. 2007, 5, 3682−3689. (23) Reyes, L.; Alvarez-Idaboy, J. R.; Mora-Diez, N. Substituent Effects in the Baeyer-Villiger Reaction of Acetophenones. A Theoretical Study. J. Phys. Org. Chem. 2009, 22, 643−649. (24) Meneses, L.; Araya, A.; Pilaquinga, F.; Fuentealba, P. Relation between the Electrophilicity and p Hammett Constant in Baeyer-Villiger Reactions. Chem. Phys. Lett. 2008, 460, 27−30. (25) Mora-Diez, N.; Keller, S.; Alvarez-Idaboy, J. R. The Baeyer–Villiger Reaction: Solvent Effects on Reaction Mechanisms. Org. Biomol. Chem. 2009, 7, 3682−3690. (26) Reyes, L.; Díaz-Sánchez, C.; Iuga, C. Effects of Alkyl Groups in the Rate Determining Step of the Baeyer-Villiger Reaction of Phenyl Alkyl Ketones: A Quantum Chemistry Study. J. Phys. Chem. A 2012, 116, 7712-7717. (27) Hawthorne, M. F.; Emmons, W. D. A Re-examination of the Peroxyacid Cleavage of Ketones. II. Kinetics of the Baeyer-Villiger Reaction. J. Am. Chem. Soc. 1958, 80, 6398–6404. (28) Mora-Diez, N.; Keller, S.; Alvarez-Idaboy, J. R. The Baeyer–Villiger Reaction: Solvent Effects on Reaction Mechanisms. Org. Biomol. Chem. 2009, 7, 3682−3690. (29) Liu, B.; Meng, X.G.; Li, W.-Y.; Zhou, L.-C.; Hu, C.-W. Theoretical and Experimental Studies on Selective Oxidation of Aromatic Ketone by Performic Acid. J. Phys. Chem. A 2012, 116, 2920−2926. (30) 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 B.01; Gaussian, Inc.: Wallingford, CT, 2009. (31) Zhao, Y.; Truhlar, D. G. Hybrid Meta Density Functional Theory Methods for Thermochemistry, Thermochemical Kinetics, and Noncovalent Interactions: The MPW1B95 and MPWB1K Models and Comparative Assessments for Hydrogen Bonding and van der Waals Interactions. J. Phys. Chem. A 2004, 108, 6908–6918. (32) Benson, S. W. The Foundations of Chemical Kinetics; McGraw-Hill: New York, 1960. (33) Galano, A. Influence of Silicon Defects on the Adsorption of Thiophene-like Compounds on Polycyclic Aromatic Hydrocarbons:  A Theoretical Study Using Thiophene + Coronene as the Simplest Model. J. Phys. Chem. A 2007, 111, 1677–1682. (34) Bach, R. D.; Dmitrenko, O. Electronic Requirements for Oxygen Atom Transfer from Alkyl Hydroperoxides. Model Studies on Multisubstrate Flavin-Containing Monooxygenases. J. Phys. Chem. B 2003, 107, 12851–12861.

11 ACS Paragon Plus Environment

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 12

TOC Graphic

12 ACS Paragon Plus Environment