Orde Q.
Munro
*,
Sandra D.
Strydom
and
Craig D.
Grimmer
School of Chemical and Physical Sciences, University of Natal, Pietermaritzburg, Private Bag X01, Scottsville, 3209, South Africa. E-mail: munroo@nu.ac.za; Fax: +27 33 260 5009; Tel: +27 33 260 5270
First published on 27th October 2003
The X-ray crystal structure of a novel mono(pyrrole)-containing tridentate Schiff base ligand, N-[(1E)-1H-pyrrol-2-ylmethylene]benzene-1,2-diamine (3), shows unexpected hydrogen bond complementarity between neighbouring pairs of molecules in the solid state (monoclinic space group C2). The supramolecular structure is best described as a hinged dimer in which the dihedral angle between the two ligand planes measures 65(1)° and is stabilized by a pair of hydrogen bonds involving the pyrrole N–H proton on one molecule and the N atom of the aryl amino group on the neighbouring molecule. The experimental H(pyrrole)⋯N(amine) H-bond distances are equivalent at 2.40(2)
Å. The 1H NMR spectrum of 3 in CDCl3 shows concentration-dependent features that are consistent with endergonic dimerization by intermolecular hydrogen bonding at higher concentrations [KD=
0.89(16) M−1 at 25
°C]. DFT calculations at the B3LYP/6-31G** level of theory correctly predict the main geometrical features of the dimer, which has slightly inequivalent H(pyrrole)⋯N(amine) distances of 2.31 and 2.33 Å in the calculated structure. The DFT-calculated dimer shows some conformational differences from the X-ray structure that primarily reflect the fact that the individual molecules of 3 in the in vacuo model are twisted about the C
N–C
C torsion angle involving the imine group and adjacent phenyl ring (C5–N2–C6–C7) by 4–8° more than is the case for the experimental structure. These conformational differences reflect the role that crystal packing plays in fine-tuning the supramolecular structure of 3.
![]() | ||
Scheme 1 Schiff bases containing pyrrole groups. |
Ligand groups such as 1 are attractive as synthons since metallation of the chelate with concomitant deprotonation of the pyrrole ring affords an anionic ligand, a useful strategy for stabilizing metal ions in high oxidation states. Indeed, a search of the Cambridge Structural Database (CSD) reveals that there are well over 100 crystallographically characterized complexes of transition and main group metal ions that have been complexed with macrocycles or polydentate chelates based on fragment 1. Two simple, yet exemplary, derivatives of this system are the 3d transition metal complexes bis(N-t-butylpyrrole-2-carbaldimino)copper(II)1 and cis-dichlorobis[2-(ethyliminomethyl)pyrrolide]titanium(IV).2
One of our current objectives is to use the tetradentate Schiff base ligand 2 {N,N′-bis[(1E)-1H-pyrrol-2-ylmethylene]benzene-1,2-diamine} as a porphyrin-like chelate for the complexation of metal ions to complement the work we have being doing with porphyrins.3,4 Our interest in compound 2 is based on two observations. First, porphyrins may formally be considered to comprise two trans pyrrole nitrogens and two trans imine-type nitrogens—the donor atom combination of 2 is thus an analogue of this pattern that has the imine nitrogen pair cis or adjacent in the ligand. Second, the electronic structure of the ligand is not unlike that of a porphyrin since the pyrrole and imine nitrogens are also capable of π-bonding with the metal dπ orbitals. There are some obvious differences between porphyrins and 2, including the 5-membered chelate rings of 2 (porphyrins contain 6-membered chelate rings) and the fact that 2 is not a macrocycle. These fundamental differences are in fact of some significance because they may provide tractable clues, such as differences in metal ion affinity constants, as to the functional importance of the four equivalent 6-membered chelate rings and the macrocyclic structure that typify porphyrins.
Tetradentate Schiff base derivatives such as 2 are normally prepared by reacting 2 equiv. of pyrrole-2-carboxaldehyde and the appropriate diamine in ethanol or dichloromethane.5 It occurred to us during this work that the tridentate mono(pyrrole) Schiff base ligand 3 {N-[(1E)-1H-pyrrol-2-ylmethylene]benzene-1,2-diamine} would also be easy to prepare by reaction of the diamine with 1 equiv. of the appropriate aldehyde. Compound 3 is interesting because it is an asymmetric tridentate ligand that may be further functionalized by reaction with a range of aldehydes of varying complexity. In effect, 3 is a potentially important synthon for the synthesis of asymmetric or chiral tridentate/tetradentate ligands with applications in enantioselective catalysis, currently a highly active field of inquiry.6–10
Schiff base 3 crystallizes from dichloromethane–hexane in the monoclinic space group C2 and displays an interesting supramolecular structure that involves complementary hydrogen bonding between neighbouring pairs of molecules (Fig. 1). In this paper, we describe the synthesis and X-ray crystal structure of 3 as well as density functional theory (DFT) calculations aimed at elucidating the intermolecular interactions between molecules of 3 in the solid state. Moreover, we have been able to show that Schiff base 3 forms H-bonded dimers at high concentrations in CDCl3 by 1H NMR spectroscopy. Significantly, the spectroscopic evidence suggests that the solid state supramolecular structure is probably also favoured in solution.
![]() | ||
Fig. 1 Schematic illustration of the H-bond complementarity found in the X-ray crystal structure of 3. |
Anal. calcd for C11H11N3: C, 71.3; H, 5.99; N, 22.7; found: C, 70.9; H, 5.97; N, 22.5. UV-vis (CH2Cl2)
λmax/nm: 245, 301, 365. δH
(500 MHz, CDCl3): 4.26 (2H, s, br, a), 6.30 (1H, dd, 3J1 3.4, 3J2 2.5, b), 6.71 (1H, dd, 3J 3.4, 4J 1.6, c), 6.74 (1H, m, d), 6.79 (1H, dd, 3J 8.1, 4J 1.2, e), 6.85 (1H, td, 3J 7.8, 4J 1.2, f), 7.07 (1H, dd, 3J 7.8, 4J 0.93, g), 7.11 (1H, td, 3J 8.1, 4J 1.2, h), 8.33 (1H, s, i), 10.25 (1H, s, br, j). δC
(125 MHz, CDCl3): 110.16 (C-1), 115.38 (C-2), 116.54 (C-3), 117.58 (C-4), 118.62 (C-5), 123.34 (C-6), 126.67 (C-7), 130.77 (C-8), 137.55 (C-9), 141.21 (C-10), 148.62 (C-11). IR (KBr pellet, cm−1): 3454w νa(NH2), 3352w νs(NH2), 3296w br ν(N–H, pyrrole), 3103w ν(C–H), 3024w ν(C–H), 1616s ν(CN), 1599s δ(NH2), 1493m, 1416m, 1326m, 1298w, 1262m, 1121m, 1088m, 1031s, 883w, 824w, 741vs, 604w, 593w, 469w.
The structure was solved in the monoclinic space group C2 using direct methods in WinGX's12 implementation of SHELXS-97.13 All non-H atoms were located in the E-map and refined anisotropically with SHELXL-97.13 The data set was of sufficiently high quality that all H atoms were located in the final difference Fourier synthesis cycle. The coordinates of the H atoms attached to the pyrrole and amino nitrogens of 3 were refined isotropically without constraints; however, the isotropic Uij values were fixed at 1.2 times the Ueq values of the nitrogen atoms. All other H atoms were calculated using the standard riding model of SHELXL-97 (HFIX 44 instruction), which allows the X–H distances to vary during refinement. The absolute configuration of 3 could not be determined unambiguously from the value of the Flack parameter.14 Selected crystal data and refinement details are given in Table 1.‡
a
w![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
|
---|---|
Chemical formula | C11H11N3 |
M r | 185.23 |
Cell setting | Monoclinic |
Space group | C2 |
a/Å | 18.373(8) |
b/Å | 5.489(4) |
c/Å | 11.369(4) |
β/° | 124.22(3) |
U/Å3 | 948.1(9) |
Z | 4 |
μ/mm−1 | 0.08 |
T/K | 122 (2) |
Total reflections | 4084 |
Independent reflections | 2065 |
Observed reflections [I![]() ![]() |
1936 |
R int | 0.030 |
R
[F2![]() ![]() |
0.036 |
wR (F2)a | 0.102 |
The X-ray crystal structure of 3
(Fig. 2) confirms that the E isomer of the Schiff base is obtained from the condensation reaction. Inspection of the space-filling plot of the structure reveals that this isomer is probably preferred on steric grounds since the Z isomer would require rotation about the N2–C5 bond by 180° and this would lead to rather severe nonbonded repulsion between H7 and H1A. Interestingly, the E isomer of 3 is not completely planar. There is a 4.0(2)° twist about the C7–C6–N2–C5 torsion angle that apparently reflects a conformational adjustment to alleviate an unfavourable steric interaction between H7 and H5 (H7⋯H5=
2.05 Å). Moreover, as shown in the schematic diagram of Fig. 2(b), the out-of-plane displacements of the non-H atoms of 3 are consistent with the molecule adopting a somewhat bowed conformation in which the two aromatic rings partly dip below the 14-atom mean plane. (The imine group atoms N2 and C5 are displaced above the mean plane.)
![]() | ||
Fig. 2 (a) Labelled ORTEP diagram of 3. Displacement ellipsoids are drawn at the 60% probability level; H atoms are drawn as spheres with equal, though arbitrary radii. (b) Diagram illustrating the perpendicular displacements of selected atoms from the 14-atom (non-H) mean plane of 3. (c) Partly labelled space-filling view (van der Waals radii) of 3 approximately perpendicular to the phenyl ring plane. |
The refined coordinates of the amino group H atoms show that the N atom has mainly sp3-hybridized character. However, it is interesting to note that the values of the bond angles C11–N3–H3A and C11–N3–H3B (Table 2) are actually between those expected for an unsubstituted sp3 hybridized N atom (107°) and an sp2 hybridized N atom (120°). Thus, although H1A and H1B are located above the 14-atom mean plane of 3, the amino group is somewhat flatter than expected for a purely sp3-hybridized N atom. The X-ray data therefore suggest that the hybridization of N3 is best described as an sp3/sp2 admixture. This interpretation is in fact confirmed by the in vacuo DFT-calculated structure of 3 (vide infra).
a The s.u. of each value is given in parentheses. | |||
---|---|---|---|
C (1)–N(1) | 1.367(2) | C(6)–C(11) | 1.414(2) |
C(1)–C(2) | 1.383(2) | C(7)–C(8) | 1.392(2) |
C(2)–C(3) | 1.416(2) | C(8)–C(9) | 1.394(2) |
C(3)–C(4) | 1.388(2) | C(9)–C(10) | 1.394(2) |
C(4)–N(1) | 1.375(2) | C(10)–C(11) | 1.404(2) |
C(4)–C(5) | 1.439(2) | C(11)–N(3) | 1.386(2) |
C(5)–N(2) | 1.282(2) | N(1)–H(1A) | 0.88(2) |
C(6)–C(7) | 1.399(2) | N(3)–H(3A) | 0.88(2) |
C(6)–N(2) | 1.414(2) | N(3)–H(3B) | 0.88(2) |
N(1)–C(1)–C(2) | 108.3(1) | C(10)–C(9)–C(8) | 120.8(1) |
C(1)–C(2)–C(3) | 107.2(1) | C(9)–C(10)–C(11) | 120.0(1) |
C(4)–C(3)–C(2) | 107.2(1) | N(3)–C(11)–C(10) | 121.8(1) |
N(1)–C(4)–C(3) | 107.8(1) | N(3)–C(11)–C(6) | 118.7(1) |
N(1)–C(4)–C(5) | 121.9(1) | C(10)–C(11)–C(6) | 119.4(1) |
C(3)–C(4)–C(5) | 130.1(1) | C(1)–N(1)–C(4) | 109.3(1) |
N(2)–C(5)–C(4) | 120.8(1) | C(1)–N(1)–H(1A) | 125.5(1) |
C(7)–C(6)–N(2) | 126.2(1) | C(4)–N(1)–H(1A) | 125.2(1) |
C(7)–C(6)–C(11) | 119.6(1) | C(5)–N(2)–C(6) | 121.8(11) |
N(2)–C(6)–C(11) | 114.2(1) | C(11)–N(3)–H(3A) | 118(1) |
C(8)–C(7)–C(6) | 120.7(1) | C(11)–N(3)–H(3B) | 116(1) |
C(7)–C(8)–C(9) | 119.5(1) | H(3A)–N(3)–H(3B) | 114(2) |
The unit cell packing for 3 (Fig. S1, Electronic supplementary information) clearly indicates that the central pair of molecules within the unit cell constitutes a hydrogen-bonded dimer in which there is a complementary interaction between the amino group lone pair (N3) of one molecule and the pyrrole NH group H atom (H1A) of the neighbouring molecule. The angle between the non-H 14-atom mean planes of the two molecules in the H-bonded pair measures 65(1)°. This reflects the projection of the lone pair of electrons that belong to N3 at an angle that is close to orthogonal to the benzene ring. Such a location of the H-bond acceptor site, which has significant sp2 hybrid character as a result of hyperconjugation of the lone pair of electrons with the aryl ring π-electrons, essentially precludes any co-planar alignments for the hydrogen bond donor molecule.
The intriguing arrangement of H-bonding interactions in the crystal structure of 3 is illustrated graphically in Fig. 3. The two complementary H bonds between the molecules that constitute the dimer have an atomic separation that measures 2.386(2)
Å
(N3⋯H1Ai) and an interaction angle of 160.11(7)°
(N1i–H1Ai⋯N3)
[symmetry code: (i)
−x+
1, y, −z
+
1]. There is also a second, more unusual type of H-bonding interaction between each member of the dimer and the closest neighbouring molecule in the unit cell. More specifically, the X-ray data suggest that a somewhat weaker hydrogen bond between the pyrrole nitrogen lone pair of a neighbouring molecule (which is formally part of the delocalised π-electron system of the ligand) and the donor hydrogen atom H3A is likely. The H-bond distance measures 2.740(2)
Å
(N1ii⋯H3A) and has an interaction angle of 156.19(7)°
(N3–H3A⋯N1ii)
[symmetry code: (ii)
−x
+
1, y
−
1, −z
+
1]. A noteworthy caveat is that the latter contact distance is very close to the sum of the van der Waals radii of H and N (2.75 Å). Furthermore, the true position of the H atom in question is never completely certain even from high-resolution low-temperature X-ray diffraction data. Nevertheless, the geometry of this unconventional H bond is not altogether unreasonable if one considers the electrostatics of the interaction. Indeed, it may even be argued that hydrogen atom H3A interacts to some extent with the delocalized π-electron cloud of the neighbouring pyrrole ring.
![]() | ||
Fig. 3 H-bonding interactions between neighbouring molecules of 3 in the solid state. |
Interestingly, the recent X-ray crystal structure of 4,5-dimethyl-1,2-bis(2-pyrrolylmethylimino)phenylene, the 4,5-dimethyl analogue of 2, also shows a complementary H-bonding relationship between two Schiff base units within a discrete dimer.5 The H bonding is, however, quite different in that four H bonds exist between the pair of molecules in the dimer, consistent with there being two imine type nitrogen H-bond acceptor sites and two pyrrole NH group H-bond donor sites in the tetradentate Schiff base derivative. Furthermore, the tridentate Schiff base 1-{(E)-[(2-hydroxyethyl)imino]methyl}-2-naphthol, which has the donor atom combination ONO, also forms intermolecular H-bonded dimers in the solid state by virtue of a suitable combination of H-bond donor and acceptor sites within the molecule.21 A number of tri- and tetradentate Schiff base systems (both free bases and metal complexes22) therefore appear to be predisposed towards self-recognition by intermolecular H-bonding.23
Fig. 4 shows selected proton NMR spectra of 3 as a function of the total concentration of the Schiff base in solution. At low concentrations, the pyrrole NH proton resonance is very broad (line width=
275
±
83 Hz at 0.047 M) as a result of rapid relaxation in CDCl3 solution. Association by hydrogen bonding as the concentration of 3 in solution increases leads to a downfield shift of the signal and concomitant narrowing of the line width (to 67
±
4 Hz at 0.90 M), consistent with a slower relaxation and/or exchange rate for the pyrrole NH proton(s) in the dimer complex. It is also noteworthy that a single resonance is observed for the pyrrole NH proton at all concentrations. This suggests that the exchange rate for the monomer
↔
dimer equilibrium is fast on the 500 MHz 1H NMR time scale and that the spectrum is a weighted time-average of that due to each species in solution. Importantly, the methine proton resonance of the imine group (δH
=
8.33) retains its Larmor frequency and line width at all concentrations. Indeed, protons that do not change their magnetic environment with concentration are expected to give concentration independent signals. This signal therefore serves as an intense, visible reference point in the 1H NMR spectrum of 3 and highlights the marked changes in the pyrrole NH proton resonance that are brought about by intermolecular hydrogen bonding.
![]() | ||
Fig. 4 Selected scans of the pyrrole NH proton resonance for Schiff base 3 as a function of the total concentration in CDCl3 at 25![]() |
The 1H NMR spectra strongly suggest that Schiff base 3 dimerizes in solution according to the simple equilibrium shown in Scheme 2. If we assume that the observed chemical shift for the pyrrole NH proton is the equilibrium-weighted average of the chemical shifts of the monomer and dimer, then eqn. (1) can be used to fit the concentration dependence of the resonance in question:
δobs![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() | (1) |
![]() | (2) |
![]() | (3) |
fD![]() ![]() ![]() ![]() | (4) |
![]() | ||
Scheme 2 |
![]() | ||
Fig. 5 Unconstrained (solid line) and constrained (dashed line) non-linear least-squares fits of eqn. (5) to the concentration dependence of the pyrrole NH proton resonance line width in CDCl3
(25![]() |
As noted above, the pyrrole NH proton resonance shows a marked decrease in line width with increasing concentration. Eqn. (1) may therefore be rewritten to fit the concentration dependence of the line width:
log(Δν1/2)![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() | (5) |
A further point worth mentioning is that the line widths increased somewhat at Schiff base concentrations >1.0 M [e.g., log(Δν1/2)=
1.89(3) at a concentration of 1.54 M]. Evidently, additional line broadening over and above that expected from intrinsic transverse relaxation occurs at the highest concentrations studied. The most likely origin of this effect is instrumental since larger volumes of solvent were used to acquire spectra from the most concentrated solutions and some magnetic field inhomogeneity is likely if part of the solution lies outside the NMR probe receiver coils. We also examined the effect of adding water to the solution and found that although the pyrrole NH proton resonance broadened significantly with increasing H2O content, the chemical shift of the signal remained constant (Fig. S2, ESI). The latter experiments confirmed (1) that our procedure for drying CDCl3 by passage down a short column of activated basic alumina is adequate for removal of all but a trace quantity of H2O and (2) that the plot of chemical shift vs. Schiff base concentration shown in Fig. 4 indisputably reflects dimerization of 3 in this system rather than a water-dependent exchange process. However, we cannot rule out the possibility that some of the line width variation evident in Fig. 5 might be due to slight variations in the trace water content of the solutions used for the NMR study.
![]() | ||
Fig. 6 Least-squares fit of the DFT-calculated (grey) and experimental (black) structures of 3. The RMSD between the two conformations is 0.374 Å for all atoms. |
As noted earlier, the ArNH2 group N atom has mixed sp2 and sp3 hybrid character if the experimental bond angles subtended at the nitrogen atom are taken as being sufficiently accurate. We find that the values of the bond angles C11–N3–H3A and C11–N3–H3B measure 114.0° and 116.3°, respectively, for the in vacuo DFT-calculated conformation of 3. These calculated bond angles are within 2–4° of the experimental angles and support the notion that the amino group of 3 has a mixed hybridization state. An intriguing question is whether H bonding with the lone pair of electrons in the dimer structure of 3 changes the C–N–H bond angles, that is the relative contributions of sp2 and sp3 hybrid character for the amino group N atom. As discussed below, we have been able to obtain evidence in support of this idea from the calculated structure of the H-bonded dimer of 3.
![]() | ||
Fig. 7 Least-squares fit of the DFT-calculated (grey) and experimental (black) dimer structure of 3. The RMSD between the dimer structures is 0.413 Å for all atoms. |
Three geometrical features of the dimer were reasonably well modelled in the calculated structure: (1) the H bond distances, (2) the conformation of each monomer unit making up the dimer and (3) the hinged arrangement of the interacting monomer units. Thus, the mean calculated N3⋯H1A distance was 2.32(1) Å while the dihedral angle between the 14-atom mean planes of the two calculated monomer units measured 79.8°. The H bond distance deviates from the experimental distance by only 0.07 Å and is well within the expected level of performance of the B3LYP method for H-bond calculations.25 The calculated dihedral angle between the two planes of the interacting monomers deviates from the experimental value by some 15°. We attribute this mainly to the absence of steric or other crystal packing interactions in the simulation. In particular, it appears that the absence of packing interactions in the in vacuo simulation allows a significant degree of twisting of the molecule about the C7–C6–N2–C5 torsion angle (as noted above in the case of the monomer structure of 3). The driving force for this intramolecular twist is probably the nonbonded repulsion between H7 and H5, which is not damped to any significant extent in the simulated structure since the molecules are not in close contact with neighbouring molecules.
Interestingly, the conformation of each molecule within the dimer is calculated to be somewhat flatter than the in vacuo structure of the monomer itself. More specifically, the dihedral angle C7–C6–N2–C5 measures 8.6° and 12.8° for the two molecules making up the dimer. (The X-ray structure of the dimer comprises symmetry equivalent molecules, while the calculated structure of the dimer has C1 symmetry and thus independent monomer conformations.) These calculated values are within 4.5° and 8.7° of the experimental values, respectively. The better agreement between the calculated and experimental conformations of 3 when H bonding to at least one of the neighbouring molecules in the lattice is taken into account clearly shows that H bonding and nonbonded intermolecular interactions fine-tune the crystallographically observed conformation of this compound.
The mean calculated C–N–H angle for the ArNH2 group of 3 measures 113(2)° and compares favourably with the mean experimental value of 117(1)°. Interestingly, the calculated structure of the H-bonded dimer exhibits an amino group geometry that is somewhat closer to being pyramidal (sp3 hybridized) than planar (sp2 hybridized). This finding suggests that the H bond to the nitrogen lone pair of electrons slightly changes the character of the nitrogen atom relative to the in vacuo monomer. Indeed, comparison of the natural atomic orbital populations for N3 in the monomer and dimer structure of compound 3 shows that the 2s and 2p orbital populations are highest for the dimer (Table 3), consistent with a more localized set of hybrid orbitals on the nitrogen atom. Ultimately, the lower valence electron population for N3 of the monomer is best viewed as reflecting a slightly higher degree of hyperconjugation of the N atom lone pair with the aryl ring π-electron system.
Natural atomic orbital | Monomer | Dimer |
---|---|---|
2s | 1.3436 | 1.3699 |
2px | 1.3773 | 1.6956 |
2py | 1.7294 | 1.3781 |
2pz | 1.3819 | 1.4249 |
Σ2p | 4.4885 | 4.4986 |
Σ(valence electrons) | 5.8322 | 5.8685 |
The natural atomic charge distributions for the monomer and dimer (Fig. 8) clearly indicate the favourable (complementary) electrostatics of the H-bonding interactions in the dimer. Thus, the amino group nitrogen atom (N3) carries a significant negative charge that approaches a full electron. The pyrrole NH group hydrogen atom (H1A) is seen to carry the highest fractional positive charge in both the monomer and the dimer. It therefore stands to reason that these regions of high opposite fractional charge should interact favourably in the H-bonded dimer and, we propose, succinctly explain the driving force behind the formation of the dimer (at least in the solid state). Interestingly, N3 and H1A achieve a slightly higher charge polarization in the dimer relative to the monomer. The increase in negative charge on the amino group nitrogen atom in the dimer amounts to about 0.037 of an electron and possibly reflects a slight increase in sp3 hybrid character for this N atom.
![]() | ||
Fig. 8 Natural atomic charges for (a) the monomer and (b) the H-bonded dimer structures of compound 3. |
The DFT-calculated energies in Table 4 may be used to estimate the energy of formation of the H-bonded dimer, ΔEcomplexation, in the gas phase. Taking into account the appropriate counterpoise correction26 to eliminate basis set superposition errors, we found that formation of the dimer was strongly favoured on thermodynamic grounds in the gas phase (ΔEcomplexation=
−9.33 kcal
mol−1). The favourable complexation energy is, furthermore, paralleled by an increase in the frontier orbital energy gap, ΔE(LUMO–HOMO), of 0.04 eV over that for the monomer, consistent with net stabilization of the system. If one compares the calculated gas phase complexation energy to the value of ΔG measured for the associative interaction in solution by NMR spectroscopy (vide supra), then it is clear that solvation has a pronounced effect on the process and severely diminishes the stability of the H-bonded complex.
Monomer | Dimer | Molecule A | Molecule B | |
---|---|---|---|---|
a All energies are in eV. b Formation energy of the dimer in the gas phase calculated as the total energy of the dimer minus the sum of the energies of its non-interacting constituent molecules (Molecules A and B). The BSSE (basis set superposition error) correction to this energy is +0.1563 eV. | ||||
E total | −16056.4141 | −32113.2963 | −16056.3651 | −16056.3706 |
HOMO![]() ![]() |
−6.81787 | −5.91499 | −6.85732 | −6.86739 |
HOMO![]() ![]() |
−6.48833 | −5.90655 | −6.68916 | −6.66603 |
HOMO![]() ![]() |
−5.75743 | −5.04993 | −5.78056 | −5.77702 |
HOMO | −4.94734 | −5.01510 | −4.97564 | −4.97646 |
LUMO | −1.27269 | −1.30180 | −1.35106 | −1.34561 |
LUMO![]() ![]() |
0.55947 | −1.28058 | 0.55648 | 0.55484 |
ΔE(HOMO![]() ![]() |
3.67465 | 3.71329 | 3.62459 | 3.63084 |
ΔEcomplexationb | −0.5606 |
Finally, it is noteworthy that the dimer structure of 3 does not comprise two separate molecules at the electronic level. As shown by a plot of the HOMO for the dimer (Fig. 9), the highest filled molecular orbital has π-symmetry and spans both molecules. Importantly, the MO coefficients for chemically equivalent atoms in each molecule are significantly different, consistent with the symmetry of the dimer and the idea that this system does not simply consist of two “isolated monomers”. Notwithstanding the fact that we have used a basis set without diffuse functions that may have partly limited the accuracy of the calculations as a result, the best description for the dimer structure of 3 is that of a single supramolecular system with substantially delocalized π-electrons. As indicated by the data in Table 4, the HOMO is near degenerate with the HOMO−
1. The latter orbital is roughly the inversion-related analogue of the HOMO. This trend is evident for the other frontier orbitals, which are seen to occur in near-degenerate pairs. In the limit of exact inversion symmetry for the dimer, one would expect these orbitals to be doubly degenerate.
![]() | ||
Fig. 9 Drawing of the HOMO for the geometry-optimized (gas phase) dimer of 3. |
Footnotes |
† Electronic supplementary information (ESI) available: unit cell packing diagram for 3 (Fig. S1) and 1H NMR spectra of 3 as a function of the concentration of added H2O in CDCl3 solution (Fig. S2). See http://www.rsc.org/suppdata/nj/b3/b305946d/ |
‡ CCDC reference number 219663. See http://www.rsc.org/suppdata/nj/b3/b305946d/ for crystallographic data in .cif or other electronic format. |
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2004 |