Anna
Ignaczak
*a,
Stanisław
Porwański
*b and
Michalina
Szyszka
b
aDepartment of Theoretical and Structural Chemistry, Faculty of Chemistry, University of Lodz, Pomorska 163/165, 90-236 Łódź, Poland. E-mail: anignacz@uni.lodz.pl
bDepartment of Organic and Applied Chemistry, Faculty of Chemistry, University of Lodz, Tamka 12, 91-403 Łódź, Poland. E-mail: porwany@chemia.uni.lodz.pl
First published on 1st December 2016
The article presents the results of our more extended studies performed on one of the carbohydrate derivatives of azacrown ethers that were recently synthesized in our laboratory. An attempt is made to explain the interactions between 1,10-N,N′-bis-(β-D-ureidocellobiosyl)-4,7,13-trioxa-1,10-diazacyclopentadecane (host) and the aspirin molecule using theoretical methods. The structure and other properties of the complex formed by the two molecules revealed by NMR, IR spectrometry and microanalysis are compared to the results of quantum calculations performed with the semiempirical and density functional methods for several selected geometries.
In this paper, we briefly review derivatives of carbohydrate azacrown ethers that we had described before.5c–e This work also describes assessment of the strength of interaction between the receptor that contains in its structure two units of saccharides and one molecule of azacrown ether (host) and the aspirin molecule, using theoretical methods.
We have had the opportunity to demonstrate new ligand 1 (Fig. 1) in which the 4,7,13,16-tetraoxa-1,10-diazacyclooctadecane was substituted by two cellobiosyl units.5c This water soluble new receptor has been obtained by the tandem SAW reaction in high yield. It was established experimentally that the podand 1 interacts efficiently both with busulfan (1,4-butanediol-dimethylsulfonate), a powerful antitumor drug in leukemia, and with basic amino acids (L-lysine and L-arginine) to form 1:
1 supramolecular species.
The stoichiometries of the complexes were established on the basis of the Job plot method (also known as continuous variation method). Our studies showed two kinds of interactions between ligand 1 and busulfan (as guest): (1) van der Waals forces between lipophilic alkyl chain protons of the guest and azacrown protons; (2) hydrogen bonds between sulphonic groups and urea bridges.
Then we decided to obtain new macrocycles in which 4,7,13,16-tetraoxa-1,10-diazacyclooctadecane was replaced with 1,4,10-trioxa-7,13-diazacyclopentadecane.5d The received receptors include two cellobiosyl or two glucosyl units that are connected by urea linkers to the diazacrown ether platform (compounds 2 and 3 in Fig. 1). The binding studies allowed us to determine experimentally that such derivatives can form when dissolved in water host–guest type complexes with aromatic guests (aspirin and paracetamol). The complexation behaviors of receptors 2 and 3 towards neutral molecules were explored by 1H NMR spectroscopy in the mixture D2O/pyridine-d5 at 293 K. The molar ratio of the host/guest can be 1:
1 or 2
:
1 and is dependent on the structures of the guest and the host.
We have also demonstrated how the solvent and the size of the ether cavity affect the stability of the host:guest type complexes.5e We have used receptors 1–5 as hosts and p-toluenesulfonamide as the guest molecule. Compounds 2 and 3 have been applied as model ligands to study the solvent effect on chemical shifts for the selected proton signals in the 1H NMR spectrum. The biggest chemical shifts have been observed in the mixture D2O/pyridine-d5 for the cellobiose derivative 2 and in the mixture CD3OD/pyridine-d5 for the glucose derivative 3. Therefore we have studied the stoichiometry of compounds 3, 5/p-toluenesulfonamide in CD3OD/pyridine-d5 and 1, 2, 4/p-toluenesulfonamide complexes in D2O/pyridine-d5. The macrocycles 1, 2, 4, and 5 are able to complex with one molecule of the guest. Only ligand 3 formed the complex in 2:
1 stoichiometry (host/guest). The most stable complex has been obtained for receptor 2. We have confirmed that the supramolecular species are formed through weak non-covalent interactions.
In the present article we focus on the complex formed by cellobiose derivative 2 with aspirin. The host molecule 2 was synthesized in a high yield of 92% from per-O-acetyl-azido-β-D-cellobiose 6 and aza-crown ether 7 over two steps via a tandem Staudinger–aza-Wittig reaction, as shown in Scheme 1.5d
The host–guest properties of receptor 2 toward aspirin (Fig. 2) were explored by 1H and 13C NMR spectroscopy in D2O at 293 K. On the basis of the Job plot method the stoichiometry of the complex was established to be 1:
1. It was not possible to determine experimentally the exact structures of either the host molecule or its complex with aspirin because we did not obtain single crystals of them to perform X-ray diffraction analysis. Therefore, to expand the knowledge of these compounds, we have decided to use quantum methods. The theoretical approach allows one to explore various properties of chemical systems at the molecular level and also provides some values that can be compared to experimental data.
As no information is available about the structure of the host 2, it would be desirable to find the ground state of this molecule. However, this would require very extended studies, which we intend to perform in future. For our preliminary tests we have considered just three different types of possible structures. In model A the two cellobiosyl fragments are stretched away, so no hydrogen bonds can be formed between them. In the second model (model B) the cellobiosyl tails are oriented such that some interaction occurs between them, but they do not come very close. The third model (model C) mimics the very close proximity of the cellobiosyl parts and significant deformation of the azacrown ring in comparison to model A. Each structure was first optimized at the PM6 level, and then used as a starting point in the molecular dynamics conformational search combined with the PM6 method. The goal of this stage was to allow the molecule to adopt the geometry more stable in terms of energy, but still resembling the initial structure. The conformational search was performed using the tool available in the program Gabedit12 combined with the MOPAC program. We used the Stochastic Dynamics method via the Verlet algorithm. In each simulation the conformational space was explored at T = 1000 K during the period of 10 ps with a time step 1 fs. From the trajectory 30 lowest energy conformations were selected and subsequently optimized at the PM6 level. As a result we obtained three different low energy conformers A, B, C of the host molecule, which are shown in Fig. 6 in the Results section.
These structures were used in our further evaluation of the complexation energy in the host:guest complex. The number of significantly different structures of the complex is very large; therefore we limited our search to the cases where the aspirin molecule can interact with both the azacrown and cellobiosyl fragments. Several orientations of the aspirin molecule with respect to the host structures A, B, C were considered; for each the conformational search (using the same method as for the host molecule) was performed. The complexation enthalpy was then calculated as the difference between the standard heats of formation of the optimized products and substrates: ΔHcompl = HXf(host–guest) − Hf(host) − Hf(guest). Here X stands for A′, B′ or C′ – the type of the complex structure, similar to those shown in Fig. 7. The reference level was always the same, i.e. Hf(host) is the lowest value found from our tests for the heat of formation of the host molecule (for example, in the case of PM6 calculations this was structure C in Fig. 6).
The structures obtained from the PM6 tests served as starting configurations for calculations performed with two other semiempirical methods, PM6-DH2 and PM7, as well as with the hybrid density functional B3LYP-GD2 method. The latter is the standard B3LYP functional13 including the Grimme empirical pair-wise long range corrections.14 These calculations were performed with the 6-31G(d,p) basis set and the ultrafine integration grid using the Gaussian 09 program.15
In the case of PM7 and B3LYP-GD2 for each system full optimization (no constrains) was performed, while the PM6-DH2 energies were obtained from the single point SCF calculations, according to the strategy suggested by the authors of the MOPAC program. The DFT complexation energies were obtained using the same strategy as in semiempirical calculations; at this stage additionally the basis set superposition errors were calculated using the counterpoise correction (CP).16
For the structures optimized with the DFT method we have performed the vibrational analysis, which confirmed the stationary points to be true minima. Additionally, for all compounds the H and C NMR shielding tensors were calculated using the B3LYP/6-31++G(d,p) and the GIAO method. The H and C NMR chemical shifts δ (relative to TMS) were calculated using the procedure proposed by Tantillo17 according to the formula δX = (IX − σX)/(−SX), where X denotes H or C and σ represents the computed isotropic values. S and I are scaling factors defined as the slope and intercept of the line obtained for a given computational method from the linear regression analysis of experimental and computed chemical shift values. It was shown in a series of articles18 that such empirical scaling allows one to correct for systematic errors present in chemical shifts obtained from quantum calculations and achieve high-accuracy results. Using the files provided by Tantillo in ref. 17 for the test set of molecules we calculated the scaling factors for the methods we used (optimization with B3LYP-GD2/6-31G(d,p), followed by GIAO NMR calculations with B3LYP/6-31++G(d,p)). The following parameters were obtained in vacuo: SH = −1.0412, IH = 31.7502, SC = −0.9428, IC = 188.8146; and in water (PCM): SH = −1.0567, IH = 31.6557, SC = −0.9687, IC = 189.8139.
In our previous work5d some chemical shifts were given only in groups, without exact identification. In Table 1 we list chemical shifts we managed to assign to specific atoms of the host molecule. Together with other values from ref. 5d they constitute the data for comparison with computed results. In Fig. 3 the numbering of atoms used in assignment of chemical shifts is shown, while in Fig. 4 the overlapped NMR spectra for aspirin, 1,10-N,N′-bis-(β-D-ureidocellobiosyl)-4,7,13-trioxa-1,10-diazacyclopentadecane molecule (host 2) and their complex are presented. Signs of interaction were first detected by the chemical-induced shifts (CIS) of protons of the guest signals as compared to those of the free compound in D2O. The signals of both aromatic and methyl protons of aspirin are upfield shifted (lower ppm values) as shown in Fig. 4a. The anomeric (H-1), H-1′ and H-6 protons of the host are also shifted in the direction of lower ppm values (Fig. 4b). The values of changes in chemical shifts for selected atoms of the 2/aspirin complex are given later, in Table 3, together with the results of our quantum calculations. We had published before that in D2O/pyridine-d5 (9:
1) compound 2 formed the complex with aspirin also in 1
:
1 stoichiometry, but the signals of both aromatic and methyl protons of aspirin were downfield shifted.5d
1H | H-1 | H-2 | H-3 | H-4 | H-5 | H-6a | H-6b | H-1′ | H-2′ | H-3′ | H-4′ | H-5′ | H-6′a | H-6′b |
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
δ | 4.87 | 3.47 | 3.51 | 3.64 | 3.59 | 3.95 | 3.75 | 4.52 | 3.33 | 3.49 | 3.43 | 3.53 | 3.92 | 3.79 |
Fig. 5 presents the IR spectrum obtained for the host molecule. The lines for ν = 3420 and 1636 cm−1 were assigned to vibrations of O–H and CO bonds, respectively. More detailed assignment of other lines observed will be made on the basis of theoretical results.
In vacuo | In water | ||||||
---|---|---|---|---|---|---|---|
Type | PM6 | PM6-DH2 | PM7 | B3LYP-GD2 | PM6 | PM7 | B3LYP-GD2 |
A′ | −17.65 | −29.84 | −25.32 | −17.44 | −5.10 | −17.00 | −22.59 |
B′ | −12.42 | −27.74 | −33.13 | −42.05 | 8.15 | −6.98 | −38.24 |
C′ | −17.72 | −37.01 | −30.80 | −28.79 | 10.07 | −4.50 | −26.60 |
Aspirin | Calc. Δδ (A′-asp.) | Calc. Δδ (B′-asp.) | Calc. Δδ (C′-asp.) | Exp. Δδ |
---|---|---|---|---|
H-3as | −0.0078 | −0.0116 | 1.1655 | −0.0206 |
H-4as | −0.0811 | −0.0528 | 0.4452 | −0.0333 |
H-5as | −0.0929 | 0.6640 | −0.0300 | −0.0206 |
H-6as | −0.4869 | −0.0618 | −0.5812 | −0.0490 |
H-9as (CH3) | 0.0929 | 0.3636 | −0.2964 | −0.0107 |
Host | A′–A | B′–B | C′–C | |
---|---|---|---|---|
H-1 | −0.1396 | 0.2854 | 0.2444 | −0.0022 |
H-1′ | 0.0074 | −0.4668 | 0.0908 | −0.0031 |
H-6a | 0.0278 | −0.1147 | −0.1282 | −0.0020 |
H-6b | 0.0046 | 0.0186 | 0.0117 | −0.0020 |
As can be seen, the results are very inconsistent, even within the set obtained with semiempirical methods. Both PM6 and PM6-DH2 methods indicate that the most stable structures are the host and complex structures C and C′, while the lowest PM7 heats of formation are obtained for B and B′. The binding enthalpies obtained with PM6-DH2 and PM7 methods are much higher than the PM6 values. Obviously the results strongly depend on the method used.
The results obtained in vacuo at the highest computational level used in the present work, B3LYP-GD2/6-31G(d,p), corroborate the PM7 predictions, indicating conformers B and B′ (Fig. 8 and 9) as the lowest energy structures, but the agreement is only qualitative. When comparing values obtained with different methods one should remember that semiempirical methods are parametrized so as to give heats of formation at 298 K, while the DFT results are total energies corresponding to the temperature 0 K. The DFT complexation energy in B′ appears to be much higher than for two other structures.
The interaction between the host and guest molecules in the liquid phase will be certainly perturbed by the presence of solvent molecules. Therefore the PM6 and PM7 optimized structures were re-optimized in the presence of water modelled with the Conductor-like Screening Model (COSMO) using the values of 78.39 and 1.3 Å for the dielectric constant and the effective radius, respectively. The results differ significantly from those obtained in vacuo. First of all, both methods PM6 and PM7 predict the host conformer A (and also the complex A′) to be much more stable in water. According to the PM6 results in solution only formation of the complex A′ appears to be profitable; for the two other forms ΔHcompl values are positive. The complexation enthalpy of A′ in water is much smaller than that in vacuo. The PM7 method predicts much larger complexation energies than PM6, but substantially smaller than the corresponding values in vacuo. Both semiempirical methods predict the same trend in the complexation energies: A′ > B′ > C′, suggesting that in aqueous solutions the more “open” structure A′ is expected to be formed.
These findings were further verified at the DFT level. The starting structures obtained in vacuo with the B3LYP-GD2 method were re-optimized in the presence of water that was described by the polarizable continuum model (PCM) of solvent. The DFT results disagree with the semiempirical predictions. As in vacuo, also in water the host molecule B appears to have the lowest total energy, but among the three conformers studied the energy differences are much less pronounced than in vacuo (∼1–2 kcal mol−1).
For the complex the DFT results in aqueous solution again indicate the structure B′ to be the most stable, similar to the one in vacuo. The presence of water affects the total energies of the complexes A′, B′ and C′, and ΔEcompl in solution appears to be less differentiated than in the gas phase. In all structures studied the complex formation involves various interactions, among them there is always formation of some hydrogen bonds between aspirin and the cellobiosyl fragments. As shown in Fig. 9, in B′ three hydrogen bonds between aspirin and the host are observed, while there are only two in C′ and one in A′.
Of course, the continuum solvent model is very approximate and accounts basically for electrostatic interactions between the system and the solvent. Also, we used rather a modest level of theory. Inclusion of diffuse functions into the basis set at the optimization stage may affect the total and complexation energies. The reader should also remember that all the results presented above refer to some arbitrarily chosen structures, and thus should be treated as rather qualitative evaluation of binding energies in these complexes. The configurational space of such systems is vast, and therefore much more extended studies are necessary to determine the lowest energy structures for both the host molecule and the host–guest complex.
As can be seen in Fig. 10A–C, no exact match is observed for any of the conformers, although all exhibit some qualitative agreement. A closer resemblance seems to appear for the more compact structures B and C, except for the CO stretching mode in B, which corresponds to two lines separated by 55 cm−1.
The computed frequencies and intensities can be used to assign particular peaks to vibrational modes and this is partially done in Fig. 10. More details can be found in Table S1 in the ESI,† where vibrations corresponding to the highest (locally) intensities are listed. The vibrations quite often have a complex character, so the specified frequency corresponds to several different vibrations. In the area of low frequencies the dominant motion is the torsion of the CCOH angle, accompanied by the out-of-plane NH bend and various CH bending deformations. The higher intensity peak, covering the range 1028–1166 cm−1, corresponds to combined vibrations: mainly CC and CO stretches, coupled to COH and CCH bending. The next part of the spectrum, from 1288 to 1458 cm−1, covers again combined modes, in which prevail the COH and CCH bending modes. The peak at 1541 cm−1 refers to the CNH and NCH bending modes, while the next one refers mainly to the CO stretch, coupled however to the CNH and NCH bending. In this region are also present scissor modes of CH2 groups belonging to the azacrown fragment. The higher frequency modes are less complex and easier to assign: the CH stretch (2881, 2920 cm−1), the OH (strong) and NH (much weaker) stretches (broad peak around 3420 cm−1).
We also made an attempt to compare the changes in 1H chemical shifts obtained from the DFT calculations in water with the available experimental estimates in D2O. The DFT values presented in Table 3 are computed from δ averaged over atoms identified in measurements. A more detailed version of Table 3, with values for individual atoms in vacuo and in water, is given in Table S2 in the ESI.† Because in real solution different conformers can be present in various fractions, the change Δδ for a given complex X′ (X = A, B, C) was calculated with respect to the corresponding host structure X.
According to the experimental data the complex formation in D2O causes relatively small signal shifts toward smaller δ for both aspirin and host atoms. The calculated Δδ are in most cases larger and often have an opposite sign. Only the complex A′ gives Δδ < 0 for all aromatic H of aspirin, but the shift of the H-6as signal is too large. No agreement is found for aspirin CH3 and also selected H of the host molecule, the smallest discrepancy appearing again for A′, but the experimental Δδ are much smaller and are expected to be sensitive to subtle changes in the system.
There may be several sources of large discrepancy between experimental and computed NMR values. First of all, our theoretical study probes just a few structures of a certain type. The measured δ and Δδ values certainly depend on the equilibrium between various host and complex structures present in real solution. Another source of disagreement can be the implicit solvent model used in our calculations. An inclusion of specific interactions with neighboring water molecules and possible hydrogen bonds formed are also expected to affect the results. Additionally, unlike the calculated results, the experimental values are obtained for the deuterated host molecule. The isotope effect on chemical shifts is discussed in a number of works (e.g.ref. 19) and has been shown to be important in the systems containing hydrogen bonds, but mainly for substituted hydrogens or heavy atoms.
Footnote |
† Electronic supplementary information (ESI) available: More detailed analysis of vibrations observed in theoretical and experimental IR spectra; computed changes in chemical shifts upon complex formation. See DOI: 10.1039/c6nj03089k |
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017 |