O. A.
Stasyuk
a,
H.
Szatylowicz
a,
T. M.
Krygowski
b and
C.
Fonseca Guerra
*c
aFaculty of Chemistry, Warsaw University of Technology, Noakowskiego 3, Warsaw 00-664, Poland
bDepartment of Chemistry, University of Warsaw, Pasteura 1, 02-093 Warsaw, Poland
cDepartment of Theoretical Chemistry and Amsterdam Center for Multiscale Modeling, Vrije Universiteit Amsterdam, De Boelelaan 1083, 1081 HV Amsterdam, The Netherlands. E-mail: c.fonsecaguerra@vu.nl
First published on 15th January 2016
The substituent effect of the amino and nitro groups on the electronic system of benzene has been investigated quantum chemically using quantitative Kohn–Sham molecular orbital theory and a corresponding energy decomposition analysis (EDA). The directionality of electrophilic substitution in aniline can accurately be explained with the amount of contribution of the 2pz orbitals on the unsubstituted carbon atoms to the highest occupied π orbital. For nitrobenzene, the molecular π orbitals cannot explain the regioselectivity of electrophilic substitution as there are two almost degenerate π orbitals with nearly the same 2pz contributions on the unsubstituted carbon atoms. The Voronoi deformation density analysis has been applied to aniline and nitrobenzene to obtain an insight into the charge rearrangements due to the substituent. This analysis method identified the orbitals involved in the C–N bond formation of the π system as the cause for the π charge accumulation at the ortho and para positions in the case of the NH2 group and the largest charge depletion at these same positions for the NO2 substituent. Furthermore, we showed that it is the repulsive interaction between the πHOMO of the phenyl radical and the πHOMO of the NH2 radical that is responsible for pushing up the πHOMO of aniline and therefore activating this π orbital of the phenyl ring towards electrophilic substitution.
In this work, we would like to study the substituent effects of NH2 and NO2 groups on benzene using Kohn–Sham density functional theory. As is known, these two substituents have completely opposite electronic properties and therefore, opposite effects on the electrophilic substitution reactions of benzene.1 The amino group is activating and ortho- and para-directing, whereas the nitro substituent is known to be deactivating and meta-directing. This difference in reactivity can be made plausible by the Lewis structures of aniline and nitrobenzene (Scheme 1), but it does not give an explanation of the influence of the substituent on the molecular σ and π orbitals located on the phenyl ring. The resonance structures suggest that the amino group increases the electron density only at ortho- and para-positions of the ring, making them more preferred to the electrophilic attack. In turn, the nitro group, according to the resonance structures, promotes the formation of a meta-product by decreasing the π-electron density at the ortho and para positions (Scheme 1). The field effect was deduced to be also important for the nitro substituent and to explain its deactivating properties.1a,6a In previous studies,6b–d some attempts were done to separate the total substituent effect of the nitro group into σ- and π-contributions. However, the outcome was not conclusive, in particular about an importance of the resonance effect, that is the directing effect, in nitrobenzene.
In this study, the resonance and inductive effects of the substituents on the benzene ring are analyzed and separated using Kohn–Sham Molecular Orbital (KS-MO) theory and the accompanying Energy Decomposition Analysis7 (EDA). The quantitative KS-MO model together with the EDA have proven to be able to elucidate the nature of different types of bonds: the nature of resonance-assisted hydrogen bonding8 and halogen bonding9 has been defined. Moreover, the KS-MO model was successfully used for explaining the organic reaction mechanism,10 the concept of aromaticity and other properties of chemical compounds.11 This method uses symmetry which allows separating the interactions into σ- and π-electronic systems of the benzene ring and the substituents. Previously, Fernandez et al.11d found a correlation between the Hammett constants and the π electronic energy term. We will also use the Voronoi Deformation Density (VDD) charge analysis, which enables us to track the reorganization in the σ and π electron densities of the phenyl ring due to the substituent and to determine the resonance and inductive effects on the ring. Based on frontier molecular orbital theory to explain chemical reactivity, our quantitative KS-MO results together with the VDD charges clarify the orientation of the electrophilic aromatic substitution and the change in reactivity of the benzene ring caused by the NH2 or NO2 substituents.
The interaction energy is examined in the substituted benzene in the framework of the KS-MO model using a quantitative energy decomposition analysis (EDA)7a into electrostatic interactions, Pauli repulsive orbital interactions and attractive orbital interactions:
ΔEint = ΔVelstat + ΔEPauli + ΔEoi | (1) |
The orbital interaction energy can be further decomposed into the contributions from each irreducible representation Γ of the interacting system (eqn (2)):7
ΔEoi = ΔEσ + ΔEπ | (2) |
(3) |
The VDD method also allows us to analyze the electronic redistributions within two polyatomic fragments when a chemical bond is formed between these two molecular fragments. Previously, we have shown that this method is most informative to analyze resonance-assisted hydrogen bonding appropriately as it allows decomposition of the total changes into σ and π electronic rearangements.8a,b The change in VDD atomic charges ΔQA is defined by eqn (4), which relates this quantity directly to the deformation density ρPh–NX2(r) − ρPh˙(r) − ρ˙NX2(r) associated with the formation of the overall molecule from the two fragments Ph˙ and ˙NX2 (with X = H or O).8a,b,14
(4) |
This functionality is extended by the decomposition into σ and π components (for the planar, Cs symmetric molecules):
(5) |
For the π system, this functionality is additionally extended by the decomposition of the electronic redistribution per atom ΔQA into a component associated with the Pauli repulsion ΔEPauli and a component associated with the bonding orbital interactions ΔEoi.
ΔQπA = ΔQπA,Pauli + ΔQπA,oi | (6) |
Δρπ(r) = ΔρπPauli(r) + Δρπoi(r) | (7) |
The change in π atomic charge caused by Pauli repulsion between the monomers in the complex is defined by eqn (8), and the corresponding change caused by charge transfer and polarization is given by eqn (9).
(8) |
(9) |
ΔQπA,Pauli = ΔQA2A,Pauli + ΔQB1A,Pauli | (10) |
ΔQπA,oi = ΔQA2A,oi + ΔQB1A,oi | (11) |
NX2 | d(CN)/Å | ΔVelstat | ΔEPauli | ΔEσa | ΔEA1 | ΔEB2 | ΔEπa | ΔEA2 | ΔEB1 | ΔEoi | ΔEint | |
---|---|---|---|---|---|---|---|---|---|---|---|---|
a ΔEσ = ΔEA1 + ΔEB2 and ΔEπ = ΔEA2 + ΔEB1. | ||||||||||||
˙NH2 | 0 | 1.389 | −160.0 | 288.5 | −247.0 | −239.5 | −7.5 | −25.1 | −0.2 | −24.9 | −272.1 | −143.6 |
˙NH2 | 90 | 1.389 | −165.5 | 292.3 | −249.5 | −238.0 | −11.5 | −11.0 | −0.1 | −9.2 | −260.5 | −133.7 |
˙NH2 | 0 | 1.492 | −124.2 | 211.0 | −208.9 | −204.1 | −4.8 | −17.3 | −0.1 | −17.2 | −226.2 | −139.4 |
˙NO2 | 0 | 1.492 | −161.2 | 330.0 | −224.4 | −217.7 | −6.7 | −18.5 | −0.7 | −17.8 | −242.9 | −74.1 |
˙NO2 | 90 | 1.492 | −159.5 | 330.3 | −229.6 | −219.9 | −9.7 | −11.0 | −0.7 | −9.1 | −239.6 | −68.8 |
˙NO2 | 0 | 1.389 | −211.6 | 445.0 | −276.0 | −265.9 | −10.1 | −27.9 | −1.0 | −26.9 | −303.9 | −70.5 |
The energy decomposition analysis of the electron-pair bond formation between Ph˙ and the substituent ˙NX2 shows that interaction energy for the formation of nitrobenzene is about half of the formation of aniline (−74.1 kcal mol−1 and −143.6 kcal mol−1, respectively) and the C–N bond is also shorter in aniline than in nitrobenzene (1.389 Å and 1.492 Å, respectively). The longer C–N distance in nitrobenzene is caused by the larger Pauli repulsion (330.0 kcal mol−1 for Ph–NO2 and 288.5 for Ph–NH2). Compression of the C–N distance in nitrobenzene to the distance in aniline makes the Pauli repulsion go up from 330.0 kcal mol−1 to 445.0 kcal mol−1, whereas in aniline the Pauli repulsion is only 288.5 kcal mol−1 at 1.389 Å. The shortening of the C–N distance in nitrobenzene results in an increase of 115.0 kcal of the Pauli repulsion and only in an extra attraction of −111.4 kcal mol−1 of the bonding components (electrostatic and orbital interaction). The σ MO diagram shows the repulsive interaction between the occupied σHOMO−1 of Ph˙ and σHOMO−1 of ˙NO2, which is absent in the σ MO diagram of aniline (see Fig. 1 and 2). The largest bonding contribution to the interaction energy is the A1 component of the orbital interaction (that is the formation of the electron-pair bond), which amounts to −239.5 kcal mol−1 for aniline and −217.7 kcal mol−1 for nitrobenzene. The electrostatic interaction is of similar strength for both systems: −160.0 kcal mol−1 for aniline and −161.2 kcal mol−1 for nitrobenzene. Note that the ΔEA2 and ΔEB2 only account for polarization within each fragment as the A2 and B2 orbitals have no amplitude on the atoms located on the C2 rotation axis and therefore do not overlap.
The stabilization caused by the donor–acceptor interactions in the B1 irreducible representation (that is in the π electronic system) is larger for aniline (−24.9 kcal mol−1) than for nitrobenzene (−17.8 kcal mol−1). Compression of the C–N bond in Ph–NO2 to the distance of Ph–NH2 enlarges the ΔEB1 to −26.9 kcal mol−1.
For both substituents, NH2 and NO2, the formation of the electron-pair bond in the σ electronic system causes the singly occupied molecular orbital (SOMO) of the phenyl radical to donate 0.2 electrons to the SOMO of the substituent accounting for the inductive effect of the substituent. In the case of nitrobenzene, the bond formation is accompanied by the Pauli repulsive interaction between occupied orbitals on the Ph˙ and ˙NO2 fragments (see Fig. 2).
The donor–acceptor interactions in the π electronic system of aniline and nitrobenzene have an opposite effect on the phenyl radical: NH2 is electron donating and NO2 is electron-withdrawing. The MO diagrams illustrate this clearly. For aniline, the π-orbital of the NH2 group donates 0.17 electrons to the πLUMO+1 orbital of the benzene ring (see Fig. 1). This counteracts the charge-flow in the σ system. There is also Pauli repulsion between the πHOMO on Ph˙ and ˙NH2, which pushes up the energy of the πHOMO of aniline to −4.6 eV and accounts for the activation towards electrophilic substitution and not the donor–acceptor interactions. This mechanism has also been encountered in the study of the substituent effects on the optical properties of naphthalene diimides.11c
In the case of nitrobenzene, the electron transfer goes in the opposite direction. The πLUMO of the NO2 group accepts 0.11 electrons from the π-system of the benzene ring increasing the deactivation for electrophilic substitution (see Fig. 2). The πHOMO of nitrobenzene lies at −7.0 eV, which is 2.4 eV lower than the same orbital in aniline and explains the lower reactivity of nitrobenzene.
Another way to investigate the effect of the substituents is to rotate the substituents by 90 degrees, that is, perpendicular to the plane of the benzene ring and analyze the capability of the substituents to participate in the π-electron delocalization of the ring. The donor–acceptor interactions are either lost as in the nitrobenzene or become negligible as in aniline (the amino group donates only 0.03 electrons, see Fig. S1 and S2, ESI†). Rotation of the substituent NX2 (the rotation barriers are given in Table S1, ESI†) switches off the donor–acceptor interactions in the π electronic system, however the energetic loss is quite small in both cases. For nitrobenzene, ΔEπ goes from −18.5 kcal mol−1 to −11.0 kcal mol−1 and for aniline from −25.1 kcal mol−1 to −11.0 kcal mol−1. Clearly, the π orbitals of the phenyl ring can also interact with the perpendicular substituent and the π delocalization is therefore not switched off completely (see Fig. S1 and S2 in the ESI,† for the MO diagrams for the perpendicular conformation).
After this explanation of the activating and deactivating effects of the NH2 and NO2 substituents on the benzene ring based on the MO diagrams, we would also like to understand the directionality of the electrophilic attack. Recently, Fievez et al.10c showed that the orbital interaction is the driving force towards selectivity in electrophilic aromatic substitution for activating substituents, but it is not the main factor for the NO2 substituent. They studied the interaction energy and its components (eqn (1)) between monosubstituted benzene derivatives and a model electrophile at the onset of the reaction in a plane parallel to the molecular plane.
Here, we would like to analyze the molecular orbitals of the substituted benzenes in terms of the contribution of the 2pz orbitals on the carbon, nitrogen and oxygen atoms to the molecular π orbitals, allowing us to explain the directionality from the molecular orbital perspective (see Fig. 3). The high contribution of a 2pz orbital on a certain carbon atom to the πHOMO of the substituted benzene will direct the electrophilic attack towards that atom because of the larger overlap between the accepting orbital of the electrophile and the πHOMO of the substituted benzene. The most favorable interaction occurs at the atom with the highest 2pz contribution to the πHOMO.
Aniline has only one highest occupied π orbital at −4.6 eV with the largest gross Mulliken contributions on the meta (14%) and the para (22%) positions. The electrophilic attack will therefore occur at these positions. Nitrobenzene has two almost degenerate highest occupied π orbitals at −6.9 eV and −7.0 eV. The πHOMO directs the electrophilic attack towards ortho and meta positions with 2pz contributions of 23% on these two positions, and the πHOMO−1 towards the para position with a 2pz contribution of 31% on the para (contributions on the ortho and meta positions are small, 5% and 8% respectively). Thus, the π orbitals of nitrobenzene are deactivated (because they are low-lying) and the experimentally observed meta directing effect of the nitro substituent for electrophilic aromatic substitution cannot be explained from the character of the πHOMO and πHOMO−1.
Fig. 4 VDD atomic charges (milli-electrons) of benzene, aniline and nitrobenzene (eqn (3)). |
The VDD charge rearrangements, as defined in eqn (4), allows analysis of the charge redistributions within two polyatomic fragments when a chemical bond is formed between these two molecular fragments. Furthermore, it allows decomposition of the total changes into the components of different irreducible representations. In Fig. 5, the VDD charge rearrangements (ΔQ) due to the formation of the C–N bond in aniline and nitrobenzene are given. The bond formation causes the phenyl ring of aniline to be activated with the largest electronic charge accumulations on the ortho and para positions, and the phenyl ring of nitrobenzene to be deactivated with the smallest electronic charge depletions at the meta positions. Decomposition into σ and π charge rearrangements, shown in Fig. 5, reveals that the π electronic system, and thus the π bond formation, is responsible for the regioselectivity of electrophilic substitution, which is in line with previous research.6c Negative numbers on the atoms mean an accumulation of π-electron density on those atoms and positive numbers mean a depletion of π-electron density at atoms. The π C–N bond formation in aniline causes ΔQπ to become more negative at the ortho and para positions (−50 and −38 milli-electrons) experimentally leading to a mixture of ortho and para products, and in nitrobenzene to become more positive at all positions, but less positive at the meta position (only 13 millielectrons are lost at the meta position), thus the nitro group is a deactivating meta-directing substituent.15,16 The VDD electronic rearrangements are in line with the experimentally observed regioselectivity of electrophilic aromatic substitution. To determine which factor is more important, orbital interactions or charge redistribution, an additional investigation of the transition state of an electrophile with the substituted benzene is necessary.
Moreover, the VDD charge rearrangements allow for decomposition into the charge rearrangements caused by the Pauli repulsion and orbital interactions. As aniline and nitrobenzene are C2v symmetric molecules, we decomposed the π charge rearrangements into the A2 and B1 contributions. Only the B1 charge rearrangements are caused by the C–N bond formation, whereas the A2 charge rearrangements are only due to polarization (no A2 orbital has amplitude on the atoms lying on the C2 symmetry axis). Fig. 6 and 7 clearly show that the directionality of the electrophilic substitution is caused by the B1 charge rearrangements in the orbital interactions, ΔQB1oi. The largest charge accumulations of ΔQB1oi are at the ortho and para positions for aniline and the largest charge depletions of ΔQB1oi are at the ortho and para positions for nitrobenzene. For aniline, this finding confirms our results from the molecular orbital analysis, whereas for nitrobenzene the VDD charge rearrangements are essential to pinpoint the orbitals responsible for the directionality, namely the B1 orbitals are responsible for the C–N bond formation.
The donor–acceptor interactions from the πHOMO of ˙NH2 to the πLUMO of the phenyl radical for the C–N bond formation in aniline are accompanied by the repulsive interaction between the two π HOMOs on the two fragments. It is this repulsive interaction between the occupied orbitals, which pushes up the πHOMO on the phenyl ring of aniline and activates it to electrophilic substitution. The directionality of aniline can be understood from the 2pz contributions (located on the unsubstituted carbon atoms) to the πHOMO of aniline, which have the largest amplitude at the ortho and para positions.
In the case of nitrobenzene, the donor–acceptor interactions lead to a lowering of the πHOMO on the phenyl ring, which deactivates this orbital. Further inspection of the πHOMO and πHOMO−1 orbitals does not explain the directionality of nitrobenzene clearly. It is the VDD charge analysis which shows that the π electronic rearrangements due to C–N bond formation in nitrobenzene are responsible for the largest deactivation at the ortho and para positions of nitrobenzene, making the nitro group meta directing for electrophilic aromatic substitution.
Footnote |
† Electronic supplementary information (ESI) available: MO diagrams for nitrobenzene and aniline in a perpendicular orientation; interaction energies and rotational barriers for substituted benzenes; and Cartesian coordinates of nitrobenzene and aniline at the BLYP/TZ2P level. See DOI: 10.1039/c5cp07483e |
This journal is © the Owner Societies 2016 |