Type-I dyotropic rearrangement for 1,2-disubstituted cyclohexanes: substitution effect on activation energy

Nadeem S. Sheikh
Department of Chemistry, Faculty of Science, King Faisal University, P.O. Box 380, Al-Ahsa 31982, Saudi Arabia. E-mail: nsheikh@kfu.edu.sa; Fax: +966 13-588-6437; Tel: +966 13-589-5393, +966 58-589-9574

Received 1st December 2015 , Accepted 11th January 2016

First published on 19th January 2016


Abstract

For an array of judiciously selected 1,2-disubstituted cyclohexane motifs, the migratory aptitude and contribution of specific structural features of synthetically valuable functional groups and halogen atoms in type-I dyotropic rearrangement are reported by employing quantum chemical calculations. This double migration process requires higher activation barriers for carbon, silicon and nitrogen bearing substituents, however, it is thermally allowed and a feasible approach for several important moieties including oxygen, sulphur, phosphorous and halogen atoms bearing migrating fragments. Strong positive correlations are observed by plotting representative activation energy trends against inductive sigma (σI), field (σF), steric substituent (Es) and polar substituent (σ*) constants. Also, an effect of the asymmetric combinations of substituents is presented which provides an interesting insight into such substitution patterns.


Introduction

Type-I dyotropic rearrangement1 is a highly stereo- and regioselective pericyclic process, which proceeds reversibly through a cyclic transition state in a concerted manner.2 A distinctive feature of this rearrangement involves one pi (π) system and a simultaneous intramolecular exchange of two sigma (σ) bonds (Fig. 1). Based on molecular-orbital and valence-bond analyses, this double migration process for 1,2-disubstitued ethane has been explicitly proposed to proceed either via an ethylenic scaffold or a diradical species.3 Contrary to this, type-II dyotropic rearrangement4 occurs via migration of two substituents to new sites in a molecule. In addition to a usual C–C5 stationary scaffold, dyotropic reaction has been successfully applied to C–B,6 C–N,7 C–O,8 C–Al,9 C–Si,10 C–S,8,11 C–Cu,12 C–Zn,13 C–Zr,14 C–Pd,15 C[double bond, length as m-dash]Si,16 N–N,17 O–N,18 Si[double bond, length as m-dash]Si,19 Fe–Pt,20 and Ge[double bond, length as m-dash]Sn,21 used as static frameworks.
image file: c5ra25482e-f1.tif
Fig. 1 Generic representation of type-I dyotropic rearrangement.

This double migration approach has been articulately incorporated to construct structurally complex molecules of synthetic and biological relevance such as lacrimin A (1),22 jaspamide (2),23 zoapatanol (3),24 and azafenestrane (4, Fig. 2).25 Moreover, several other bioactive natural products such as milbemycin β3,26 discodermolide,27 xanthanolides,28 cephalostatin analogues29 and a number of synthetically significant building blocks30 have also been prepared through this methodology. Additionally, type-II dyotropic rearrangement has also found its applications towards the synthesis of bioactive derivatives including cyclic peptide TCM-95A and B.31


image file: c5ra25482e-f2.tif
Fig. 2 Synthetic applications of type-I dyotropic rearrangement towards representative bioactive molecules.

Chemical reactions proceed through one or more transition states and the rate of the reaction depends on the activation energy (Ea), which is the free energy difference between the transition state (TS) and the reactants. It is a measure of chemical reactivity and provides a very useful information about the reaction profile. Also, it is directly linked with stability of the TS, which relies on several decisive factors, not just limited to (a) electronic interactions between the reacting centres, (b) charge distributions, (c) steric hindrance caused by substituents, (d) mesomeric, inductive and anomeric effects. In case of type-I dyotropic rearrangement, hybridization of the atoms of the migrating groups directly attached with the stationary scaffold has a pronounced influence on Ea due to associated electronic nature and steric effect. In general, sp3 hybridized atoms are electron donating moieties whereas both sp2 and sp hybridized atoms tend to withdraw electronic density from a molecule. In case of substituted aromatic systems, the electronic influence is controlled by the substituent attached with the aromatic system. The transition states are stabilized/destabilized as a result of all these factors, which have a direct impact on the Ea; greater the stability of the TS, lower will be the Ea. Quantum chemical calculations provide an insightful information towards the elucidation of reaction mechanisms and chemical reactivity descriptors.32 Application of computational investigations in synthesis has emerged as an indispensable research element and indeed, is a time and cost effective approach. It has been eloquently applied to dyotropic rearrangement and is reported for several structural units such as dithienylethenes,33 arene/allene cycloadducts,34 phosphate and sulphate anions,35 azines,36 pentalenene,37 vicinal dibromides,38 β-lactones,39 dimethylaurate,40 organocuprates and organoargentates,41 and nitroso acetals.42

Undoubtedly, cyclohexane is regarded as one of the most privileged cycloalkane due to associated structural features. It exists in several interconvertible conformations along with a perfectly unstrained and highly stable chair conformation. It is a fairly unreactive molecule with limited applications43 however, substituted cyclohexane is of substantial synthetic utility and acts as a precursor to numerous organic transformations. Moreover, replacement of hydrogen atoms with larger substituents in cyclohexane skeleton can prevent the interconversion between different conformations and this phenomenon is a common observation in cyclohexane containing structures of biological importance. Considering its significance, a comprehensive investigation towards the migratory aptitude of synthetically significant functional groups/atoms in type-I dyotropic rearrangement for variously functionalized 1,2-disubstituted cyclohexane structures is described.

Results and discussion

In general, type-I dyotropic rearrangement occurs at high temperatures as reported for thermal mutarotation of 2,3-dibromo-tert-butylcyclohexane,44 and vicinal dibromides in steroidal structures.45 The results obtained from this study are in agreement with this experimental observation and 1,2-shifts for the considered groups proceed via high energy barriers up to 126 kcal mol−1. Such high barriers have also been reported for the type-I dyotropic model systems studied for 1,2-disubstituted ethane structures.3b The results for 1,2-disubstituted cyclohexanes bearing migratory groups with sp3-hybridized atoms and halogen atoms directly attached with the cyclohexane skeletons are summarized in Table 1 and the optimized structures of the corresponding four-membered transition states are provided in the ESI (Fig. 1).
Table 1 Type-I dyotropic reactions for migratory substituents with sp3-hybridized atoms and halogen atoms directly attached with the cyclohexane unit
Entry Substituents (R and R′) Activation energya (Ea), kcal mol−1 Bond distanceb (r), Å Average bond distancea,b (r), Å Average bond anglea,b (θ)
Cα–Cβ Cα/Cβ–R/R′ Cα–R–Cβ and Cα–R′–Cβ
a For a detailed information about energies, bond distances and bond angles (see ESI).b The values are for the corresponding transition states.
1 CH3 120.9 1.368 2.425 32.77
2 NH2 92.02 1.445 2.065 43.15
3 OH 76.85 1.416 1.973 42.04
4 F 58.56 1.423 1.909 43.77
5 SiH3 88.62 1.371 2.879 27.52
6 PH2 63.67 1.376 2.814 28.28
7 SH 49.00 1.418 2.492 33.04
8 Cl 36.24 1.434 2.384 35.00
9 CH2CH3 120.2 1.362 2.541 31.28
10 CH(CH3)2 109.4 1.353 2.700 29.03
11 C(CH3)3 95.83 1.345 3.107 25.00
12 CH2CHCH2 90.16 1.353[thin space (1/6-em)] 3.039 25.54
13 Si(CH3)3 97.00 1.362 3.010 26.13
14 P(CH3)2 68.22 1.361 2.934 26.81
15 OCH3 71.29 1.418 1.967 42.25
16 SCH3 50.14 1.417 2.461 33.45
17 OC(CH3)3 89.28 1.431 2.011 41.67
18 OCH2Ph 79.10 1.422 2.000 41.65
19 OCH2OCH3 77.53 1.422 2.033 40.90
20 OSi(CH3)3 71.10 1.424 1.975 42.27
21 OC(O)CH3 63.88 1.420 2.059 40.32
22 OPh 60.99 1.420 2.021 41.14
23 OC(O)Ph 50.42 1.432 2.084 40.20
24 Br 31.50 1.428 2.506 33.11


It is evident that 1,2-migration for carbon bearing groups requires higher energy barriers; the maximum for CH3 groups (entry 1) and it gradually decreases when CH3 motifs are replaced by NH2 (entry 2), OH (entry 3) and F (entry 4) substituents. In the transitions states, the bond distance between the carbon atoms (Cα and Cβ) bearing CH3 groups is 1.368 Å, which is an indication of an olefinic bond. However for NH2, OH and F migrating groups, it is in the range of 1.416–1.445 Å.46 Also, the average bond distance for Cα/Cβ–R/R′ is comparatively higher for CH3 groups (2.425 Å) as compared to NH2, OH and F moieties (1.909–2.065 Å).

This decrease in the computed activation barrier in going from CH3 to F substituents may be ascribed to the increase in the bond polarity or inductive effect, which is in an order of CH3 < NH2 < OH < F. It significantly influences to stabilize the transition states and decreases the Ea. Also, it has been proposed for 1,2-disusbstituted ethane molecules that halogen atoms as migrating groups provide better stabilization to the transition states through an additional donor–acceptor orbital interaction between the lone pairs of halogen atoms and the newly formed ethylenic π* LUMO.3b This additional stabilization completes the pericyclic circuit and consequently, a lower activation barrier is observed. In addition to this, linear free-energy correlations for the energy barriers and inductive sigma (σI) and field (σF) constants for CH3, NH2, OH and F groups/atoms are shown in Fig. 3.47 From the positive correlation coefficients obtained, it can be concluded that the energy barrier has a strong relationship with the corresponding Hammett constants. Replacing with the substituents bearing the atoms belonging to the third period of the periodic table (entries 5–8), a similar trend is observed; Ea is in an order of SiH3 > PH2 > SH > Cl. Also, the presence of a weaker C–R bond in the reactants corresponds to a lower Ea and it has been noticed for the pairs SiH3 and CH3, PH2 and NH2, OH and SH, F and Cl. Similar observations for the type-I dyotropic rearrangement barrier have been reported for the pairs CH3 and SiH3, F and Cl substituted ethane molecules and explained in terms of activation-strain and molecular-orbital analyses.3b


image file: c5ra25482e-f3.tif
Fig. 3 The correlation between the activation energy (Ea) of CH3, NH2, OH and F as migrating groups with their corresponding (a) inductive sigma constants (σI) and (b) field constants (σF).

Interestingly, sequential exchange of hydrogen atoms with CH3 groups in CH3 substituents significantly reduces the energy barrier (entries 9–11). For this particular case of carbon bearing migrating fragments, an increase in the steric crowding (by replacing hydrogen atoms with CH3 groups) results in an increase in the average bond distances (2.541–3.107 Å) between the carbon atoms (Cα and Cβ) and the substituents (R and R′) involved in the transition states. It is conceivable to infer that this increase in the bond distances leads to lower the steric interactions between the substituents and the carbon atoms, which consequently stabilize the corresponding transition states and result in decreasing the Ea for this double migration process. Notably, the bond angles between carbon atoms (Cα and Cβ) and the substituents (R and R′) seem to be related with the steric crowdedness as a gradual decrease in the bond angles is recorded from CH2CH3 to C(CH3)3 substituents (31.28–25.00°). These results are also supported by linear free-energy relationship described by the Taft equation and strong positive correlations are observed between the activation energy and steric substituent (Es) and polar substituent (σ*) constants for CH3, CH3CH2, CH(CH3)2 and C(CH3)3 as shown in Fig. 4.48


image file: c5ra25482e-f4.tif
Fig. 4 The correlation between the activation energy (Ea) of CH3, CH3CH2, CH(CH3)2 and C(CH3)3 as migrating groups with their corresponding (a) steric substituent constants (Es) and (b) polar substituent constants (σ*).

In addition to this, the migrating ability of allyl group (entry 12) requires higher Ea, which is probably because of the repulsive interaction between the π-bonds present in the allylic substituents and the newly formed π-bond in the corresponding transition state (TS12, Fig. 1, ESI). Such a repulsion leads to destabilize the TS and causes an increase in the energy barrier. However, the steric effect becomes more pronounced when there is a slight increase in the bond distances between the carbon atoms (Cα and Cβ) and the substituents (R and R′) involved in the transition states. It is the case when SiH3 (entry 5) is substituted with Si(CH3)3 (entry 13); the difference for average bond distance is of 0.131 Å and for CH3 and C(CH3)3 groups, it is 0.682 Å. It is plausible that an increase in the steric crowding by replacing H atoms with CH3 groups result in an enhanced steric repulsion between the substituents and the stationary scaffold, which destabilizes the resultant transition states. This explains why rearrangement of Si(CH3)3 substituents is linked with a higher Ea than the corresponding [1,2]-shifts for SiH3 groups. Comparing to their respective hydrogenated equivalents, similar effect towards an increase in Ea has been computed for P(CH3)2 (entry 14), OCH3 (entry 15) and SCH3 (entry 16).

At this juncture, an investigation was carried out to understand the behaviour of commonly used OH protecting groups for this [1,2]-shifts approach and a regular pattern can be envisaged. From OH (entry 3) to OC(CH3)3 (entry 17), introduction of the steric bulk significantly increases the energy barrier and this agrees well with other findings of this study, where the steric factor is a dominant contributor. The functional groups OCH2Ph (entry 18) and OCH2OCH3 (entry 19) have Ea values similar to OH groups (entry 3). For OSi(CH3)3 substituents (entry 20), a lower energy barrier of 71.10 kcal mol−1 is calculated as compared to OH migrating fragments (76.85 kcal mol−1). This decrease in the Ea may be ascribed to the presence of a more polarized O–Si bond than an O–H bond. A significant decrease in the activation barrier is computed for OC(O)CH3 migrating units (entry 21), where the TS receives better stabilization due to the presence of carbonyl groups. Stabilization of the TS caused by delocalization of the lone pairs of oxygen substituents over aromatic system renders a lower barrier (60.99 kcal mol−1) and it has been noticed for OPh groups (entry 22). As anticipated, the energy barrier for OC(O)Ph (entry 23) substituents is remarkably lower than the corresponding OH groups by 26.43 kcal mol−1 due to additional stabilization, which is clear from the TS23 (Fig. 1, ESI). It is the lowest activation barrier for the considered OH protecting groups, which is a consequence of more stabilized TS due to the presence of improved delocalization. The most feasible migrating fragments for the type-I dyotropic rearrangement among the considered systems are Br atoms (entry 24) with a reaction barrier of 31.5 kcal mol−1. This is due to weaker C–Br bonds in the reactant and it is one of the consistent aspects of this study; weaker the C–R bond in the reactants, lower will be the energy barrier.

For comparative purposes and to comprehensively document this investigation, the migratory aptitude for the groups bearing sp2- and sp-hybridized carbon atoms was explored. A range of valuable functional moieties were selected, which provides an important information about the specific behaviour of the substituents in this double migration process and the results are summarized in Table 2. In the presence of these functional groups, type-I dyotropic rearrangement for cyclohexanes as stationary scaffolds proceeds via high activation barriers (more than 100 kcal mol−1) with an exception of C(O)Cl groups. Vinyl groups as migrating substituents require the highest energy barrier (126.1 kcal mol−1, entry 1) that is probably due to an increased electronic interaction between the olefinic bonds of the substituents and the newly formed π-bond in the corresponding TS (TS25, Fig. 1, ESI). A significant decrease of 8.9 kcal mol−1 is observed when H atoms of the carbon atoms of vinyl groups attached with the cyclohexane unit are swapped by CH3 groups. By doing this, the steric crowdedness is increased, which generally results in higher Ea. However, the average distance between the newly formed π-bond in the TS and the substituents increases from 2.31 to 2.372 Å. Probably, this increase stabilizes the TS by reducing the interactions, which would lead to a lower Ea. The net effect of this replacement is a decrease in the activation barrier and it is in accordance with the findings for other carbon bearing groups such as going from CH3 to C(CH3)3 substituents, entries 1 and 11 of Table 1 respectively.

Table 2 Type-I dyotropic reactions for migratory fragments with sp2- and sp-hybridized atoms directly attached with the cyclohexane unit
Entry Substituents (R and R′) Activation energya (Ea), kcal mol−1 Bond distanceb (r), Å Average bond distancea,b (r), Å Average bond anglea,b (θ)
Cα–Cβ Cα/Cβ–R/R′ Cα–R–Cβ and Cα–R′–Cβ
a For a detailed information about energies, bond distances and bond angles (see ESI).b The values are for the corresponding transition states.
1 CHCH2 126.1 1.378 2.310 34.73
2 C(CH3)CH2 117.3 1.375 2.372 33.70
3 C(O)H 111.6 1.364 2.582 30.64
4 C(O)CH3 104.5 1.362 2.697 29.24
5 C(O)NH2 109.4 1.369 2.539 31.27
6 C(O)NHPh 105.1 1.369 2.597 30.57
7 C(O)OCH3 111.1 1.378 2.421 33.06
8 C(O)OH 110.6 1.380 2.410 33.27
9 C(O)OC(O)CH3 108.6 1.381 2.417 33.18
10 C(O)F 108.5 1.387 2.372 34.00
11 C(O)Cl 92.50 1.385 2.417 33.31
12 Ph 119.8 1.379 2.328 34.44
13 CCH 115.1 1.431 2.197 40.40
14 CN 106.9 1.410 2.183 37.69


The presence of carbonyl groups as migrating substituents withdraw the electronic density from the newly formed π-bonds in a TS, which promotes destabilization of the TS and increases the Ea. Another factor is steric hindrance offered by bulky substituents attached with the carbonyl groups, which leads to an increased bond distances between the carbon atoms (Cα and Cβ) and the substituents (R and R′) involved in the transition states. This results in lowering the steric interactions hence a decrease in activation barrier as observed for carbon bearing groups as migrating substituents. So, the carbonyl substituents behave differently and the resultant energy barrier is directly linked with the extent of electron withdrawal and the steric hindrance. The computed activation barriers for C(O)H (entry 3) and C(O)CH3 (entry 4) are 111.6 and 104.5 kcal mol−1 respectively. These values are lower than vinyl groups (entry 1) due to increased bond distances and lower repulsive interactions for carbonyl bearing groups as compared to vinyl substituents. The average computed bond distance for C(O)CH3 substituents is 2.697 Å, while for C(O)H groups is 2.582 Å. Also, CH3 groups tend to reduce the electron withdrawing ability of the carbonyl group due to hyperconjugation, which leads to stabilize the TS and consequently Ea is decreased. In case of nitrogen containing substituents like C(O)NH2 (entry 5) and C(O)NHPh (entry 6), there is a difference of 4.3 kcal mol−1. This may be ascribed to a possible extended delocalization of electronic cloud through the involvement of π-electrons from the Ph ring, which can reduce the electron withdrawing ability of the carbonyl group (TS30, Fig. 1, ESI). This can assist to stabilize the TS as compared to the situation when no Ph group is present and consequently, Ea is significantly decreased.

For oxygen bearing carbonyl groups such as C(O)OCH3 (entry 7), C(O)OH (entry 8) and C(O)OC(O)CH3 (entry 9), the energy barriers are 111.1, 110.6 and 108.6 kcal mol−1 respectively. Moving from C(O)OC(O)CH3 substituents to C(O)OCH3, a difference of 2.5 kcal mol−1 is observed that is probably related to better stabilization of newly formed π-bond in the TS by the oxonium species through hyperconjugation. A significant difference of 16.0 kcal mol−1 in the energy barrier is observed for C(O)F (entry 10) and C(O)Cl (entry 11), which seems to be due to the repulsive interaction of lone pairs on F and Cl atoms with the newly formed π-bond. The average bond distances between the substituents and the carbon atoms (Cα–Cβ) suggest that these interactions are higher for C(O)F substituents as it is 2.372 Å compared to 2.417 Å for C(O)Cl migrating groups. In addition to this, higher electronegativity of F over Cl atom can destabilize the corresponding TS by withdrawing electronic density from the newly formed π-bond, and causes a higher Ea for C(O)F substituents.

An installation of Ph groups (entry 12) as migrating substituents show higher value of Ea, 119.8 kcal mol−1, which is significantly lower than the vinyl substituents (entry 1). This seems to be as a result of delocalized π-bonds in the aromatic system, which renders lower electronic interactions between the newly formed π-bond and the π-bonds in the Ph rings. For other synthetic transformations, the existence of ortho/para and/or meta directing groups on an aromatic ring plays a decisive rule to determine the reactivity and selectivity. In case of type-I dyotropic rearrangement for 1,2-disubstituted cyclohexane, there seems to be a very minor contribution offered by these ortho/para and meta directors located on Ph substituents.49 For CCH (entry 13) and CN (entry 14) groups, activation barriers of 115.1 and 106.9 kcal mol−1 and bond distances (Cα–Cβ) of 1.431 and 1.41 Å are observed respectively.46 This decrease in Ea in going from CCH to CN may be rationalized on the basis of their optimized transition state structures (TS37 and TS38, Fig. 1, ESI). It is reasonable to suggest that in the case of CN as migrating groups, planarity of the substituents in the TS renders better stabilization and lower repulsive interactions in the TS as compared to CCH groups. Consequently, a lower activation barrier is noticed for CN migrating groups in type-I dyotropic rearrangement of cyclohexanes.

Finally, an investigation towards the energy barriers for asymmetric combinations of substituents was carried out and the results are summarized in Table 3. It provides an interesting information towards captodative substitution, which demonstrates a combined effect of electron withdrawing (captor) and electron donating (dative) substituents attached to the same substrate.50

Table 3 Effect of asymmetric combinations of substituents on the activation barriers of 1,2-disubstituted cyclohexanes
Entry Substituents Activation energya (Ea), kcal mol−1 Bond distanceb (r), Å
R R′ Cα–Cβ
a For a detailed information about energies, bond distances and bond angles (see ESI).b The values are for the corresponding transition states.
1 CH3 NH2 105.0 1.399
2 CH3 OH 90.5 1.388
3 CH3 F 77.2 1.385
4 NH2 OH 72.7 1.449
5 NH2 F 57.3 1.443
6 OH F 62.0 1.421
7 CH3 C(O)CH3 113.2 1.367
8 NH2 C(O)CH3 99.3 1.404
9 OH C(O)CH3 86.9 1.389
10 F C(O)CH3 73.8 1.385
11 CH3 CN 94.7 1.394
12 NH2 CN 73.5 1.463
13 OH CN 86.4 1.427
14 F CN 81.5 1.415
15 C(O)CH3 CN 93.0 1.387
16 F Cl 47.5 1.432
17 F Br 45.2 1.430
18 Cl Br 33.9 1.431


Certain captodative substitution patterns have been found effective in decreasing the activation barriers for several pericyclic transformations.51 An explanation of the results presented in Table 3 appear somewhat intricate however, couple of aspects can be clearly seen:

(1) For most of the substitution patterns studied, the activation energies for unsymmetrical disubstituted cyclohexanes are in between the barrier values obtained for the corresponding individual symmetrical disubstituted cyclohexane systems.

(2) An acceleration effect is observed for certain combinations where the energy barriers of asymmetric 1,2-disubstituted cyclohexanes are significantly lower than the analogous symmetrical disubstituted cyclohexanes. These combinations include the presence of two mesomerically electron donating substituents such as NH2 and OH (entry 4), an electron donating and an electron withdrawing species like CH3 and CN (entry 11) and both electron withdrawing migrating fragments such as C(O)CH3 and CN (entry 15).

Conclusions

A comprehensive investigation towards type-I dyotropic rearrangement for a range of 1,2-disubstituted cyclohexane molecules is presented. The results obtained for the considered groups/atoms are in good agreement with the experimental observations and reported DFT based calculations for the related systems. To establish a free-energy linear relationship, representative activation energy trends are plotted against inductive sigma (σI), field (σF), steric substituent (Es) and polar substituent (σ*) constants, which provide a strong positive correlation. The salient conclusions drawn from this double migration process include:

(a) Inductive effect of the substituents plays a substantial role in determining the feasibility of this approach and it is a thermally allowed process for 1,2-dihalogenated cyclohexanes due to added stabilization by the interaction of halogen lone pairs with the newly formed π-bond.

(b) Oxygen, sulphur and phosphorus bearing migrating fragments require moderate levels of activation barriers and these have also been computed for the most common OH protecting groups. The most convenient OH protecting group for type-I dyotropic reaction of 1,2-disubstituted cyclohexane is O(CO)Ph with an activation barrier of 50.42 kcal mol−1, which is 26.43 kcal mol−1 lower than the corresponding OH substituted system.

(c) Installation of carbon, silicon and nitrogen bearing migrating groups proceed via higher energy barriers; the maximum is for carbon atom.

(d) The presence of carbonyl groups as migrating units result in higher activation energies for the considered groups because of enhanced destabilization of the transition states.

(e) Phenyl (Ph) rings as migrating substituents have higher energy barriers and there is a slight difference in the Ea caused by the synergistic effect of electron donating and withdrawing groups.

(f) Increase in steric bulk of the substituents lead to higher energy barriers and increase in the bond distances between the carbon atoms (Cα and Cβ) and the substituents (R and R′) involved in the transition states decreases the activation barrier.

(g) Asymmetric combinations of substituents usually proceed with an energy barrier related to the corresponding individual symmetrically disubstituted systems however for certain unsymmetrical combinations, such as NH2 and OH, CH3 and CN, and C(O)CH3 and CN, an extra acceleration effect is observed which leads to a significant decrease in the activation energy of 1,2-asymmteric disubstituted cyclohexane molecules.

Methods

All calculations were performed with the Gaussian 09W programme52 and the results were produced with GaussView 5.0. Density functional theory (DFT)53 calculations using the B3LYP functional54 were used to locate all the stationary points involved. Geometries were optimized at B3LYP/6-31+G(d,p) level of theory, which has been successfully applied to related dyotropic reactions.37,42 The frequency calculations were run at the same level of theory to confirm each stationary point to be either a minimum or a transition structure. The transition states were also linked to their corresponding minima through the intrinsic reaction coordinate (IRC)55 calculations (see ESI for details), which confirm the connection of transition structures with the reactants and products. Zero-point energy (ZPE) values were computed from the optimized geometries at 298.15 K and are not corrected.

Acknowledgements

The author gratefully acknowledges the support offered by the King Faisal University, Saudi Arabia. Also, D. Leonori (University of Manchester, UK) is thanked for the technical discussions.

Notes and references

  1. M. T. Reetz, Angew. Chem., Int. Ed. Engl., 1972, 11, 129–130 CrossRef CAS.
  2. For a selection of reviews concerning dyotropic rearrangement, see: (a) M. F. Croisant, R. van Hoveln and J. M. Schomaker, J. Org. Chem., 2015, 27, 5897–5907 Search PubMed; (b) I. Fernandez and F. M. Bickelhaupt, Chem. Soc. Rev., 2014, 43, 4953–4967 RSC; (c) E. M. Greer and C. V. Cosgriff, Annu. Rep. Prog. Chem., Sect. B: Org. Chem., 2013, 109, 328–350 RSC; (d) O. Gutierrez and D. J. Tantillo, J. Org. Chem., 2012, 77, 8845–8850 CrossRef CAS PubMed; (e) D. J. Tantillo and J. K. Lee, Annu. Rep. Prog. Chem., Sect. B: Org. Chem., 2011, 107, 266–286 RSC; (f) I. Fernandez, F. P. Cossio and M. A. Sierra, Chem. Rev., 2009, 109, 6687–6711 CrossRef CAS PubMed; (g) I. D. Gridnev, Coord. Chem. Rev., 2008, 252, 1798–1818 CrossRef CAS; (h) M. T. Reetz, Adv. Organomet. Chem., 1977, 16, 33–65 CrossRef CAS.
  3. (a) H. M. Buck, Open J. Phys. Chem., 2013, 3, 119–125 CrossRef CAS; (b) I. Fernandez, F. M. Bickelhaupt and F. P. Cossío, Chem.–Eur. J., 2012, 18, 12395–12403 CrossRef CAS PubMed.
  4. M. T. Reetz, Angew. Chem., Int. Ed. Engl., 1972, 11, 131–132 CrossRef.
  5. For C–C stationary scaffold, see selected examples: (a) D. C. Braddock, D. Roy, D. Lenoir, E. Moore, H. S. Rzepa, J. I.-C. Wu and R. Schleyer, Chem. Commun., 2012, 48, 8943–8945 RSC; (b) J. G. Buchanan, G. D. Ruggiero and I. H. Williams, Org. Biomol. Chem., 2008, 6, 66–72 RSC; (c) F. Stöckner, R. Beckert, D. Gleich, E. Birckner, W. Günther and H. Görls, J. Org. Chem., 2007, 8, 1237–1243 Search PubMed; (d) F. Stöckner, C. Käpplinger, R. Beckert and H. Görls, Synlett, 2005, 643–645 Search PubMed; (e) G. D. Ruggiero and I. H. Williams, Chem. Commun., 2002, 732–733 RSC; (f) K. Sato, Y. Yamashita and T. Mukai, Tetrahedron Lett., 1981, 22, 5303–5306 CrossRef CAS.
  6. (a) E. Hupe, D. Denisenko and P. Knochel, Tetrahedron, 2003, 59, 9187–9198 CrossRef CAS; (b) G. Zweifel and H. Arzoumanian, J. Am. Chem. Soc., 1967, 89, 5086–5088 CrossRef CAS.
  7. J. W. Davies, J. R. Malpas and M. P. Walker, Chem. Commun., 1985, 686–687 RSC.
  8. (a) M. T. Reetz, M. Kliment, M. Plachky and N. Greif, Chem. Ber., 1976, 108, 2728–2742 CrossRef; (b) M. T. Reetz, M. Kliment and M. Plachky, Chem. Ber., 1976, 108, 2716–2727 CrossRef; (c) M. T. Reetz, M. Kliment and M. Plachky, Angew. Chem., Int. Ed. Engl., 1974, 13, 813–814 CrossRef.
  9. A. Alexakis, J. Hanaizi, D. Jachiet, J.-F. Normant and L. Toupet, Tetrahedron Lett., 1990, 31, 1271–1274 CrossRef CAS.
  10. (a) W. Uhl, J. Bohnemann, M. Layh and E.-U. Wuerthwein, Chem.–Eur. J., 2014, 20, 8771–8781 CrossRef CAS PubMed; (b) A. Naka, S. Ueda, J. Ohshita, A. Kunai, T. Miura, H. Kobayashi and M. Ishikawa, Organometallics, 2008, 27, 2922–2928 CrossRef CAS; (c) Y. Yu and S. Feng, Int. J. Quantum Chem., 2007, 107, 105–115 CrossRef CAS; (d) Y. Yu and S. Feng, J. Phys. Chem. A, 2006, 110, 12463–12469 CrossRef CAS PubMed; (e) Y. Yu, S. Feng and D. Feng, J. Phys. Chem. A, 2005, 109, 3663–3668 CrossRef CAS PubMed; (f) L. Claes, J.-P. François and M. S. Deleuze, J. Am. Chem. Soc., 2003, 125, 7129–7138 CrossRef CAS PubMed; (g) H. Bürger and P. Moritz, Organometallics, 1993, 12, 4930–4939 CrossRef; (h) R. Köster, G. Seidel, R. Boese and B. Wrackmeyer, Chem. Ber., 1990, 123, 1013–1028 CrossRef; (i) M. T. Reetz and N. Greif, Angew. Chem., Int. Ed. Engl., 1977, 16, 712–713 CrossRef; (j) W. I. Bevan, R. N. Haszeldine, J. Middleton and A. E. Tipping, J. Organomet. Chem., 1970, 23, C17–C19 CrossRef CAS.
  11. A. Haas, Angew. Chem., Int. Ed. Engl., 1965, 4, 1014–1023 CrossRef CAS.
  12. N. J. Rijs, B. F. Yates and R. A. J. O'Hair, Chem.–Eur. J., 2010, 16, 2674–2678 CrossRef CAS PubMed.
  13. T. Harada, D. Hara, K. Hattori and A. Oku, Tetrahedron Lett., 1988, 29, 3821–3824 CrossRef CAS.
  14. (a) G. Erker and R. Petrenz, Organometallics, 1992, 11, 1646–1655 CrossRef CAS; (b) G. Erker and R. Petrenz, Chem. Commun., 1989, 345–346 RSC; (c) E. Negishi, K. Akiyoshi, B. O'Corner, K. Takagi and G. Wu, J. Am. Chem. Soc., 1989, 111, 3089–3091 CrossRef CAS; (d) A. S. Ward, E. A. Mintz and M. P. Kramer, Organometallics, 1988, 7, 8–12 CrossRef CAS; (e) G. Erker and F. Rosenfeldt, J. Organomet. Chem., 1980, 188, C1–C4 CrossRef CAS; (f) G. Erker and F. Rosenfeldt, Angew. Chem., Int. Ed. Engl., 1978, 17, 605–606 CrossRef.
  15. J. Faessler and S. Bienz, Organometallics, 1994, 13, 4704–4707 CrossRef CAS.
  16. K. M. Baines, A. G. Brook, R. R. Ford, P. D. Lickiss, A. K. Sexana, W. J. Chatterton, J. F. Sawyer and B. A. Behnam, Organometallics, 1989, 8, 693–709 CrossRef CAS.
  17. E. Gellermann, U. Klingebiel, T. Pape, F. Dall'Antonia, T. R. Schneider and S. Schmatz, Z. Anorg. Allg. Chem., 2001, 627, 2581–2588 CrossRef CAS.
  18. (a) C. Ekber, S. Schamatz, F. Diedric and U. Klingebiel, Silicon Chem., 2003, 2, 117–123 CrossRef; (b) F. Diedric, U. Klingebiel and M. Schäfer, J. Organomet. Chem., 1999, 588, 242–246 CrossRef; (c) T. Albercht, G. Elter and A. Meller, Z. Anorg. Allg. Chem., 1999, 625, 1453–1456 CrossRef.
  19. (a) H. B. Yokelson, D. A. Siegel, A. J. Millevolte, J. Maxka and R. West, Organometallics, 1990, 9, 1005–1010 CrossRef CAS; (b) H. B. Yokelson, D. A. Siegel and R. West, J. Am. Chem. Soc., 1986, 108, 4239–4241 CrossRef CAS.
  20. (a) P. Braunstein, M. Knorr, G. Reinhard, U. Schubert and T. Stährfeldt, Chem.–Eur. J., 2000, 6, 4265–4278 CrossRef CAS PubMed; (b) P. Braunstein, M. Knorr, B. Hirle, G. Reinhard and U. Schubert, Angew. Chem., Int. Ed. Engl., 1992, 104, 1583–1585 CrossRef.
  21. A. Sekiguchi, R. Izumi, V. Y. Lee and M. Ichinohe, Organometallics, 2003, 22, 1483–1486 CrossRef CAS.
  22. A. Takle and P. Kocienski, Tetrahedron, 1990, 46, 4503–4516 CrossRef CAS.
  23. P. Kocienski and C. Barber, Pure Appl. Chem., 1990, 62, 1933–1940 CrossRef CAS.
  24. P. Kocienski, C. J. Love, R. J. Whitby, G. Costello and D. A. Roberts, Tetrahedron, 1989, 45, 3839–3848 CrossRef CAS.
  25. (a) S. E. Denmark, J. I. Montgomery and L. A. Kramps, J. Am. Chem. Soc., 2006, 128, 11620–11630 CrossRef CAS PubMed; (b) S. E. Denmark and J. I. Montgomery, Angew. Chem., Int. Ed., 2005, 44, 3732–3736 CrossRef CAS PubMed.
  26. (a) P. Kocienski, C. Yates, S. D. A. Street and S. F. Campbell, J. Chem. Soc., Perkin Trans. 1, 1987, 2183–2187 RSC; (b) P. Kocienski, S. D. A. Street, C. Yates and S. F. Campbell, J. Chem. Soc., Perkin Trans. 1, 1987, 2189–2194 RSC.
  27. E. Lemos, F.-H. Poŕee, A. Commercon, J.-F. Betzer, A. Pancrazi and J. Ardisson, Angew. Chem., Int. Ed., 2007, 46, 1917–1921 CrossRef PubMed.
  28. W. Ren, Y. Bian, Z. Zhang, H. Shang, P. Zhang, Y. Chen, Z. Yang, T. Luo and Y. Tang, Angew. Chem., Int. Ed., 2012, 51, 6984–6988 CrossRef CAS PubMed.
  29. W. Li, T. G. LaCour and P. L. Fuchs, J. Am. Chem. Soc., 2002, 124, 4548–4549 CrossRef CAS PubMed.
  30. (a) V. C. Purohit, A. S. Matla and D. Romo, J. Am. Chem. Soc., 2008, 130, 10478–10479 CrossRef CAS PubMed; (b) I. Arrastia, B. Lecea and F. P. Cossío, Tetrahedron Lett., 1996, 37, 245–248 CrossRef CAS; (c) J. Mulzer, A. Pomtner, R. Strasser, K. Hoyer and U. Nagel, Tetrahedron Lett., 1995, 36, 3679–3682 CrossRef CAS; (d) T. H. Black and T. S. McDermott, J. Chem. Soc., Chem. Commun., 1991, 184–185 RSC; (e) T. H. Black, S. A. Eisenbeis, T. S. McDermott and S. L. Maluleka, Tetrahedron, 1990, 46, 2307–2316 CrossRef CAS; (f) T. H. Black, J. A. Hall and R. G. Sheu, J. Org. Chem., 1988, 53, 2371–2374 CrossRef CAS; (g) T. H. Black and J. D. Fields, Synth. Commun., 1988, 29, 1747–1750 CAS; (h) J. Mulzer and G. Brütrup, Angew. Chem., Int. Ed. Engl., 1979, 18, 793–794 CrossRef.
  31. S. Lin and S. J. Danishefsky, Angew. Chem., Int. Ed., 2002, 41, 512–515 CrossRef CAS.
  32. S. M. Bachrach, Computational Organic Chemistry, Wiley, 2nd edn, 2014 Search PubMed.
  33. T. Mahmood, M. Arshad, M. A. Gilani and K. Ayub, J. Mol. Model., 2015, 21, 1–9 CrossRef CAS PubMed.
  34. H. V. Pham, A. S. Karns, C. D. Vanderwal and K. N. Houk, J. Am. Chem. Soc., 2015, 137, 6956–6964 CrossRef CAS PubMed.
  35. S. Yamabe, G. Zeng, W. Guan and S. Sakaki, J. Comput. Chem., 2014, 35, 2195–2204 CrossRef CAS PubMed.
  36. Maria, M. Hanif, T. Mahmood, R. Ludwig and K. Ayub, J. Mol. Model., 2014, 20, 2304–2307 CrossRef CAS PubMed.
  37. M. W. Lodewyk, D. Willenbring and D. J. Tantillo, Org. Biomol. Chem., 2014, 12, 887–894 CAS.
  38. I. Fernandez, M. A. Sierra and F. P. Cossio, Chem.–Eur. J., 2006, 12, 6323–6330 CrossRef CAS PubMed.
  39. (a) C. A. Leverett, V. C. Purohit, A. G. Johnson, R. L. Davis, D. J. Tantillo and D. Romo, J. Am. Chem. Soc., 2012, 134, 13348–13356 CrossRef CAS PubMed; (b) R. L. Davis, C. A. Leverett, D. Romo and D. J. Tantillo, J. Org. Chem., 2011, 76, 7167–7174 CrossRef CAS PubMed.
  40. N. J. Rijs, G. B. Sanvido, G. N. Khairallah and R. A. J. O'Hair, Dalton Trans., 2010, 39, 8655–8662 RSC.
  41. N. J. Rijs and R. A. J. O'Hair, Organometallics, 2010, 29, 2282–2291 CrossRef CAS.
  42. R. L. Davis and D. J. Tantillo, J. Org. Chem., 2010, 75, 1693–1700 CrossRef CAS PubMed.
  43. A major application of cyclohexane is an industrial scale preparation of adipic acid and caprolactam, which are used to manufacture nylon.
  44. D. H. R. Barton and A. J. Head, J. Chem. Soc., 1956, 932–937 RSC.
  45. C. A. Grob and S. Winstein, Helv. Chim. Acta, 1952, 35, 782–802 CrossRef CAS.
  46. This exceptional increase in the bond distance is due to higher synchronicity values and computed synchronicity values for NH2, OH and CN are 0.61, 0.85 and 0.63 a.u. respectively.
  47. C. Hansch, A. Leo and R. W. Taft, Chem. Rev., 1991, 91, 165–195 CrossRef CAS.
  48. E. V. Anslyn and D. A. Dougherty, Modern Physical Organic Chemistry, University Science Books, 2006, p. 455 Search PubMed.
  49. Compared to Ph groups substituted cyclohexane, a slight difference in the energy barriers is noticed for the CH3, NH2, OH, OCH3, Cl, C(O)H, C(O)CH3 and COOH groups bearing Ph substituted cyclohexanes, both at meta and para positions separately.
  50. For a selection of reviews concering captodative substitution effect, see: (a) H. G. Viehe, Z. Janousek, R. Merenyi and L. Stella, Acc. Chem. Res., 1985, 18, 148–154 CrossRef CAS; (b) R. Sustmann and H. G. Korth, Adv. Phys. Org. Chem., 1990, 26, 131–178 CrossRef CAS.
  51. For selected examples of captodative substitution in pericyclic transformations, see: (a) T. J. Greshock and R. L. Funk, J. Am. Chem. Soc., 2006, 128, 4946–4947 CrossRef CAS PubMed; (b) T.-Q. Yu, Y. Fu, L. Liu and Q.-X. Guo, J. Org. Chem., 2006, 71, 6157–6164 CrossRef CAS PubMed; (c) N. A. Magomedov, P. L. Ruggiero and Y. Tang, J. Am. Chem. Soc., 2004, 126, 1624–1625 CrossRef CAS PubMed; (d) J. Mertes and J. Mattay, Helv. Chim. Acta, 1988, 71, 742–748 CrossRef CAS; (e) J. L. Boucher and L. Stella, Tetrahedron, 1986, 42, 3871–3885 CrossRef CAS; (f) J. L. Boucher and L. Stella, Tetrahedron Lett., 1985, 26, 5041–5044 CrossRef CAS; (g) J. L. Boucher and L. Stella, Tetrahedron, 1985, 41, 875–887 CrossRef CAS.
  52. M. J. Frisch, et al., GAUSSIAN09, (Revision D.01), Gaussian, Inc., Wallingford CT, 2013 Search PubMed; the full reference can be found in the ESI..
  53. R. G. Parr and W. Yang, Density-Functional Theory of Atoms and Molecules; Oxford University Press, U.K., 1989 Search PubMed.
  54. (a) P. J. Stephens, F. J. Devlin, C. F. Chabalowski and M. J. Frisch, J. Chem. Phys., 1994, 98, 11623–11627 CrossRef CAS; (b) A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef CAS; (c) A. D. Becke, J. Chem. Phys., 1993, 98, 1372–1377 CrossRef CAS; (d) C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785–789 CrossRef CAS.
  55. (a) C. Gonzalez and H. B. Schlegel, J. Phys. Chem., 1990, 94, 5523–5527 CrossRef CAS; (b) C. Gonzalez and H. B. Schlegel, J. Phys. Chem., 1989, 90, 2154–2161 CrossRef CAS; (c) K. Fukui, Acc. Chem. Res., 1981, 14, 363–368 CrossRef CAS; (d) Due to unsuccessful IRC calculations for some cases, the transition state structures were linked to their corresponding minima via perturbing their structures in the direction along the vibrational coordinate linked with the imaginary frequency, followed by optimization.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra25482e

This journal is © The Royal Society of Chemistry 2016