Insight into structural and π–magnesium bonding characteristics of the X2Mg⋯Y (X = H, F; Y = C2H2, C2H4 and C6H6) complexes

Si-Yi Li, Di Wu, Ying Li*, Dan Yu, Jia-Yuan Liu and Zhi-Ru Li
Institute of Theoretical Chemistry, Jilin University, Changchun 130023, P. R. China. E-mail: liyingedu@jlu.edu.cn

Received 20th September 2016 , Accepted 24th October 2016

First published on 24th October 2016


Abstract

Quantum chemical calculations have been performed to study the nature of interaction of complexes formed by MgX2 (X = H, F) molecules with acetylene, ethylene, and benzene, respectively. Results indicate that nonbonded interactions, namely π–magnesium bonds, contribute to the stability of the resulting dimers. An intriguing structural evolution has been found for the complexes with different π electron donors. Upon complexation, both the Mg–X and C–C bonds lengthened, accompanied by redshifted X–Mg–X and C–C stretching vibrations. NBO analysis reveals that the main charge-transfer from the conjugated molecules to MgX2 is from the πCC bonding orbital to the empty lone pair orbital of Mg. Energy decomposition analysis indicates that the stability of the topic complexes mainly comes from the attractive electrostatic interaction and polarization, similar to the case of π–beryllium bonds. When compared with other nonbonded interactions, it is found that π–magnesium bonds are stronger than π–lithium and π–sodium bonds. Especially, they are comparable in strength to π–beryllium bonds with MgF2 playing the role of Lewis acid.


Introduction

Noncovalent interactions play a key role in chemistry and biochemistry research,1–4 especially as far as molecular recognition5,6 and molecular medicine7,8 are concerned. In recent years, numerous experimental and theoretical endeavors have been devoted to understanding the origin and strength of such interactions. Hydrogen bond is one of the most crucial and common noncovalent interactions. The formation of hydrogen bond can be depicted as X–H⋯Y where X–H represents the hydrogen bond donor and the hydrogen acceptor Y may be an atom or an anion, or a fragment, or a molecule. In 1971, Morokuma9 reported the first theoretical evidence of a π–hydrogen bonding complex H2CO⋯2H2O in which the C[double bond, length as m-dash]O group acts as the proton acceptor. As an analog of hydrogen bonds, the existence of lithium bond was theoretically predicted in 197010 and experimentally confirmed by Ault and Pimentel11 in 1975. During the past decades, lithium bond has been identified in a variety of systems, and many non-traditional lithium bonding interactions have been put forward, such as π–lithium bond.12 In the same vein, the concept of π–sodium bond has been proposed lately.13 Note that a similar interaction in the X3M⋯π (M = B, Al; X = H, Cl, Br) complexes was also elaborated by Grabowski.14

Most recently, Yañez et al.15 have demonstrated that there is an interaction between BeX2 (X = H, F, Cl, OH) and different Lewis bases, namely NH3, H2O, FH, PH3, SH2, ClH and BrH. They examined this special kind of weak interactions and defined it as “beryllium bonds” in a similar way as the definition of hydrogen bonds and lithium bonds. Subsequently, the concept of beryllium bond attracted much attention and research interest. Eskandari explored the nature of beryllium bonds by a QTAIM study.16 The investigation of interaction between squaric acid and BeH2 by Montero-Campillo and coworkers17 has revealed significant cooperative effects between the beryllium bond and the BeH⋯HO dihydrogen bond. Yañez's group18 focused on the ternary BeR2:C6X6:Y complexes (R = H, F and Cl, X = H and F, and Y = Cl and Br) and provided clear evidence for favorable cooperativity in beryllium bond and aromatic ring:anion interaction. They also proposed the existence of π–beryllium bonds between ethylene and acetylene and BeX2 (X = H, F, and Cl) derivatives, and found that the π–beryllium bonds are about four times stronger than π–hydrogen bonds.19 In a very recent paper, Yu et al. theoretically characterized a kind of beryllium bonding complexes with radicals playing the role of Lewis base.20

The extensive studies of beryllium bonds, as well as the similarity between beryllium and magnesium elements, naturally lead to a question: does magnesium bond, as a congener of beryllium bond, exist? The answer is already known. In a very recent work, Yang and co-workers have demonstrated that there is another weak interaction other than lithium bond between LiNH2 and MgH2 molecules, and consequently proposed the existence of magnesium bond.21 Compared with the well-studied beryllium bonds, the systematic study of magnesium bonds is of additional significance. On the one hand, magnesium-containing complexes are nontoxic, which facilitates the experimental analysis. On the other hand, the element of magnesium widely exists in organisms and plays an essential role in a variety of biochemical processes. For instance, magnesium ions are important in the formation of DNA origami structures22 and affect their conductivity as well.23 Hence, the interactions between magnesium derivatives and conjugated systems are of great research value.

Here, we focused on the formation of π–magnesium bond and systematically studied the interaction between conjugated molecules (C2H2, C2H4 and C6H6) and MgX2 (X = H, F) molecule by using quantum chemical methods. Herein, ethylene, acetylene and benzene have played the role of π electron donors in the previous investigations of π–hydrogen,24 π–lithium,12 and π–halogen bonds.25 The main objectives of this paper are (1) to understand how the MgX2 and conjugated molecules respond to each other and explore the origin of π–magnesium bond, (2) to reveal different geometrical features of the X2Mg⋯C2H2, X2Mg⋯C2H4 and X2Mg⋯C6H6 complexes, (3) to compare the strength of π–magnesium bond with other nonbonding interactions concerned, including conventional magnesium bonds, π–beryllium bonds, π–lithium bonds and π–sodium bonds, and (4) to analyze the relationship between π–magnesium bonding strength and different π electron donors. We hope that the results from this work not only provide reference to the researches in the field of biochemistry and physiology in view of the important role that magnesium element plays therein, but also help to understand the interaction between conjugated π–systems and other metal atoms, for example the behavior of transition metal atoms adsorbed on graphene or graphene oxide sheets, etc.26–28

Computational details

The optimized structures of the X2Mg⋯C2H2, X2Mg⋯C2H4 and X2Mg⋯C6H6 (X = H, F) complexes were obtained at the MP2/aug-cc-pVTZ level by using the counterpoise procedure.29,30 Then the vibrational frequency calculations were performed to ascertain that these geometries are true minima on the potential energy surface. Intermolecular interaction energies (Eint) of the complexes were calculated using the coupled cluster theory with single and double excitations and perturbative contribution of connected triple excitations [CCSD(T)] method with the aug-cc-pVTZ basis set. The Eint values have been calculated as the difference between the energy of the complex (EAB) and the sum of the energy of the monomer subunits (EA, EB), by the following formula,
 
Eint = EAB(MAB) − EA(MAB) − EB(MAB) (1)

To eliminate the basis set superposition error (BSSE) effect in the interaction energy given by eqn (1), we use the same basis set, MAB, for the monomers calculation as for the complex calculation. Besides, the total interaction energy can be decomposed into four components using the LMOEDA method,31 as shown in eqn (2)

 
ΔEint = ΔEelstat + ΔEex+rep + ΔEpol + ΔEdisp (2)
where ΔEelstat, ΔEex+rep, ΔEpol and ΔEdisp correspond to electrostatic, exchange-repulsion, polarization, and correlation terms, respectively. These components were computed at the SCF/aug-cc-pVTZ level except that the ΔEdisp term was obtained at the MP2 level, by using the GAMESS program.32

The bonding in the complexes was investigated based on two different but complementary methods, namely the quantum theory of atoms in molecules (QTAIM)33,34 and natural bond orbital (NBO) analysis.35 Using the AIM2000 and AIMALL program,34,36–38 we have obtained the molecular graphs for these complexes and the electron densities of π–magnesium bond critical points (BCPs).

Unless otherwise specified, all the calculations were performed with the Gaussian 09 program package.39

Results and discussion

Geometrical structures and π–magnesium bond

Using the counterpoise procedure, the optimized geometries of the X2Mg⋯C2H2, X2Mg⋯C2H4 and X2Mg⋯C6H6 (X = H, F) complexes with all real frequencies have been obtained at the MP2/aug-cc-pVTZ level, which are presented in Fig. 1. The important geometrical parameters are listed in Table 1. For comparisons, the optimized structures of corresponding monomers are given in Fig. S1 in the (ESI). The NBO charges of these complexes at the MP2 level are listed in Table 2. Note that we also obtained the hydrogen bonding structures of the X2Mg⋯C2H2 and X2Mg⋯C2H4 dimers (see Fig. S2, ESI), which are 4.45–9.50 kcal mol−1 higher in energy than π–magnesium bonding structures, so the latter are considered to be global minima on the interaction potential energy surface.
image file: c6ra23368f-f1.tif
Fig. 1 Optimized structures of the (I) H2Mg⋯C2H2, (II) H2Mg⋯C2H4, (III) H2Mg⋯C6H6, (IV) F2Mg⋯C2H2, (V) F2Mg⋯C2H4, and (VI) F2Mg⋯C6H6 complexes at the MP2/aug-cc-pVTZ level (P and O denote the centers of C1–C2 bond and the C6H6 ring, respectively).
Table 1 Main geometrical parameters of the X2Mg⋯C2H2, X2Mg⋯C2H4 and X2Mg⋯C6H6 (X = H, F) complexes. Bond lengths in Å and bond angles in degrees. P and O denote the centers of the C1–C2 bond and the C6H6 ring, respectively
Complex Symmetry dMg–P dMg–C1 RC1–C2 RMg–X ∠X1MgX2 ∠X1MgP ∠MgPO ∠X–MgP–C θa
a See Fig. 1 for definition.
H2Mg⋯C2H2 C2v 2.617 2.687 1.216 1.721 162.2 98.9   0.0 0.0
F2Mg⋯C2H2 C2v 2.460 2.534 1.216 1.785 159.5 100.3   0.0 0.9
H2Mg⋯C2H4 C2v 2.702 2.784 1.340 1.719 161.3 99.3   0.0 −0.4
F2Mg⋯C2H4 C2 2.523 2.611 1.341 1.783 157.1 101.5   61.8 −1.3
H2Mg⋯C6H6 Cs 2.683 2.773 1.401 1.725 159.4 105.9 90.6 90.0 −0.5
F2Mg⋯C6H6 Cs 2.504 2.601 1.404 1.789 151.5 115.9 86.0 90.0 0.5
H2Mg D∞h       1.707 180.0        
F2Mg D∞h       1.768 180.0        
C2H2 D∞h     1.212           0.0
C2H4 D2h     1.333           0.0
C6H6 D6h     1.394           0.0


Table 2 The NBO charges (in |e|) of selected atoms in the six X2Mg⋯C2H2, X2Mg⋯C2H4 and X2Mg⋯C6H6 (X = H, F) complexes. Δq = qcomplexqmonomer
  Mg X1/X2 C1–C2 QCTa
qcomplex qmonomer Δq qcomplex qcomplex qmonomer Δq
a QCT denotes the electron transfer from the conjugated molecule to the MgX2 molecule.
H2Mg⋯C2H2 1.199 1.253 −0.054 −0.621 −0.478 −0.450 −0.028 0.043
H2Mg⋯C2H4 1.184 1.253 −0.069 −0.619 −0.716 −0.682 −0.034 0.054
H2Mg⋯C6H6 1.188 1.253 −0.065 −0.620/−0.629 −0.488 −0.382 −0.106 0.061
F2Mg⋯C2H2 1.771 1.829 −0.058 −0.902 −0.510 −0.450 −0.060 0.033
F2Mg⋯C2H4 1.754 1.829 −0.075 −0.904 −0.756 −0.682 −0.074 0.054
F2Mg⋯C6H6 1.757 1.829 −0.072 −0.902/−0.903 −0.536 −0.382 −0.154 0.048


From Fig. 1 and Table 1, these dimers show intriguing structural features. The X2Mg and the acetylene molecules are coplanar in the X2Mg⋯C2H2 (X = H, F) complexes, which have C2v symmetry. The binding distance (dMg–P) of two monomers, defined as the distance between Mg atom and the center of C1–C2 bond, are 2.617 and 2.460 Å for the H2Mg⋯C2H2 and F2Mg⋯C2H2 complexes, respectively. As to the X2Mg⋯C2H4 species, H2Mg⋯C2H4 shows a similar structure to that of X2Mg⋯C2H2, that is, the H2Mg molecule is coplanar with the C[double bond, length as m-dash]C bond. Whereas, the situation is different for the F2Mg⋯C2H4 complex. A deflection of the F2Mg molecule is found when it interacts with the ethylene molecule. From Table 1, the dihedral angle ∠F–MgP–C is 61.8°. As a result, the F2Mg⋯C2H4 complex is of C2 symmetry. The repulsive interaction between negatively charged F and C atoms is considered to be the reason for the unusual deflection of the F2Mg molecule.

From Table 2, the C atom carries much more negative charge in ethylene than in acetylene and benzene molecules. Consequently, the F2Mg molecule deviates from the frontal plane to reduce the F⋯C repulsion interaction. The dMg–P distances of the X2Mg⋯C2H2 dimers are shorter than those of corresponding C2H4-based species, implying that acetylene is a stronger π electron donor than ethylene and has a stronger interaction with the X2Mg molecules. According to previous report, the LiF molecule favours the face center geometry while HF prefers the bond center geometry with benzene.24 Similar to the case of HF, the X2Mg molecule interacts with one of the C–C bond (partial double) instead of interacting with the center of the benzene ring (see Fig. 1). Besides, the C1–C2 bond is perpendicular to the X2Mg molecular plane and leads to Cs-symmetric structure of X2Mg⋯C6H6. From Table 1, the dMg–P distances are 2.683 and 2.504 Å for the H2Mg⋯C6H6 and F2Mg⋯C6H6 dimers, respectively, which are slightly shorter than those of corresponding X2Mg⋯C2H4 ones.

From Table 1, all the dMg–P distances vary in the range of 2.460–2.702 Å, which are longer than the conventional magnesium bond distances of 2.07–2.10 Å (ref. 21) in the LiNH2⋯HMgX (X = H, F, Cl, Br, CH3, OH, and NH2) dimers. Nevertheless, the Mg⋯C1 distances of 2.534–2.784 Å are much shorter than the sum of van der Waals radii of Mg and C atoms (3.7 Å),40 reflecting a strong interaction between X2Mg and the conjugated molecules, in the present six complexes. As shown in Table 2, the C1–C2 charges in the free C2H2, C2H4, and C6H6 molecules are negative (−0.382 to −0.682 |e|), and the Mg atom carries positive charges of 1.253 |e| and 1.829 |e|, in the MgH2 and MgF2 monomers, respectively. Thus, the nature of the interaction between X2Mg and conjugated molecules, namely π–magnesium bond, may mainly derive from the electrostatic attraction interaction. During the π–magnesium bond formation, the Mg atom becomes less positive, whereas the negative charge on the C1–C2 bond increases. Note that similar situation occurred for the interaction between FH and C6H6 molecules.41 From the table, 0.033–0.061 |e| electron transfer occurs from the conjugated molecules to the MgX2 molecule. It is noted that the charge transfer is ca. 0.02 |e| in the BeX2 (X = H, F) complexes of benzene,19 indicating a smaller charge transfer interaction in the π–beryllium bonding complexes.

Table 3 lists the main orbitals involved in charge transfer and corresponding second order stabilization energies of these π–magnesium bonding complexes, which are obtained on the basis of SCF electron densities. From the table, the main charge-transfer from the conjugated molecules to MgF2 is from the πCC bonding orbital to the empty 3s and 3p orbitals of Mg. As for the H2Mg-based series, the situation is a bit complex. From Table 3, there are two main orbital interactions for the H2Mg⋯C6H6 complex, namely πCC → p(Mg) and πCC → s(Mg), just like the case of F2Mg⋯π dimers. By contrast, the largest second order stabilization energy only involved electron donation from the πCC bonding orbital to the empty 3p orbital of Mg for the H2Mg⋯C2H4 and H2Mg⋯C2H2 complexes. No πCC → s(Mg) charge transfer has been found because the 3s orbital is mixed with a 3p orbital of Mg to form two sp hybrids involved in the H–Mg–H bond for these two complexes. Instead, there are some secondary donor–acceptor interactions, including the electron donation from πCC bonding orbital to the image file: c6ra23368f-t1.tif antibonding orbital, from σCC and σCH bonding orbitals to the empty 3p orbitals of Mg.

Table 3 Second order stabilization energies (kcal mol−1) of the orbital interactions in the X2Mg⋯C2H2 X2Mg⋯C2H4 and X2Mg⋯C6H6 (X = H, F) complexes
complex πCC → p(Mg) πCC → s(Mg)

image file: c6ra23368f-t2.tif

σCC → p(Mg) σCC → s(Mg) σCH → p(Mg)
F2Mg⋯C6H6 10.88 8.62   2.39    
F2Mg⋯C2H4 5.58 20.63        
F2Mg⋯C2H2 9.10 10.98        
H2Mg⋯C6H6 12.52 10.94   5.22 1.33 3.48
H2Mg⋯C2H4 18.39   5.76 1.55   7.44
H2Mg⋯C2H2 19.81   4.74 2.24   6.50


Table 1 shows that the binding distance dMg–P increases in the order 2.460 Å (F2Mg⋯C2H2) < 2.504 Å (F2Mg⋯C6H6) < 2.523 Å (F2Mg⋯C2H4) and 2.617 Å (H2Mg⋯C2H2) < 2.683 Å (H2Mg⋯C6H6) < 2.702 Å (H2Mg⋯C2H4). The data suggest that C2H2 is the best π electron donor among the three conjugated molecules and may form the strongest π–magnesium bond with the X2Mg molecules. On the other hand, the dMg–P distances of the F2Mg-based complexes are obviously shorter than those of H2Mg-based ones, reflecting that the π–magnesium bonds are stronger in the former species. This is reasonable because MgF2 is a better Lewis acid than MgH2.

Due to the π–magnesium bond formation, the Mg–X bonds of the MgX2 molecules elongate 0.012–0.021 Å. For instance, the Mg–H bond lengths are 1.721, 1.719 and 1.725 Å in the H2Mg⋯C2H2, H2Mg⋯C2H4 and H2Mg⋯C6H6 complexes, respectively, while it is 1.707 Å in the free MgH2 molecule. Meanwhile, both Mg–X bonds are seen to bend in a direction opposite to the conjugated molecules, leading to a bending angle ∠X1MgX2 of 151.5–162.2°.

For the unsaturated moiety, the structural distortion is not obvious when compared with the corresponding monomer. From Table 1 and Fig. S1 (ESI), the C1–C2 bond slightly lengthens upon the complex formation. For instance, the RC1–C2 length of 1.261 Å for X2Mg⋯C2H2 is 0.05 Å longer than that of the isolated C2H2 molecule. In contrast, the length of RC1–C2 only elongates 0.01 Å due to the interaction between X2Mg and the benzene ring. Such bond lengthenings indicate slightly weaker C1–C2 bonds in the complexes, which is confirmed by their lesser Wiberg bond index than those in corresponding monomers (see Table S1, ESI). Besides, as can be seen from Table 1, the angle θ (the deflection of C–H bond from the horizontal plane that contains C–C bond) is 0.9° in the F2Mg⋯C2H2 complex, showing that the C2H2 subunit is not linear herein. Similarly, the C2H4 and C6H6 subunits are not strictly planar in the studied complexes with θ ranging from −1.3° to 0.5°.

Interaction energies

Interaction energy is one of the most important measurement of the intermolecular interaction. Table 4 collects the BSSE-corrected interaction energies of the X2Mg⋯C2H2, X2Mg⋯C2H4 and X2Mg⋯C6H6 (X = H, F) complexes. The CCSD(T) interaction energies of these complexes range from −7.22 to −15.80 kcal mol−1. The BSSE of 0.31–1.61 kcal mol−1 is less than 10.2% of the total interaction energy. The MP2 results agree very well with those from the CCSD(T) method, and the difference between the both does not exceed 7.1%. So the MP2 method is also a good choice to characterize the strength of π–magnesium bond. From the table, electron correlation is important for interaction energy calculations of the H2Mg-based species, which amounts to 18.7–47.6% of the CCSD(T) interaction energy. In contrast, it contributes much less to the interaction energy of the F2Mg-based complexes.
Table 4 Interaction energies (kcal mol−1) of the six complexes at different levels with the aug-cc-pVTZ basis set. Electron correlation contribution has been computed by the formula EC = |CCSD(T) − SCF|/CCSD(T) × 100%
Complex SCF MP2 CCSD(T) BSSE EC dMg–P
F2Mg⋯C6H6 −13.74 −15.52 −15.80 1.61 13.0% 2.504
F2Mg⋯C2H4 −13.08 −13.16 −13.47 0.90 2.9% 2.523
F2Mg⋯C2H2 −15.71 −15.00 −15.66 0.81 0.3% 2.460
H2Mg⋯C6H6 −4.45 −9.09 −8.49 0.70 47.6% 2.683
H2Mg⋯C2H4 −5.11 −7.48 −7.22 0.34 29.2% 2.702
H2Mg⋯C2H2 −7.24 −8.97 −8.91 0.31 18.7% 2.617


It has been reported that the interaction energies of π–lithium12 and π–beryllium19 bonding complexes usually increase in a linear manner with the decreasing binding distances between two monomers. So the binding distances (dMg–P) of the π–magnesium bonding dimers are also shown in Table 4 to facilitate comparison. From the table, the absolute interaction energy increases in the order H2Mg⋯C2H4 (−7.22 kcal mol−1) < H2Mg⋯C6H6 (−8.49 kcal mol−1) < H2Mg⋯C2H2 (−8.91 kcal mol−1), which accords with the decreasing Mg⋯P distance from H2Mg⋯C2H4 to H2Mg⋯C2H2.

However, there is an exception in the MgF2-based complexes. As shown in Table 4, the interaction energy increases in the order F2Mg⋯C2H4 (−13.47 kcal mol−1) < F2Mg⋯C2H2 (−15.66 kcal mol−1) < F2Mg⋯C6H6 (−15.80 kcal mol−1), that is, the F2Mg⋯C6H6 complex exhibits larger interaction energy than that of F2Mg⋯C2H2 though the latter has a shorter binding distance. As discussed later, this inconsistency can be interpreted on the basis of AIM analysis. Besides, the interaction energies of F2Mg-based complexes are nearly twice of those of the H2Mg-based ones, indicating that the former species are much more stable than the latter. This is consistent with the shorter binding distances in the F2Mg-based species.

In order to gain more insight into the origin and nature of π–magnesium bonds, the LMOEDA analysis has been performed. The results indicate significant electrostatic and polarization energies in the complexation between MgX2 and π electron donors. The former is at least half and the latter accounts for ca. 28–45%, of the total attractive interaction energy (see Table 5). Similarly, the Yañez's group19 also reported that the electrostatic term constituted the largest stabilizing contribution to the formation of π–beryllium bonding complexes. It is worth mentioning that the dispersion energy also makes a non-negligible contribution (17.6%) to the stabilization of the H2Mg⋯C6H6 complex. Obviously, the MgF2-based complexes exhibit larger attractive interaction energies but smaller exchange-repulsion energies than the MgH2-based ones.

Table 5 LMO–EDA partition terms (kcal mol−1) for the X2Mg⋯C2H2, X2Mg⋯C2H4 and X2Mg⋯C6H6 (X = H, F) complexes
  Eelst Epol Edisp Eex+rep
F2Mg⋯C6H6 −15.35 −13.83 −1.82 15.45
F2Mg⋯C2H4 −18.53 −9.91 −0.11 15.38
F2Mg⋯C2H2 −23.14 −10.47 0.68 17.90
H2Mg⋯C6H6 −13.06 −8.68 −4.64 17.26
H2Mg⋯C2H4 −15.82 −7.50 −2.37 18.18
H2Mg⋯C2H2 −19.16 −8.11 −1.74 20.00


As reported in previous works, the interaction energies of the FLi⋯Y and HNa⋯Y (Y = C2H2, C2H4 and C6H6) complexes vary in the ranges of −8.35 to −10.83 kcal mol−1 (calculated at the B3LYP/6-311++G(d, p) level)12 and −4.32 to −6.90 kcal mol−1 (at the MP2/6-311++G(d, p) level),13 respectively. Obviously, the interaction energies of the topic complexes are much larger than those of corresponding π–lithium and π–sodium bonding species. The F2Mg⋯C2H2 and F2Mg⋯C2H4 complexes possess comparable interaction energies to those of π–beryllium bonding X2Be⋯C2H2/C2H4 (X = H, F, Cl) species, whose interaction energies vary in the range of −11.90 to −14.76 kcal mol−1 at the CCSD(T)/aug-cc-pVTZ level.19 In contrast, the H2Mg⋯C2H2 and H2Mg⋯C2H4 complexes exhibit much smaller interaction energies. In addition, the interaction energies are −10.00 and −7.10 kcal mol−1 for the F2Be⋯C6H6 and H2Be⋯C6H6 complexes at the PBE/TZ2P level,42 as reported by Feng's group. Hence, the X2Mg⋯C6H6 complex has larger interaction energy than X2Be⋯C6H6, whether X represents F or H atom. As far as conventional magnesium bonding complexes are concerned, Yang and co-workers reported considerable binding energies of 63.2–66.5 kcal mol−1 for the LiNH2⋯HMgX (X = H, F, Cl, Br, CH3, OH, and NH2) complexes.21 This is not surprising in view of the fact that these complexes are stabilized not only by magnesium bonds but also by concomitant lithium bonds.

Harmonic vibrational frequencies and Raman activity

The calculated harmonic vibrational frequencies for typical intermolecular vibrational modes of the X2Mg⋯C2H2, X2Mg⋯C2H4 and X2Mg⋯C6H6 (X = H, F) complexes are shown in Table 6, the corresponding data of monomers are collected in Table S2 in ESI. In view of the fact that the C–C and X–Mg–X symmetric stretching vibrations do not have infrared intensity, the Raman intensities of the representative vibrations are shown in the tables. From Table 6, the frequencies of the stretching mode of the π–magnesium bonds π⋯Mg are in the range of 141.2–240.7 cm−1. The symmetric stretching frequency of the MgH2 molecule decreases as much as 42.8 cm−1, 38.7 cm−1, and 51.7 cm−1 when interacting with acetylene, ethylene and benzene molecules, respectively, accompanied by a decrease in Raman intensity. In comparison, the MgF2 symmetric stretching vibration is much less red-shifted in the complexes. Meanwhile, all the X–Mg–X antisymmetric stretching vibrations undergo substantial redshift of 44.1–61.1 cm−1 along with the π–magnesium bond formation. The C–C stretching vibrations of C2H2 and C2H4 molecules are also red-shifted when complexed with the MgX2 molecule, and the decreased vibrational frequencies are 14.7–20.0 cm−1. Effect of the complexation with MgX2 molecule on the C1–C2 stretching vibration of C6H6 is relatively small. From Table 6, the C–C stretching vibration frequency of C6H6 only decreases 7.6 and 7.0 cm−1 in the H2Mg⋯C6H6 and F2Mg⋯C6H6 dimers, respectively. Note that the redshifts of C–C and Mg–X stretching vibration modes accord well with the bond lengthening, both indicating the weakened C–C and Mg–X bonds upon π–magnesium bond formation.
Table 6 Main harmonic vibrational frequencies (v, cm−1) and corresponding Raman intensities (RI, au) of the X2Mg⋯C2H2, X2Mg⋯C2H4, and X2Mg⋯C6H6 (X = H, F) complexes
  X–Mg–X sym. stretch X–Mg–X antisym. stretch C–C stretch π··· Mg stretch
v Δv ΔRI v Δv ΔRI v Δv ΔRI v
H2Mg⋯C2H2 1591.3 −42.8 −90.6 1606.3 −50.6 16.6 1952.0 −15.7 16.6 168.2
F2Mg⋯C2H2 540.8 −10.7 −1.4 815.2 −46.0 0.2 1953.1 −14.7 0.2 240.7
H2Mg⋯C2H4 1595.5 −38.7 −70.7 1610.7 −46.2 24.4 1661.1 −16.9 24.4 145.0
F2Mg⋯C2H4 546.2 −5.2 −1.4 817.0 −44.1 0.2 1658.0 −20.0 0.2 222.7
H2Mg⋯C6H6 1582.5 −51.7 −151.9 1603.3 −53.6 55.6 1007.4 −7.6 55.6 141.2
F2Mg⋯C6H6 545.3 −6.1 −2.0 800.0 −61.1 0.3 1005.2 −7.0 0.3 208.0


AIM analysis

Bader's quantum theory of atoms in molecules (QTAIM)33,34 is useful in providing significant insight into the nature and strength of intermolecular interactions. This theory is based on the topological analysis of electron density and its Laplacian at the bond critical point (BCP). The Laplacian is the sum of λ1, λ2, and λ3, the three eigenvalues of the Hessian matrix of electron density.

In order to gain more information about the weak interactions in π–magnesium bonding dimers, the calculated topological parameters at the π⋯Mg BCPs for the six complexes are exhibited in Table 7, where the electron densities ρ(r)BCP and their Laplacians ∇2ρ(r) are also given. From Table 7, the ρ(r) values increase in the order 0.0116 au for H2Mg⋯C6H6 < 0.0125 au for H2Mg⋯C2H4 < 0.0138 au for H2Mg⋯C2H2 and 0.0160 au for F2Mg⋯C6H6 < 0.0175 au for F2Mg⋯C2H4 < 0.0190 au for F2Mg⋯C2H2, demonstrating that acetylene is the best Lewis base and can form the strongest π–magnesium bond herein. By contrast, the C6H6 molecule is a weaker π electron donor in view of the relatively smaller ρ(r) values of the π⋯Mg BCPs.

Table 7 The topological properties at the π⋯Mg bond critical point (all units are atomic units)
Complex ρ(r) λ1 λ2 λ3 2ρ(r)
H2Mg⋯C2H2 0.0138 −0.0132 −0.0073 0.0828 0.0624
F2Mg⋯C2H2 0.0190 −0.0208 −0.0155 0.1333 0.0971
H2Mg⋯C2H4 0.0125 −0.0117 −0.0055 0.0644 0.0471
F2Mg⋯C2H4 0.0175 −0.0189 −0.0133 0.1096 0.0774
H2Mg⋯C6H6 0.0116 −0.0077 −0.0051 0.0568 0.0440
F2Mg⋯C6H6 0.0160 −0.0162 −0.0076 0.0989 0.0750


The molecular graphs of the X2Mg⋯C2H2, X2Mg⋯C2H4 and X2Mg⋯C6H6 complexes are presented in Fig. 2. The figure shows that there only exists one bond path (BP) linking the magnesium atom with C[double bond, length as m-dash]C or C[triple bond, length as m-dash]C bond by a bond critical point (BCP) for the H2Mg⋯C2H2, H2Mg⋯C2H4, F2Mg⋯C2H2 and F2Mg⋯C2H4 complexes. For the X2Mg⋯C2H2 complexes, the bond path actually connects the Mg atom and the non-nuclear attractor (NNA) situated between two BCPs of the CC bond (Fig. 2). The similar case occurred in the interaction between AlF3 and C2H2 molecules.14 Whereas, the situation is more complex for the F2Mg⋯C6H6 complex, where exists four kinds of interactions between the F2Mg and C6H6 molecules. The first one is the magnesium bond formed between Mg and the center of C1–C2 bond via a BCP (ρ(r) = 0.0160 au). The second interaction is cage path (CP) linking Mg, F and π electron of benzene containing a cage critical point (CCP) with ρ(r)CCP of 0.0052 au. The third one is two ring paths (RP) linking Mg, F and some C atoms by ring critical point (RCP) with ρ(r)RCP of 0.0059 au. Besides, we can also see another bond path linking F1 atom and one of the C–C bond, indicating an interaction between F atom and the benzene ring and the corresponding ρ(r)BCP value is 0.0063 au. These additional interactions between MgF2 and C6H6 molecules may well explain why the F2Mg⋯C6H6 complex exhibits a larger interaction energy than that of F2Mg⋯C2H2 although it has a longer binding distance and a smaller Mg⋯π ρ(r)BCP value. Similarly, there are one RCP and one CCP between the H2Mg molecule and benzene ring in addition to the Mg⋯π BCP (ρ(r) = 0.0116 au), and the corresponding ρ(r)RCP and ρ(r)CCP values are both equal to 0.0050 au. As a result, the H2Mg⋯C6H6 complex is more stable than H2Mg⋯C2H4 even though the π–magnesium bonding interaction is stronger in the latter. As far as the π–beryllium bonding species is concerned, there is only one bond path was found linking the Be atom to one carbon atom of benzene, and the ρ(r)BCP values are 0.0190 and 0.0223 au for the H2Be⋯C6H6 and F2Be⋯C6H6 complexes, respectively.42 By contrast, the Mg⋯π ρ(r)BCP values are slightly smaller (see Table 7). Nevertheless, the X2Mg⋯C6H6 complexes show larger interaction energies than corresponding X2Be⋯C6H6 ones due to the combined effect of multiple interactions between X2Mg and benzene molecules.


image file: c6ra23368f-f2.tif
Fig. 2 The molecular graphs of the X2Mg⋯C2H2, X2Mg⋯C2H4 and X2Mg⋯C6H6 (X = H, F) complexes (red, green and blue dots denote BCP, CCP and RCPs, respectively).

Conclusion

The structural characteristics and interaction nature of complexes formed between MgX2 (X = H, F) and conjugated molecules have been investigated in detail. The stability of the topic species originates from the formation of π–magnesium bonds, where 0.033–0.061 |e| electron transfer occurs and the conjugated molecules play the role of electron donor. NBO analysis reveals that the main charge-transfer involves the πCC bonding orbital and the empty lone pair orbitals of Mg. Interesting geometrical features have been found for these complexes. The angles between the MgF2 molecular plane and C1–C2 bond are 0°, 61.8° and 90° with Lewis bases being C2H2, C2H4 and C6H6, respectively. During the π–magnesium bond formation, all the Mg–X bonds lengthened. Accordingly, the X–Mg–X stretching vibrations are red-shifted. The same holds for the C–C stretching modes. According to the computed interaction energies, the MgF2-based complexes are more stable than the MgH2-based ones, which may be attributed to the less positive Mg charge of the MgH2 molecule. AIM analysis indicates that MgX2 forms stronger π–magnesium bond with acetylene compared to ethylene and benzene. Meanwhile, several additional interactions have been found between the MgX2 and C6H6 molecules, explaining why the F2Mg⋯C6H6 complex shows the largest interaction energy value.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant No. 21173095, 21303066, 21573089) and State Key Development Program for Basic Research of China (Grant No. 2013CB834801).

References

  1. R. W. Saalfrank, H. Maid and A. Scheurer, Angew. Chem., Int. Ed., 2008, 47, 8794 CrossRef CAS PubMed.
  2. A. Llanes-Pallas, C.-A. Palma, L. Piot, A. Belbakra, A. Listorti, M. Prato, P. Samorì, N. Armaroli and D. Bonifazi, J. Am. Chem. Soc., 2008, 131, 509 CrossRef PubMed.
  3. A. Bauzá and A. Frontera, Angew. Chem., Int. Ed., 2015, 54, 7340 CrossRef PubMed.
  4. P. Politzer, J. S. Murray and T. Clark, J. Mol. Model., 2015, 21, 1 CrossRef CAS PubMed.
  5. G. Jones, P. Willett and R. C. Glen, J. Mol. Biol., 1995, 245, 43 CrossRef CAS PubMed.
  6. M. G. Chudzinski, C. A. McClary and M. S. Taylor, J. Am. Chem. Soc., 2011, 133, 10559 CrossRef CAS PubMed.
  7. J. Itskovitz-Eldor, M. Schuldiner, D. Karsenti, A. Eden, O. Yanuka, M. Amit, H. Soreq and N. Benvenisty, Mol. Med., 2000, 6, 88 CAS.
  8. S. K. Silverman and T. R. Cech, Biochemistry, 1999, 38, 8691 CrossRef CAS PubMed.
  9. K. Morokuma, J. Chem. Phys., 1971, 55, 1236 CrossRef CAS.
  10. P. A. Kollman, J. F. Liebman and L. C. Allen, J. Am. Chem. Soc., 1970, 92, 1142 CrossRef CAS.
  11. B. S. Ault and G. C. Pimentel, J. Phys. Chem., 1975, 79, 621 CrossRef CAS.
  12. K. Yuan, Y. Liu, L. LÜ, Y. Zhu, J. Zhang and J. Zhang, Chin. J. Chem., 2009, 27, 697 CrossRef CAS.
  13. Z. Li, X. Zhang, H. Li, Y. Zhu and X. Yang, Chem. Phys. Lett., 2011, 510, 273 CrossRef CAS.
  14. S. J. Grabowski, Molecules, 2015, 20(6), 11297–11316 CrossRef CAS PubMed.
  15. M. Yáñez, P. Sanz, O. Mó, I. Alkorta and J. Elguero, J. Chem. Theory Comput., 2009, 5, 2763 CrossRef PubMed.
  16. K. Eskandari, J. Mol. Model., 2012, 18, 3481 CrossRef CAS PubMed.
  17. M. M. Montero-Campillo, A. M. Lamsabhi, O. Mo and M. Yañez, J. Mol. Model., 2013, 19, 2759 CrossRef CAS PubMed.
  18. M. Marin-Luna, I. Alkorta, J. Elguero, O. Mo and M. Yañez, Molecules, 2015, 20, 9961–9976 CrossRef CAS PubMed.
  19. E. F. Villanueva, O. Mo and M. Yañez, Phys. Chem. Chem. Phys., 2014, 16, 17531 RSC.
  20. D. Yu, Y. Li, D. Wu and S. Y. Li, Theor. Chem. Acc., 2016, 135, 112 CrossRef.
  21. X. Yang, Q. Li, J. Cheng and W. Li, J. Mol. Model., 2013, 19, 247 CrossRef CAS PubMed.
  22. E. P. Gates, J. K. Jensen and J. N. Harb, et al., RSC Adv., 2015, 5(11), 8134–8141 RSC.
  23. C. Y. Li, E. A. Hemmig and J. Kong, et al., ACS Nano, 2015, 9(2), 1420–1433 CrossRef CAS PubMed.
  24. S. S. C. Ammal and P. Venuvanalingam, J. Chem. Phys., 1998, 109(22), 9820–9830 CrossRef CAS.
  25. R.-Y. Li, Z.-R. Li, D. Wu, Y. Li, W. Chen and C.-C. Sun, J. Phys. Chem. A, 2005, 109, 2608 CrossRef CAS PubMed.
  26. A. V. Krasheninnikov, P. O. Lehtinen and A. S. Foster, et al., Phys. Rev. Lett., 2009, 102(12), 126807 CrossRef CAS PubMed.
  27. C. Xu, X. Wang and J. Zhu, J. Phys. Chem. C, 2008, 112(50), 19841–19845 CAS.
  28. F. Li, P. Li and m. Yuan, et al., RSC Adv., 2014, 4, 55781 RSC.
  29. S. F. Boys and F. Bernardi, Mol. Phys., 1970, 19, 553 CrossRef CAS.
  30. I. Alkorta and J. Elguero, J. Phys. Chem. A, 1999, 103, 272 CrossRef CAS.
  31. P. Su and H. Li, J. Chem. Phys., 2009, 131, 014102 CrossRef PubMed.
  32. M. W. Schmidt, K. K. Baldridge, J. A. Boatz, S. T. Elbert, M. S. Gordon, J. H. Jensen, S. Koseki, N. Matsunaga, K. A. Nguyen, S. Su, T. L. Windus, M. Dupuis and J. A. Montgomery, J. Comput. Chem., 1993, 14, 1347 CrossRef CAS.
  33. F. Cortés-Guzmán and R. F. Bader, Coord. Chem. Rev., 2005, 249, 633 CrossRef.
  34. R. F. Bader and C. F. Matta, J. Phys. Chem. A, 2004, 108, 8385 CrossRef CAS.
  35. A. E. Reed, L. A. Curtiss and F. Weinhold, Chem. Rev., 1988, 88, 899 CrossRef CAS.
  36. R. F. W. Bader, J. Phys. Chem. A, 1998, 102, 7314 CrossRef CAS.
  37. T. Keith, TK Gristmill Software, Overland Park KS, USA, 2011 Search PubMed.
  38. F. Biegler-König and J. Schönbohm, J. Comput. Chem., 2002, 23, 1489 CrossRef PubMed.
  39. G. W. T. M. J. Frisch, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09 Revision A.02, Gaussian, Inc., Wallingford, CT, 2009 Search PubMed.
  40. S. Batsanov, Inorg. Mater., 2001, 37, 871 CrossRef CAS.
  41. S. J. Grabowski and J. M. Ugalde, J. Phys. Chem. A, 2010, 114(26), 7223–7229 CrossRef CAS PubMed.
  42. Q. Zhao, D. Feng, Y. Sun and J. Hao, Comput. Theor. Chem., 2011, 964, 188 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra23368f

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.