Saeedreza Emamian*a,
Tian Lub and
Farid Moeinpourc
aChemistry Department, Shahrood Branch, Islamic Azad University, Shahrood, Iran. E-mail: s_emamian@iau-shahrood.ac.ir; saeedreza_em@yahoo.com; Fax: +98 23 32390537; Tel: +98 91 21735085
bBeijing Kein Research Center for Natural Sciences, Beijing 100022, P. R. China
cDepartment of Chemistry, Bandar Abbas Branch, Islamic Azad University, Bandar Abbas, Iran
First published on 9th July 2015
The [3 + 2] cycloaddition reaction of the simplest azomethine betaines, AZBs, as a class of three-atom-components, TACs, toward ethylene was theoretically studied at the DFT-B3LYP/6-31G(d) level. The high reactivity of AZBs in the studied reactions is reflected by the corresponding relatively low activation total electronic energies calculated at the very accurate MP4(SDTQ)/6-311++G(d,p) level. The singlet-diradical character of AZBs was estimated according to the two different descriptors based on the HF symmetry-broken approach and also by using the natural resonance theory (NRT). The high linear correlation coefficients found between the singlet-diradical character of AZBs and corresponding activation energies clearly show that the singlet-diradical character is responsible for the high reactivity of AZBs. An assessment of molecular orbital shapes in AZBs indicates that delocalization of the central nitrogen lone pair into the Pz atomic orbitals of terminal atoms accounts for the singlet-diradical character in AZBs. Our results also imply that the singlet-diradical characters are inversely proportional with the electronegativity of heteroatoms in AZBs. The ground electronic density transfer (GEDT), calculated at the transition state geometries using four different population schemes, predicts a low polar character for the investigated reactions which is in excellent agreement with the low difference in the global electrophilicity index of the reagents. Furthermore, the molecular mechanism involved in the studied reactions was characterized with the aid of ELF topological analysis supporting a non-concerted one-step mechanism; not a pericyclic concerted one as is provided in most textbooks.
The electronic structure of an A-TAC can generally be represented by five Lewis structures I–V, as given in Scheme 2.
As shown in Scheme 2, Lewis structures I through IV have closed-shell zwitterionic electronic structures, while Lewis structure V has an open-shell singlet-diradical electronic structure. It is worth noting that Lewis structures III and IV with a very low contribution among structures I through V (see later) are unique 1,3-dipoles.2 Moreover, it has been well-recognized that the high reactivity of A-TACs toward unsaturated bonds is a direct consequence of their singlet-diradical character; the higher the singlet-diradical character, the more reactive is the TAC.2,5 Consequently, it is clear why the “1,3-dipole” phrase is meaningless to describe this class of compounds when they participate in a cycloaddition reaction. It is also interesting to know more about the meaning of the “singlet-diradical” character exhibited by some compounds in chemistry. A singlet-diradical system is defined as a molecular arrangement in which all electrons are paired, while a weakly coupled pair of these electrons occupy different parts of space with a small sharing region.6 The occupation arrangement of molecular orbitals for a hypothetical compound displaying a singlet-diradical character is depicted in Scheme 3.
![]() | ||
Scheme 3 A schematic representation for the occupation arrangement of molecular orbitals in a hypothetical species with singlet-diradical character. |
Owing to the special importance of singlet-diradical character in TACs, having a simple and, in particular, straightforward theoretical methodology by which such character can easily be estimated has its own unique attraction. With the aid of the “breathing-orbital valence bond (BOVB)” method, Braida computed the weight of singlet-diradical structure V (see Scheme 2) at the resonance hybrid in a given series of TACs involved in [3 + 2] cycloadditions toward ethylene and acetylene.5 Unfortunately, the method used by Braida is excessively complex. Instead, the method proposed by Bachler6 is much simpler to use to evaluate singlet-diradical character in a given species and can be performed using a very familiar GAUSSIAN software package. In this method, which is also known as the HF and/or DFT symmetry-broken approach, the restricted wave function for a given system is initially checked to find a feasible instability. If an instability is found, the symmetry of restricted wave function is broken and a new wave function is rebuilt by a mix of restricted closed-shell and unrestricted open-shell wave functions. In this way, the magnitude of the total spin operator, 〈S2〉, of the newly rebuilt wave function can be considered as the tendency of the system under study to behave as a singlet-diradical species. With the value of 〈S2〉 at hand, the singlet-diradical percentage can easily be calculated using the following equation:
![]() | (1) |
Additionally, based on the symmetry-broken approach, there is another singlet-diradical descriptor index which is formally expressed in the case of the spin-projected UHF (PUHF) theory.7 This descriptor provides the singlet-diradical percentage as %Y defined by the following equation:
![]() | (2) |
![]() | (3) |
In order to explain the reactivity of TACs in [3 + 2] cycloaddition reactions, a distortion/interaction based model has recently been introduced by Ess and Houk.8 According to this model, activation energy is considered as a sum of two different factors; distortion energy and interaction energy. While the former indicates the required energy to distort reagents from their ground state geometries to the corresponding geometries in TS, the latter implies the energy corresponding to the interaction between two fragments in TS. The relationship between activation, distortion, and interaction energies is depicted in Scheme 4. The applicability of this mode was examined in [3 + 2] cycloaddition reactions of different P- and A-TACs toward ethylene and acetylene, they found a good correlation coefficient, R2 = 0.97, between B3LYP/6-31G(d) activation enthalpies and distortion energies in the studied reactions. They also concluded that distortion energy, which is mainly associated with the distortion of TACs, plays a key role in the reactivity differences of the considered TACs. It is worth noting that the model proposed by Ess and Houk is not able to address why the activation energy depends on the geometry or vice versa.2
![]() | ||
Scheme 4 A schematic relationship between activation, ΔE≠, distortion, ΔE≠d, and interaction, ΔE≠i, energies in the [3 + 2] cycloaddition reaction of different TACs toward ethylene and acetylene studied by Ess and Houk.8 |
In this work, [3 + 2] cycloaddition reaction of a class of A-TACs, namely azomethine betaines (AZBs), toward ethylene is theoretically taken into account in order to address the following questions:
(i) Can the high reactivity of AZBs in the [3 + 2] cycloaddition reaction toward ethylene be well explained using the singlet-diradical character descriptor indices introduced in eqn (1) and (2)? If so, a very good linear correlation is expected when the corresponding activation energies are plotted against these indices.
(ii) What is the source of singlet-diradical character in AZBs?
(iii) What molecular mechanism is involved in the studied reactions? In other words, do these [3 + 2] cycloadditions take place via a pericyclic concerted mechanism as mentioned in the textbooks and as commonly used among chemists?
Herein, the very accurate MP4(SDTQ)/6-311++G(d,p) level is employed in order to obtain activation energies corresponding to the [3 + 2] cycloaddition reactions presented in Scheme 5. Singlet-diradical character descriptor indices are calculated using the HF symmetry-broken approach, and also by using the natural resonance theory (NRT) to find, if any, a good linear correlation between the activation energies and the previously mentioned indices. Furthermore, an ELF topological analysis along the intrinsic reaction coordinate, IRC, allows the molecular mechanism involved in the studied [3 + 2] cycloadditions to be clarified in detail.
Reaction | ΔE≠1 | ΔE≠2 |
---|---|---|
AZB1 (X = CH2) + ethylene | −2.2 | 1.03 |
AZB2 (X = NH) + ethylene | 4.56 | 7.81 |
AZB3 (X = O) + ethylene | 10.36 | 12.36 |
It is worth mentioning that while the negative activation total electronic energy in the [3 + 2] cycloaddition reaction of AZB1 toward ethylene could be related to the greater reactivity of AZB1 compared with the other ones, it is more usual to obtain activation energies in positive values, as comes from the meaning of the phrase “activation energy”.
The negative activation energy in the reaction of AZB1 toward ethylene will become positive if the formation of a pre-complex between reactants is taken into account. Indeed, when the reactants approach, the potential energy is reduced (shifts into more negative values) due to the formation of a weak van der Waals pre-complex between the reactants in a very early stage of the reaction channel (see Scheme 6). As can be seen, the advantage of considering the formation of a pre-complex is to obtain a positive value for the activation energy.
For each [3 + 2] cycloaddition reaction in this work, the pre-complex can structurally be found at the first point of the corresponding IRC curve toward a TS. Taking into account a pre-complex formation, the calculated activation energies will change to 1.03, 7.81, and 12.36 kcal mol−1 for the [3 + 2] cycloaddition reactions of AZB1, AZB2, and AZB3 toward ethylene, respectively (see third column in Table 1). These relatively low activation energies, especially for AZB1, clearly indicate the very high reactivity of AZBs toward ethylene in a [3 + 2] cycloaddition reaction, so that such cyclization can easily take place at room temperature with an acceptable rate.
Values of ΔE1, presented in the second column of Table 2, are related to the quantity named “energy lowering” gained via the symmetry-broken approach which is defined as the difference between the HF symmetry-broken solution, E(b), and the restricted HF solution, E(S0). This quantity, in fact, implies the tendency of a species to exhibit a singlet-diradical character; i.e., the more negative ΔE1, the higher the tendency to provide a singlet-diradical character. The third column represents the values of the total spin operator, 〈S2〉, resulting from the last optimization step of the symmetry-broken process. Additionally, the fourth through seventh columns include percentage of singlet-diradical character, %nrad, calculated based on eqn (1), the HOMO and LUMO occupation numbers, nHOMO and nLUMO, and percentage of singlet-diradical character, %Y, calculated based on eqn (2). In the last column of Table 2, the values of the singlet-triplet energy gap, ΔE2, defined as the difference between the first high spin triplet state solution, Eu(T1), and the HF symmetry broken solution, E(b), are collected. Values of E(S0), E(b), and Eu(T1) are given in Table 3. According to the values collected in Table 2, a greater singlet-diradical character is predicted by the descriptor index defined as %nrad in comparison with that defined as %Y for the considered AZBs. Nevertheless, the singlet-diradical character of AZBs based on both the %nrad and %Y descriptors follows the trend of AZB1 (X = CH2) > AZB2 (X = NH) > AZB3 (X = O). In other words, the singlet-diradical character is inversely proportional with the electronegativity of X, which is in nice agreement with the calculated activation energies of the [3 + 2] cycloaddition of AZBs toward ethylene, AZB3 (X = O) > AZB2 (X = NH) > AZB1 (X = CH2).
AZB | E(S0)/a.u. | E(b)/a.u. | Eu(T1)/a.u. |
---|---|---|---|
AZB1 (X = CH2) | −132.97704916 | −133.00082440 | −132.98795454 |
AZB2 (X = NH) | −148.97306185 | −148.99093858 | −148.97271276 |
AZB3 (X = O) | −168.80920648 | −168.81886549 | −168.79028776 |
The inverse proportionality of singlet-diradical character with the electronegativity of atom X may be explained considering the fact that by increasing the electronegativity of atom X, the capability of the system to bear an unpaired electron on the more electronegative atom X is considerably reduced. Consequently, the stability of system significantly decreases as well. Furthermore, as expected, the singlet-triplet energy gaps, ΔE2, are reduced with the increase of singlet-diradical character of AZBs. A linear regression between indices describing singlet-diradical character of AZBs and the corresponding activation energies of [3 + 2] cycloadditions between AZBs and ethylene is presented in Fig. 2.
![]() | ||
Fig. 2 Linear regression between the singlet-diradical character descriptor indices and activation energies of the [3 + 2] cycloaddition reactions of AZBs toward ethylene. |
The high linear correlation coefficients (0.9693 and 0.9996 in the case of %nrad and %Y, respectively) clearly indicate that the high reactivity of AZBs can directly be explained by their singlet-diradical character. On the other hand, the higher correlation coefficient resulting from the linear regression of %Y with activation energies implies that, in the case of the studied AZBs, %Y is a better and more reliable index than %nrad to describe corresponding singlet-diradical character. As mentioned in the introduction, Braida has also calculated the weight of singlet-diradical structure for the studied AZBs using the BOVB method;5 the corresponding values are 41.30, 38.00, and 33.70% for AZB1–3, respectively. Interestingly, while the observed trend of singlet-diradical character, obtained via the excessively complex BOVB method, is in line with those obtained via the indices defined in this work, when singlet-diradical character values resulting from the BOVB method are plotted against %Y and %nrad indices, very high linear correlation coefficients (0.9998 and 0.9565 in the case of %nrad and %Y, respectively) are obtained as shown in Fig. 3.
![]() | ||
Fig. 3 Linear regression between the singlet-diradical character descriptor indices, defined in this work, and that obtained by Braida using the BOVB method for the studied AZBs. |
As shown in this figure, the singlet-diradical character descriptor defined based on eqn (1), %nrad, is much better correlated with that obtained by Braida, indicating more similarity in the basis used in the BOVB method and the symmetry-broken approach. Moreover, by using the natural resonance theory (NRT keyword) in Gen NBO5.0W software,23 various resonance structures and corresponding contributions can nicely be mapped for a given species. Table 4 shows the percentage of resonance structures I–V for AZBs obtained via the “NRT” keyword.
As collected in Table 4, the percentage of singlet-diradical character of AZBs follows the trend of AZB1 (23.93%) > AZB2 (20.62%) > AZB3 (17.20%) in complete agreement with the trend provided by the %nrad and %Y descriptors. However, as mentioned in the introduction, 1,3-dipole structures have a very low contribution among resonance structures (see the last column in Table 4). Interestingly, when the singlet-diradical percentage of AZBs obtained from natural resonance theory is correlated with those obtained via %nrad and %Y descriptors, as displayed in Fig. 4, very good linear correlation coefficients are obtained, especially in the case of the %nrad descriptor, indicating that natural resonance theory can also be taken into account as a powerful and helpful tool to estimate the amount of singlet-diradical character in the corresponding species.
![]() | ||
Fig. 4 Linear regression of singlet-diradical descriptor indices, %nrad and %Y, with the singlet-diradical percentage obtained via natural resonance theory for AZBs. |
![]() | ||
Fig. 5 HOMO and HOMO–n MO shapes of the studied AZBs including Pz atomic orbital coefficients of terminal atoms in HOMO MO, calculated at the HF/STO-3G//B3LYP/6-31G(d) level. |
As shown in Fig. 5, while the HOMO in all AZBs corresponds to the non-bonding n MO with a nodal plane at the central nitrogen atom, HOMO–2 in the case of AZB1 and HOMO–3 in the case of AZB2 and AZB3 are associated with the bonding π MO extended on the C–N–X (X = CH2, NH, and O) three-atom system. The electron density of HOMO MO is distributed over the Pz orbitals of two terminal atoms. The shape of the non-bonding n MO can be related with the diradical structure V in Scheme 1, whereas the π MO can be related with a resonant structure between I and II in which the nitrogen lone pair is delocalized over two adjacent carbon–carbon (in AZB1), carbon–nitrogen (in AZB2), and carbon–oxygen (in AZB3) atoms. Consequently, delocalization of the central nitrogen lone pair over two terminal atoms in AZBs is responsible for the singlet-diradical character in these TACs. The values of calculated Pz atomic orbital coefficients indicate that with the increase of the electronegativity of atom X, the Pz atomic orbital coefficient of the terminal carbon atom decreases, −0.7155 (AZB1, X = CH2) > −0.6727 (AZB2, X = NH) > −0.6521 (AZB3, X = O) which is in excellent agreement with the singlet-diradical character of AZBs. On the other hand, the Wiberg22 bond order (BO) of the C–N bond in AZBs1–3 is 1.34, 1.44, and 1.56, respectively. These values obviously indicate that the electronegativity of atom X plays a main role in the amount of singlet-diradical character of AZBs; that is, more electronegativity leads to an increase of the π delocalization in the HOMO−n MO (see Fig. 5). In consequence, the π character of the C–N bond in AZBs increases which, in turn, results in an increase of the corresponding C–N BO. In other words, the weight of resonant structures I and II (see Scheme 2) will increase and, hence, the weight of singlet-diradical structure V is shifted toward lower values. In Fig. 6, the linear regression of singlet-diradical descriptor indices, %nrad, %Y, with the Pz atomic orbital coefficient of the terminal carbon atom of AZBs is plotted.
![]() | ||
Fig. 6 Linear regression of singlet-diradical descriptor indices, %nrad and %Y, with the Pz atomic orbital coefficient of the terminal carbon atom of AZBs. |
As can be seen, the Pz atomic orbital coefficients of the terminal carbon atom in AZBs are much better correlated with the %Y descriptor than %nrad, indicating that, in the investigated AZBs, the Pz atomic orbital coefficient of terminal carbon atoms is directly proportional with the singlet-diradical character.
Species | μ | η | ω |
---|---|---|---|
AZB1 | −1.81 | 4.47 | 0.36 |
AZB2 | −2.69 | 5.01 | 0.72 |
AZB3 | −3.43 | 5.54 | 1.06 |
Ethylene | −3.37 | 7.76 | 0.73 |
The absolute values of difference in the electrophilicity index, Δω, of AZBs and ethylene are very low, varying in the narrow range of 0.01 to 0.34 eV, implying the very low polar character of the corresponding [3 + 2] cycloadditions. On the other hand, the electronic chemical potentials of AZB1, −1.81 eV, and AZB2, −2.69 eV, are higher than that of ethylene, −3.37 eV, whereas the electronic chemical potential of AZB3, −3.43 eV, is lower than that of ethylene. Therefore, while AZBs1 and 2 will act as nucleophiles, AZB3 acts as an electrophile in the [3 + 2] cycloaddition reaction toward ethylene. Consequently, along the very low polar cycloadditions, the global electron density transfer (GEDT) will take place from AZBs1 and 2 toward ethylene while in the case of AZB3, GEDT should take place in the reverse direction. The values of GEDT in the studied [3 + 2] cycloaddition reactions are calculated through various population schemes and collected in Table 6.
The values of GEDT presented in Table 6, regardless of the population scheme used, are very low, in excellent agreement with the slight polarity of the studied [3 + 2] cycloadditions as expected considering the very low values of Δω. On the other hand, among the employed population schemes, only the charges derived from the electrostatic potential, MK and ChelpG, are capable of providing a correct direction for GEDT in excellent agreement with the values of electronic chemical potentials of reactants (see Table 5). Interestingly, when GEDT values calculated via the employed population schemes are plotted against activation energies, as displayed in Fig. 7, the best linear correlation coefficients are associated with the populations based on the electrostatic potentials; i.e., the MK and ChelpG schemes show correlation coefficients of 0.9722 and 0.9936, respectively.
From Fig. 7, one can easily conclude that the ChelpG population scheme is the best scheme to describe the polar nature of the studied [3 + 2] cycloadditions both qualitatively and quantitatively. Moreover, closer attention to Fig. 7 shows that while the activation energies vary in a relatively wide range of 1.03 to 12.36 kcal mol−1, the absolute values of GEDT vary only in a narrow range of 0.008 to 0.115 e based on the ChelpG population analysis. In other words, GEDT values do not increase as fast as activation energies decrease. This, in fact, is a characteristic of pr-type [3 + 2] cycloaddition reactions in which, unlike Diels–Alder cycloadditions, GEDT does not act as a driving force. Such behavior obviously emphasizes that the singlet-diradical character of TAC is the driving force and acts as a determining factor in pr-type [3 + 2] cycloadditions.2
![]() | ||
Fig. 8 B3LYP/6-31G(d) IRC profiles including the ELF attractor positions for the most relevant points of the [3 + 2] cycloaddition of AZB1 (top), AZB2 (middle), and AZB3 (bottom) toward ethylene. |
Some appealing conclusions can be pointed out considering the IRC profiles and ELF analysis: (i) all studied [3 + 2] cycloadditions proceed via a non-concerted one step mechanism; (ii) formation of both C1–C4 and C5–X3 single bonds happen after passing the corresponding TS in all studied [3 + 2] cycloaddition reactions; (iii) formation of the first C1–C4 single bond take place at the distance of 2.052, 1.977, and 1.949 Å for [3 + 2] cyclization of AZB1, AZB2, and AZB3, respectively, toward ethylene through the coupling of two pseudoradical centers created on the interacting C1 and C4 carbon atoms. As can be seen, the increase of electronegativity of atom X from AZB1 (X = C) to AZB2 (X = N) to AZB3 (X = O) leads to a slight decrease in the distance of C1–C4 single bond formation which is in agreement with the decrease in Pz atomic orbital coefficient of C1; (iv) while the formation of a second C5–X3 single bond in the case of AZB1 and AZB2 takes place at the distance of 2.052 and 1.848 Å, respectively, via the coupling of two pseudoradical centers created on the interacting C5 and X3 atoms, a quite different pattern is observed in the case of AZB3. Indeed, in the [3 + 2] cycloaddition of AZB3 toward ethylene the second C5–X3 (X = O) single bond formation takes place at a distance of 1.718 Å in which the oxygen atom of AZB3 is nucleophilically attacked by the C5 carbon atom of ethylene. As depicted in Fig. 9, the “hole” around the oxygen atom is associated with an electron deficient region making the oxygen atom susceptible to nucleophilic attack. In other words, the ELF topological analysis for the [3 + 2] cyclization of AZB3 toward ethylene clearly indicates that the oxygen atom of AZB3 does not directly participate in the formation of a C–O single bond via sharing its valence electrons density as a pseudoradical center and, instead, prefers to be attacked by a nucleophilic center.
![]() | ||
Fig. 9 ELF isosurface map with the isovalue of 0.825 corresponding to point P7 located on the IRC profile of the [3 + 2] cycloaddition between AZB3 and ethylene. |
Footnote |
† Electronic supplementary information (ESI) available: Details of ELF topological analysis of bonding changes along the studied [3 + 2] cycloaddition reactions. Cartesian coordinates of the B3LYP/6-31G(d) optimized structures of species involved in the studied [3 + 2] cyclization reactions. See DOI: 10.1039/c5ra08614k |
This journal is © The Royal Society of Chemistry 2015 |