Role of Weak Intermolecular Interactions in the Crystal Structure of

Russian Academy of Sciences, 47, Leninsky Prosp, Moscow 119991, Russia. Cryst. Growth Des. , 2014, 14 (9), pp 4439–4449. DOI: 10.1021/cg500533f...
1 downloads 14 Views 5MB Size
Article pubs.acs.org/crystal

Role of Weak Intermolecular Interactions in the Crystal Structure of Tetrakis-furazano[3,4-c:3′,4′-g:3″,4″-k:3‴, 4‴‑o][1,2,5,6,9,10,13,14]octaazacyclohexadecine and Its Solvates Published as part of the Crystal Growth & Design Mikhail Antipin Memorial virtual special issue Kyrill Yu. Suponitsky,*,† Konstantin A. Lyssenko,† Ivan V. Ananyev,† Andrei M. Kozeev,‡ and Aleksei B. Sheremetev‡ †

A. N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 28, Vavilov Street, Moscow 119991, Russia ‡ N. D. Zelinsky Institute of Organic Chemistry, Russian Academy of Sciences, 47, Leninsky Prosp, Moscow 119991, Russia S Supporting Information *

ABSTRACT: A single crystal X-ray diffraction study of macrocycle 1 and its solvates with dichloroethane and acetonitrile was carried out. Analysis of crystal packing based on geometrical criteria and intermolecular interaction energies obtained from topological analysis of the experimental electron density and quantum chemical calculations allowed for a detailed description of peculiarities of the crystal packing of compound 1 and its modification upon solvate formation. Crystal packing of solvates of this conformationally rigid macrocycle can be successfully explained by a shape similarity principle as well as by consideration of stabilization energy of the molecule in crystal that allows one to get insight into the formation of other solvates or cocrystals of the macrocycle 1.

1. INTRODUCTION Investigation of organic macrocycles has long attracted considerable attention of scientists owing to the wide range of their applications such as in supramolecular chemistry, molecular recognition, catalysis, photochemistry, and medicine.1 The properties of macrocycles are highly dependent on the nature and number of donor atoms in it as well as on the conformation and dimensions of the cavity. Oligoazobenzenophanes, which are supramolecular structures comprised of aromatic units with azo bridges, are an interesting class of macrocyclic compounds that can adopt different conformations as a result of E-Z isomerization of the azobenzene moiety.2 An important point is that the E and Z forms of the azobenzene backbone can be easily and reversibly photo- or thermoisomerized and differ significantly in length. Therefore, azobenzene is widely used as photochromic moiety in many types of molecular switches.3 Whereas extended azobenzene macrocycles are well-known, azoheterocycle-based macrocycles are rare.4 At the same time, the growth of azoles chemistry over the last decades has been significant, mainly owing to the importance of azoles in medicinal and material chemistry. Recently, polydentate aromatic nitrogen heterocycles have emerged as building blocks for macrocycle construction.5 Most of the so far reported azo-macrocycles, in which aromatic units in the azobenzene moiety are replaced by azole units, have been based on the azofurazan backbone.6 © 2014 American Chemical Society

In 1996, the macrocycle built up of four furazan (1,2,5oxadiazole) rings linked by four azo bridges, tetrakis-furazano[3,4-c:3′,4′-g:3″,4″-k:3‴,4‴-o ][1,2,5,6,9,10,13,14]octaazacyclohexadecine 1, known as TATF (Scheme 1), was described and two methods of its preparation have been elaborated.6c,d,f,g Macrocycle 1 has a high heat of formation, and its use as a component of high energetic compositions for variety applications has been investigated.7 Electron-withdrawing properties of macrocycle 1 also make it interesting for biological applications; its utilization as an inhibitor of soluble guanylate cyclase has been patented.8 To the best of our Scheme 1

Received: April 17, 2014 Revised: July 30, 2014 Published: August 7, 2014 4439

dx.doi.org/10.1021/cg500533f | Cryst. Growth Des. 2014, 14, 4439−4449

Crystal Growth & Design

Article

Table 1. Crystallographic Data for Macrocycle 1 and Its Solvates 1·0.5DCE and 1·MeCN

a

parameter

1

1·0.5DCE

1·MeCN

empirical formula fw crystal system space group a, Å b, Å c, Å β, deg V, Å3 Z dcalc, g·cm−3 μ, mm−1 F(000) θ range, deg reflections collected independent reflections Rint refined parameters completeness to theta θ, % GOF (F2) reflections with I > 2σ(I) R1(F) (I > 2σ(I))a wR2(F2) (all data)b largest diff peak/hole, e·Å−3

C8N16O4 384.24 monoclinic C2/c 25.5408(3) 6.80990(10) 16.3783(2) 95.3730(10) 2836.17(6) 8 1.800 0.151 1536 2.50−62.88 272196 23207 0.0395 253 99.5 1.055 17087 0.0418 0.1287 0.594/−0.387

C8N16O4·0.5C2H4Cl2 433.72 monoclinic C2/c 30.4180(5) 7.08340(10) 16.0390(2) 108.6860(10) 3273.65(8) 8 1.760 0.300 1736 1.41−60.00 248117 24590 0.0343 280 99.1 1.145 17842 0.0402 0.1371 0.672/−0.589

C8N16O4·C2H3N 425.29 monoclinic P21/n 15.1136(17) 6.5756(8) 18.463(2) 113.538(2) 1682.2(3) 4 1.679 0.138 856 2.24−28.00 11417 4035 0.0404 281 99.3 1.006 2920 0.0391 0.0961 0.331/−0.244

R1 = ∑||F0| − |Fc||/∑|F0|. bwR2 = (∑[w(F02 − Fc2)2]/∑[w(F02)2]1/2.

(MeCN) solvates by the X-ray diffraction method and described the details of the crystal packing of 1 and peculiarities that govern its modification upon formation of solvates by different approaches; those include analysis of shortened contacts and intermolecular interaction energies based on experimental electron density distribution function and quantum chemical calculations.

knowledge, oxidation of the azo group to the azoxy group of the macrocycle 1 is the only known reaction of this compound.7f,9a A few analogues of macrocycle 1 in which one or two azo linkages between the furazan units are replaced by heteroatoms, such as oxygen6e,l,10 and sulfur,6l or by a heterocycles unit,6h,m,n,11 have been described. Some of those analogues were obtained in the form of solvates with DMSO6h and AcOH,6l and were characterized by X-ray crystallography. It is well-known that the solid state properties and behavior of the given compound in the form of solvate or cocrystal with other compounds can vary dramatically;12−14 this is of paramount importance if the material has a potential for the applications in medical or high energetic industry.13,14 Cocrystallization can either improve or deteriorate the desired properties; therefore, the development of approaches that can control the formation of a cocrystal is an important challenge.15 It should be also noted that attempts to predict a cocrystal formation are usually based on the use of relatively strong intermolecular interactions such as hydrogen or halogen bonds or strong π···π stacking interactions.16 In the case of compounds that form weak nonbonded interactions only, the predictive power of this approach becomes extremely low. Therefore, an analysis of intermolecular interactions in crystals of individual compounds and study of their redistribution upon formation of a solvate can provide valuable information for optimization of the pathways toward crystal structures with high density for potential applications as energetic materials. For instance, investigation of solvates of the well-known explosive HMX17 allowed the peculiarities of intermolecular interactions to be described in detail, and this information was used for the design of HMX cocrystals with other energetic materials.18 In the present work, we have investigated crystal structure of macrocycle 1 and its dichloroethane (DCE) and acetonitrile

2. EXPERIMENTAL SECTION 2.1. Synthetic Procedures. Macrocycle 1 has been synthesized by oxidation of 3,4-diaminofurazan with dibromoisocyanurate according to the published procedure.6g Red-orange crystals of compound 1 with mp 214−216 °C (lit.6c mp 208−210 °C) have been obtained by crystallization from CCl4. IR, MS, and 13C NMR spectra are identical to those reported in the literature.6c 2.2. Single-Crystal X-ray Diffraction Study. X-ray experiments were carried out using SMART APEX2 CCD (λ(Mo−Kα) = 0.71073 Å, graphite monochromator, ω-scans) at 100 K. Collected data were analized by the SAINT and SADABS programs incorporated into the APEX2 program package.19 All structures were solved by the direct methods and refined by the full-matrix least-squares procedure against F2 in anisotropic approximation. The positions of hydrogen atoms of solvents were calculated and included in the refinement within isotropic approximation by the riding model with the Uiso(H) = 1.5Ueq(Ci) for methyl groups and 1.2Ueq(Cii) for other carbon atoms, where Ueq(C) are the equivalent thermal parameters of the parent atoms. In 1·0.5DCE, the dichloroethane molecule is disordered over two positions with occupancies equal to 0.72(2):0.28(2); both parts were refined in anisotropic approximation. The refinement was carried out with the SHELXTL program.20 The details of data collection and crystal structures refinement are summarized in Table 1. In addition to ordinary refinement by the SHELXTL program given in Table 1, for TATF and its solvate with DCE, the quality of the single crystals and diffraction data allowed for the multipole refinement which was carried out within the Hansen−Coppens model.21 The total pseudostatic charge density distribution ρ(r) was calculated as a sum of pseudoatomic charge densities. The refinement in this study was 4440

dx.doi.org/10.1021/cg500533f | Cryst. Growth Des. 2014, 14, 4439−4449

Crystal Growth & Design

Article

done with XD2006 software.22 The multipole expansion was truncated at the octupole level (l = 3) for heavy atoms (for the chlorine atom, hexadecapoles were taken into account as well); for hydrogen atoms only the populations of monopoles and D10 harmonics were refined. In each case, we used low-angle reflections with sin θ/λ ≤ 0.904 Å−1 to refine multipole parameters and high angle reflections with θ/λ ≥ 0.7 Å−1 for the refinement of positions and anisotropic displacement parameters of heavy atoms. Multipole and monopole populations and corresponding κ coefficients in 1 were refined separately at the initial steps of the refinement and together at the final steps. Taking into account the presence of the statistical disorder in 1·0.5DCE, as the starting model for the refinement, we used all parameters obtained for 1. We have found that it leads to faster convergence and slightly lower R-factor values. Furthermore, such procedure facilitates the correct refinement of the disordered molecule of dichloroethane. The overall quality of the experiment and the refinement are supported by the analysis of differences of mean-squares displacement amplitudes (DMSDA) along interatomic vectors in the molecule;23 those values do not exceed 6 × 10−4 Å−2 and 4 × 10−4 Å−2 in 1 and 1·0.5DCE, respectively. The value of DMSDA for disordered DCE molecule, as had been expected, exceeded the threshold value and was equal to 2.6 × 10−3 Å−2. Electron density residuals were randomly distributed in the unit cell of 1 and did not exceed 0.20 e·Å−3, although they are slightly higher (0.3 e·Å−3) in the solvate in which this maxima is located in the neighborhood of the disordered molecule (Figure 1S, Supporting Information illustrates residual density). The refinement was carried out against F and converged to R = 0.0196, Rw = 0.0157, and GOF = 1.32 (for 12436 merged reflections with I > 3σ(I)) for 1, and to R = 0.0259, Rw = 0.0178, and GOF = 1.439 (for 17625 merged reflections with I > 3σ(I)) for 1·0.5DCE. The estimation of the kinetic energy [g(r)] was based on the Kirzhnits’s approximation24 relating it to the values of the ρ(r) and its derivatives: g(r) = (3/10)(3π2)2/3[ρ(r)]5/3 + (1/72)|▽ρ(r)|2/ρ(r) + 1/6▽2ρ(r)]. The use of this relation in conjunction with the virial theorem (2g(r) + v(r) = 1/4▽2ρ(r)) results in the values of potential energy density [v(r)] in the critical points from experimental diffraction data. The reliability of this approach in the case of closed-shell and intermediate type of interatomic interactions was demonstrated for various compounds.25−27 For the critical point search in intermolecular areas, we have used the following procedure: (1) each atom was surrounded by the cluster with the radii of 6 Å, and each contact with a distance up to 4 Å has been analyzed; (2) for all critical points found, we calculated bond paths in order to verify for what particular pair of atoms the interaction occurs. By means of this procedure, we have checked all interatomic interactions and thus obtained the molecular graph of the supramolecular organization. Selected molecular graphs illustrating the bond paths are depicted in Figure 2S, Supporting Information. The atomic charges obtained by integration of the ρ(r) function over the atomic basins (Ω) and volumes for 1 and 1·0.5DCE are summarized in Table 1S, Supporting Information. Charge leakage as small as 0.01e clearly illustrates high accuracy of the numerical integration used to obtain these values. Sum of the volumes of atoms, constituting macrocycle, becomes slightly smaller for 1·0.5DCE (350.01 Å3) in comparison to that for 1 (354.05 Å3). For the majority of atoms, the difference in atomic volumes is nearly negligible with the only exception for O4, N6, N7, N8 for which the values increase by 1.9−2.3 Å3. The atomic volume of DCE molecule is equal to 89.31 Å3. The atomic charges for all atoms in 1 and 1·0.5DCE do not differ by more than 0.06e. As an additional indication of the quality and, as a consequence, the reliability of the results obtained, the deformation electron density (DED) maps for both compounds show the expected features; those are the accumulation of DED in the areas of chemical bonds and electron lone pairs of nitrogen and oxygen atoms (Figure 3S, Supporting Information). The topological analysis of the experimental ρ(r) distribution and data visualization were performed with the WinXPRO program suites.28 2.3. Computational Details. Quantum chemical calculations have been carried out using the Gaussian program29 within density

functional theory (DFT). On the basis of reported success of M052X functional to describe the energy of intermolecular interactions30−33 and the results of our recent study on isolated molecules and molecular aggregates,34−37 here we used the M05-2X/aug-cc-pvdz level of theory.

3. RESULTS AND DISCUSSION Molecular crystals are always stabilized by numerous noncovalent interactions of a different nature. Among them, hydrogen and halogen bonds are the strongest ones.16a,b,38 In those cases when H-bonds lead to a formation of 1-D or 2-D framework, their structures can be predicted based on statistical analysis of the data retrieved from Cambridge Structural Database39 or by quantum chemical methods.40 However, stabilization of a 3-D structure is also provided by weak intermolecular interactions such as C−H···O, C−H···N, peakhole (n···π or π···π), and van-der-Waals, and an influence of those weak interactions can hardly be predicted.41 It should be noted that the presence of the strong (or relatively strong) intermolecular interactions will not necessarily lead to a dense crystal packing. For instance, in our recent studies on aminofurazans and tetrazines, it was observed that crystal packing density does not exceed 1.6 g/cm3 if relatively strong H-bonds are formed in the crystal structure.6j,42 It is logical to suggest that in the absence of strong intermolecular interactions, molecules have more freedom and a better opportunity to properly adjust their relative orientation and come closer to each other. A number of weak interactions is usually quite large, and the sum of their energies can be comparable to the energies of a few strong interactions that can be formed in a crystal. Recent studies on nitro-substituted furazans have shown that a high density of their crystal packing is provided by a large number of weak intermolecular interactions involving nitro groups43−46 which is in an agreement with Kitaigorodsky aufbau principle.47 The latter is based on the tendency of molecules to form dense crystal packing and on the shape similarity principle. It should however be noted that aufbau principle can be violated if strong Hbonds are observed in crystal structures.48 Macrocycle 1 does not have any strong acceptors of a lone pair; therefore, strong intermolecular interactions are not expected in its crystal structure. Early X-ray study of macrocycle 16f has revealed its relatively dense crystal packing (1.80 g/cm3 at 153 K), but no description of the crystal packing is given both in CSD and in the article.6f Nevertheless, it can be suggested that the crystal structure of 1 is stabilized by numerous weak intermolecular interactions. This suggestion, however, needs to be verified. Another point that is worth studying is how intermolecular interactions would be redistributed if a solvent molecule is included in the crystal structure. To answer those questions as well as to search for peculiarities and driving forces of the crystal structure formation, we have carried out a single crystal X-ray diffraction study of the individual macrocycle 1 and its two solvates with dichloroethane and acetonitrile, which also do not have functional groups able to form strong intermolecular interactions with the macrocycle. Crystallization of macrocycle 1 was carried out from a variety of solvents, namely, carbon tetrachloride, chloroform, methylene chloride, tetranitromethane, benzene, dichloroethane, and acetonitrile. Solvates of macrocycle 1 were obtained only for two latter solvents in the ratios of 1:0.5 and 1:1, respectively. From all other solvents, single crystals of the individual 4441

dx.doi.org/10.1021/cg500533f | Cryst. Growth Des. 2014, 14, 4439−4449

Crystal Growth & Design

Article

Figure 2. Projection of the crystal packing of the macrocycle 1 onto the ac plane. Molecules of the closest surroundings of the central molecule M0 are denoted by capital letters and correspond to those in Table 2. Letters in parentheses correspond to molecules located above or below the plane of the figure (obtained by a translation along the axis b). Selected layers L1, L2, L3, L4 discussed in the text are shown schematically by color lines.

the macrocycle. It is characterized by a nonplanar structure (Figure 1) which resembles that of calix[4]arene in an 1,3alternate conformation. Comparison of molecular geometries shows that the presence of a solvent has no influence on the molecular structure. Conformation of the macrocycle can be described by selected torsion angles presented in Table 2S (in the Supporting Information). Torsion angles about double N N bonds are close to 180°, and those about C−C bonds are close to 0°. The exocyclic angles C−C−NN from the one side of each furazan cycle are nearly 0°, while angles from the other side are found in the range of 125−145°. Quantum chemical calculation of the isolated macrocycle leads to nearly the same molecular structure. For the description of the crystal packing, we used the following approaches. The first one is the most popular and easy-to-use visual analysis that is based on a consideration of the shortened intermolecular distances; this is quite sufficient for the majority of goals. However, for a detailed study on peculiarities of the crystal packing, the approaches based on intermolecular interaction energies should also be used. This approach allows one to identify crystal packing motif based on its stabilization energy and was introduced by Zorky et al.49 As such, we used topological analysis of the electron density (for 1 and 1·0.5DCE, the quality of their single crystals allowed us to carry out multipole refinement to obtain experimental electron density distribution function) and estimated energies of intermolecular interactions based on their correlation with the potential energy density (V(r)) in the bond critical point (BCP) of electron density (Econt = 1/2V(r))27,50 (see also Supporting Information for use of the recently proposed formula Econt = 0.429 × G(r) based on correlation of Econt with the kinetic energy density (G(r)) in BCP).51 The third

Figure 1. General view of the macrocycle 1 (a) and its solvates with DCE (b) and MeCN (c) in representation of atoms by thermal displacements ellipsoids with 50% probability. Minor part of the disordered DCE is shown by open lines.

macrocycle 1 were grown. For the crystal structures of macrocycle 1 and solvates 1·0.5DCE, and 1·MeCN, an asymmetric unit cell contains one independent molecule of 4442

dx.doi.org/10.1021/cg500533f | Cryst. Growth Des. 2014, 14, 4439−4449

Crystal Growth & Design

Article

Table 2. Close Contacts (Å) and Intermolecular Interaction Energies (kcal/mol) as Obtained from Topological Analysis of the Electron Density (ED) and Quantum Chemical (DFT) Calculations for 1 and 1·0.5DCEa individual macrocycle 1

solvate 1·0.5DCE

N

close contact

distance

energyb (ED)

energy (DFT)

A

N2···N12 N12···N2 N3···N3 N3···O2 O2···N3 N9···N9 N9···C3 C3···N9

3.130 3.130 2.928 3.073 3.073 3.158 3.288 3.288

−6.4

−6.2

B

O4···N11 O4···N14 O4···C4 N7···C7 N13···N8c N11···O4 N14···O4 C4···O4 C7···N7 N8···N13c N1···N11 C1···N5 C2···N5 N1···N4c N15···N5 N11···N1 N5···C1 N5···C2 N4···N1c N5···N15 C1···N3 C6···N4 C5···N4

2.963 3.047 3.014 3.412 3.229 2.963 3.047 3.014 3.412 3.229 3.039 3.168 3.265 3.274 3.201 3.039 3.168 3.265 3.274 3.201 3.073 3.161 3.247

−4.2

G

N3···C1 N4···C6 N4···C5

H

I

C

D

E

F

J K L M N S0

distance

energyb (ED)

energy (DFT)

3.089 3.089 2.946 3.108 3.108 3.091 3.287 3.287 3.199 3.199 3.290 3.290 3.541

−7.2

−6.0

−4.3

N2···N12 N12···N2 N3···N3 N3···O2 O2···N3 N9···N9 N9···C3 C3···N9 N2···N5 N5···N2 N2···C5 C5···N2 O4···N8c

−0.3

−1.0

−4.2

−4.3

N8···O4c

3.541

−0.3

−1.0

−4.0

−3.1

N1···N11 C1···N5 C2···N5 N1···N4c

3.092 3.109 3.143 3.230

−3.2

−3.1

−4.0

−3.1

N11···N1 N5···C1 N5···C2 N4···N1c

3.109 3.143 3.092 3.230

−3.2

−3.1

−2.4

−2.8

−2.3

−2.4

−2.8

−3.3

−2.3

O3···N7c N7···O3c N6···N13c N13···N6c

3.202 3.202 3.309 3.309

−3.0

−2.4

−4.6

−3.2

N15···N6c N1···O3c N6···N15c O3···N1c N2···N2 N2···N4c N4···N2c O1···O1

3.293 3.287 3.293 3.287 3.157 3.406 3.406 2.791

−1.4

−1.6

−1.5

−1.3

−1.4

−1.6

−1.5

−1.3

−1.0 −0.6 −0.6 −2.1

−1.5 −0.7 −0.7 −0.3

3.172 3.084 3.096 3.216 3.172 3.084 3.096 3.216 2.994 2.994 3.137 3.137 3.258 3.258 3.105 3.333 3.105 3.333 3.185 3.602 3.602 2.832 3.387 3.546 2.563 2.997

−3.3

3.073 3.161 3.247

C1···N3 C6···N4 N16···O2 N16···N3 N3···C1 N4···C6 O2···N16 N3···N16 O4···N6 N6···O4 N7···N13 N13···N7 N7···C6 C6···N7 N15···N6 N1···O3c N6···N15 O3···N1c N2···N2 N2···N4c N4···N2c O1···O1 N14···Cl1 N11···Cl1c N13···H1Sa N8···H1Sbc

−0.9 −0.4 −0.4 −1.9 −1.5

−1.3 −0.6 −0.6 −0.2 −2.9

−1.8 −1.6

−3.6 −0.9

S1 S2

4443

close contact

dx.doi.org/10.1021/cg500533f | Cryst. Growth Des. 2014, 14, 4439−4449

Crystal Growth & Design

Article

Table 2. continued individual macrocycle 1 N

close contact

distance

b

energy (ED)

solvate 1·0.5DCE energy (DFT)

close contact c

N8···Cl1

distance

energyb (ED)

energy (DFT)

3.435

a

Interaction energies are given for the closest environment of the central molecule M0 (Figure 2) for which either a close contact or an interaction energy higher than 0.5 kcal/mol is observed (more detailed information is given in Tables 3S and 4S in the Supporting Information); energies for macrocycle···solvate contacts are given for the major part of the disordered DCE. bEnergies obtained using Econt = 1/2V(r) formula. cContacts for which BCPs were found, but interatomic distances exceeded the sum of nonbonded atomic radii53 are shown in bold.

Figure 3. General view of the columns formed along the 2-fold axis in the crystal structures of 1 (a), 1·0.5DCE (b), 1·MeCN (c).

approach is also based on the energetic criteria of intermolecular interaction, but the energies are obtained by quantum chemical calculation of a series of pairs of the central molecule (M0) with its neighboring molecules using the geometry taken from the X-ray data.52 Crystal packing of the individual macrocycle 1 is depicted in Figure 2, while close contacts and intermolecular interaction energies are given in Table 2. From the standpoint of the visual analysis, the crystal packing motif in 1 can be described in several ways based on close intermolecular contacts. Along the 2-fold axis, molecules form columns (Figure 3) by interaction of the nitrogen atoms of the furazan ring with the π-systems of the two furazan rings of the neighboring molecules (F, G) (n···π type of interaction) with a relatively high energy. Because of the nonplanar structure of the macrocycle, these columns contain voids which are filled with the molecules from the adjacent columns (Figure 3, M0···B, M0···C) thereby leading to interpenetrating double columns along the 2-fold axis. The πsystems of both the azo-bridges and furazan rings are involved in the interaction between those columns with the energy being higher than that inside the column (both n···π and π···π types

contribute to the energy of these interactions, but contribution of the n···π type is much higher). The strong interactions are also formed with the other adjacent columns (M0···A, M0···D, M0···E, M0···H), and the highest energy is observed for M0···A molecular pair. Again, the most significant contribution to this interaction comes from the n···π type. The same is true for M0···D, M0···E molecular pairs, while for M0···H, both n···π and π···π types contribute nearly equally. Along the axis a, interpenetrating columns interact with the adjacent column A to form layers (L1, see Figure 2) parallel to the ab crystallographic plane. On the other hand, along the axis c, those columns interact with molecules D and E to form double layers (L2) in the bc plane. Another double layer (L3) is obtained by interactions that include molecules A, D, E, F, G, N. One can also separate ordinary layers (L4) formed by M0··· D, M0···E, M0···F, M0···G interactions which are twice as thin as L2. On the basis of the inspection of the close contacts only, it is not clear which layers are more stable, and difference in their stability is unknown. An analysis of the crystal packing in terms of interaction energies shows that stabilization energy of the selected layers is equal to −10.9, −13.0, −12.2, −7.5 kcal/ 4444

dx.doi.org/10.1021/cg500533f | Cryst. Growth Des. 2014, 14, 4439−4449

Crystal Growth & Design

Article

Macrocycles 5 and 6 also adopt a nonplanar conformation. Instead of interpenetrating columns, they form 3-D conglomerate structures, again, by means of interactions of n···π and π···π types with the former prevailing over the latter. Before the following discussion, we have to comment on the origin of some discrepancy observed between the interaction energies obtained in two ways (Table 2). Within quantum chemical approaches, an energy of a dimer formation from its constituted monomers is usually estimated by the well-known formula Eint = Edimer − Emonomer1 − Emonomer2. An obvious deficiency of such an approach for the application to analysis of the crystal packing is consideration of the bare dimer which leads to two types of possible errors. The first one is a neglect of polarization of the given dimer by its surroundings in the real crystal. The second one is related to incorrect description of the electron density in the area located between interacting molecules. In the crystal, electron density at any given point arises from contribution of all molecules while in the bare dimer only two molecules are taken into consideration. In the case of an approach based on the topological analysis of the electron density, an uncertainty of the estimated energy might arise from an empirical nature of the Econt = 1/2V(r) formula. Therefore, one should not expect quantitative agreement between the results obtained by two approaches. At the same time, the results from Table 2 as well as comparison of the energies of the selected layers clearly demonstrate that both ways lead to comparable values, and semiquantitative agreement is observed. The worst agreement was found for the M0··· N interaction due probably to the above reasons. The following discussion is based on quantum chemical energies because for the 1·MeCN solvate, reliable electron density distribution function cannot be obtained from the X-ray experimental data. Solvate 1·0.5DCE crystallizes in the same space group as the pure compound 1 does, and two unit cell parameters (b and c) are nearly equal for 1 and 1·0.5DCE, while the axis a is ca. 5 Å longer. The DCE molecule is incorporated into the crystal structure so that it occupies the empty space inside the columns of the macrocycle (Figure 3) thereby moving apart interpenetrated columns in the crystallographic direction a. This also leads to some relative shift of the layers parallel to the bc crystallographic plane. If one would consider the structure of macrocycle 1 as formed by those layers (L4) then an inclusion of DCE would cause layers to separate. As a result, the M0··· B,C interaction becomes significantly weaker, the M0···H interaction becomes somewhat stronger, and a small decrease is observed for interaction energy inside the column (M0···F,G), while all the other intermolecular pair interactions are affected by a solvent incorporation to a much lesser extent (Table 2, Figure 4a), and the system of close contacts between molecules of corresponding pair is not significantly changed. It should be noted here that the M0···B,C intermolecular interaction (which breaks upon solvate formation) is one of the strongest in the crystal structure of 1. The inclusion of the solvent molecule causes the layers L2 and L1 which have higher stabilization energy to break, while less stabilized layers L3 and L4 remain almost unchanged that looks somewhat unexpected. However, from the standpoint of the Kitaigorodsky auf bau principle, location of a relatively small solvate molecule in the voids formed by a columnar structure along the 2-fold axis seems logical. The fact that crystallosolvates of the macrocycle were obtained from two solvents only (out of seven probed for crystallization) confirms an importance of the shape similarity

Scheme 2

mol for L1, L2, L3, L4, respectively (energy is obtained based on DFT calculations (see below) as a half of the sum of pair energies of the interaction of M0 with its closest neighbors inside the layer). Direct comparison of stabilization energies of ordinary and double layers is not straightforward; however, one can compare EL1 with EL4 and EL2 with EL3. For comparison of the two methods used for energy estimation, we also calculated layer energies based on experimental electron density distribution, using Econt = 1/ 2V(r) formula, which are equal to −10.4, −13.5, −13.1, −7.8 for L1, L2, L3, L4, respectively, and do not differ significantly from the DFT results, showing the same trends. On the basis of interaction energies, the crystal packing of 1 is better described as a 3-D conglomerate rather than a columnar or layered structure as the differences in energies in the a, b, c directions are not very pronounced. We also compared the types of intermolecular interactions that are observed for previously reported macrocycles incorporated furazan rings (Scheme 2).6l,m,9c In contrast to macrocycle 1, macrocycles 2 and 3 adopt a nearly planar conformation. In their crystal structures, both compounds form parquet layers by means of n···π interactions, while π···π interactions are observed between the layers (Figure 4Sa in the Supporting Information). The compound 2 was also obtained as a solvate with acetic acid; in this case, the π···π interaction pattern does not change significantly, while the n···π interactions are now observed between the macrocycle and the solvate molecule only. Acetic acid forms H-bonded dimers, and therefore no hydrogen bonds are observed between the solvate molecule and the macrocycle (Figure 4Sb in the Supporting Information). Macrocycle 4 adopts a nonplanar conformation that resembles that of the macrocycle 1, and its crystal structure is very different from those of 2 and 3, and it has some similarities to that of 1. It consists of interpenetrating columns that are stabilized by n···π and π···π interactions, and the same types of interactions connect columns into a 3-D structure (Figure 5S in the Supporting Information). 4445

dx.doi.org/10.1021/cg500533f | Cryst. Growth Des. 2014, 14, 4439−4449

Crystal Growth & Design

Article

Figure 4. Crystal packing fragments of 1·0.5DCE (a) and 1·MeCN (b). Molecules of the closest environment of the central molecule M0 are denoted by capital letters and correspond to those in Table 3. Letters in parentheses correspond to molecules located above or below the plane of the figure (obtained by a translation along the axis b).

and shape similarity principle, and on a consideration of the intermolecular energy) explain the observed structure of 1· 0.5DCE. However, from a consideration of the pair intermolecular energies of the individual macrocycle 1, location of a solvent in the above-mentioned voids is not evident. Changes in the crystal packing of 1 upon formation of the 1· MeCN solvate which crystallizes in P21/n space group are more pronounced (Figure 4b, Table 3). Nevertheless, columns along the 2-fold axis are similar to those in 1 and 1·0.5DCE. Again, MeCN molecules are located in the same voids as DCE (Figure 3) thereby causing layers L4 to separate. Because of some differences in the size and shape of these solvents (one DCE molecule occupies less volume than two MeCN molecules) separation between the layers is larger, and energy of M0···B,C interactions is negligible. While the packing diagrams in Figure 4 seems quite similar, more detailed analysis reveals some differences. Similar to 1 and 1·0.5DCE, the columns along the 2-fold axis are bound into

principle for the solvate or, possibly, for the cocrystal formation. As expected, DCE molecules do not form any strong intermolecular interactions with the macrocycle. However, weak Cl···N (of the n···π type) and C−H···N interactions partly compensate a loss of energy due to weakening of M0···B,C interactions. In total, the energy of interactions of the central molecule M0 with its closest macrocycle molecules is decreased by 7.9 kcal/mol. Sum of the energies of M0···solvent interactions is equal to 7.4 kcal/mol. However, the stabilization energy of 1·0.5DCE that is equal to 1/2Emacrocycle···macrocycle + Emacrocycle···solvent is increased by 3.5 kcal/mol and, therefore, the solvate appears to be more stable than the individual macrocycle 1. It might be guessed that location of the solvent in other sites would lead to weakening of more than two relatively strong intermolecular interactions and, as a consequence, would result in a decrease in the crystal structure stabilization energy. Therefore, both ways of a description of the crystal packing (those based on geometrical characteristics 4446

dx.doi.org/10.1021/cg500533f | Cryst. Growth Des. 2014, 14, 4439−4449

Crystal Growth & Design

Article

intermolecular interactions (n···π and π···π) contribute to the stabilization energy of 1·MeCN crystal; however, contribution of the n···π type is higher. Sum of the energies of the M0···macrocycle and M0···solvent interactions is equal to −25.9 and −16.0 kcal/mol. Similar to 1· 0.5DCE, stabilization of the M0 by its interactions with the macrocycle molecules decreases (by 9.5 kcal/mol); however, total stabilization energy is increased by 11.2 kcal/mol that is much higher than that obtained for 1·0.5DCE.

Table 3. Close Contacts (Å) and Intermolecular Interaction Energies (kcal/mol) as Obtained from Quantum Chemical (DFT) Calculations for 1·MeCNa N A

D

E F

G

I

J L P

Q

S0 S1

S2 S3 S4 S5

close contact

distance

O3···C3 3.203 N5···C3 3.282 N5···C4 3.314 C3···O3 3.203 C3···N5 3.282 C4···N5 3.314 N7···N5 3.125 N6···N7 3.135 N6···N14 3.173 N6···C7 3.319 C1···N1 3.188 C2···N1 3.092 N3···C1 3.092 N4···C5 3.206 N4···C6 3.190 C1···N3 3.092 C5···N4 3.206 C6···N4 3.190 N5···N7 3.125 N7···N6 3.135 N14···N6 3.173 C7···N6 3.319 N1···C1 3.188 N1···C2 3.092 no close contacts N2···O4 3.042 C2···O4 3.283 N2···N8 3.197 O4···N2 3.042 O4···C2 3.283 N8···N2 3.197 N8···H2Sb 2.786 N8···C2S N8···H2Sc 2.546 N8···C1S 3.167 C7···N1S 3.233 C8···N1S 3.256 C8···C1S 3.264 N9···H2Sa 2.785 no close contacts O1···C1S 3.154 no close contacts

energy (DFT) for 1· MeCN

energy (DFT) for 1

energy (DFT) for 1·0.5DCE

−4.5

−6.2

−6.0

−3.0

−3.1

−3.1

−2.7

−3.1

−3.1

−3.5

−2.8

−2.3

−3.5

−2.8

−2.3

−3.0

−1.6

−1.3

−2.7

−1.6

−1.3

−2.0 −0.5

−0.7 −0.1

−0.6 −0.1

−0.5

−0.3

−0.2

−5.8

−2.9

−4.9

−3.6

−1.7 −1.4 −1.2 −1.0

−0.9

4. CONCLUSIONS In the present work, we have tried to find some factors that govern crystal packing formation of the furazan-containing macrocycles based on the X-ray diffraction data, quantum chemical calculations as well as on the analysis of the crystal packing of the macrocycle 1, its solvates and previously investigated furazan macrocycles. Our results showed that n···π type of weak intermolecular interactions prevails over π···π type due probably to relatively strong acceptor properties of the furazan ring. In spite of the absence of the strong intermolecular interactions, macrocycle 1 has a relatively high density (1.800 g/cm3 at 100 K) due to numerous weak interactions that is in accord with the well-known auf bau principle47 which is based on dense packing and shape similarity principles. This successfully explains why in the crystal structure of the solvates of macrocycle a solvent occupies voids inside columnar structure causing relatively strong intermolecular interactions to break. At the same time, the majority of other intermolecular close contacts in crystallosolvate of 1 with DCE are retained, and interaction energies are nearly unchanged. In the 1·MeCN solvate, crystal packing is modified to a greater extent. While, at the first glance, crystal packing of solvates 1·MeCN and 1· 0.5DCE looks similar, the system of close contacts is completely different. However, such a redistribution does not cause significant changes in the stabilization energy, thus outlining an important role of weak intermolecular interactions. The old interactions can easily break, but the new ones can easily replace them. In the other words, crystal packing that is stabilized by weak interactions is easily adjustable to an influence of the surroundings. This fact illustrates existing problems related to a prediction of cocrystal formation15 and a more general problem of the crystal structure prediction.54,55 Both approaches (geometrical and energetical) explain the observed crystal packing of crystallosolvates, and the shape similarity principle can be utilized for further design of cocrystals of the macrocycle 1. However, from a consideration of the structure of the individual macrocyle only, one can hardly predict how a solvate or a cocrystal would be formed.

a

For a convenient comparison, intermolecular interaction energies for 1 and 1·0.5DCE structures are also given in the last two columns.



layers parallel to the ab crystallographic plane (due to different space groups, corresponding layers (L4) in 1 and 1·0.5DCE are parallel to the bc plane). The layer energy is equal to −9.2 kcal/ mol that is 1.7 kcal/mol higher than that found in the structure 1. This is related to different relative orientation of the molecules inside the layer. In 1 as well as in 1·0.5DCE, columns are related by the symmetry plane, while in 1·MeCN by the 2fold axis. The energy of the L4, which is also retained upon solvate formation, increases to a lesser extent (by 0.5 kcal/mol). However, the system of close contacts changes significantly, again, due to different relative orientations of the columns. Similar to 1 and 1·0.5DCE crystals, both types of weak

ASSOCIATED CONTENT

S Supporting Information *

Figures of residual and deformational experimental electron density, examples of dimeric interaction pattern, crystal packing diagrams, atomic charges and volumes, and energies of all intermolecular interactions. X-ray crystallographic information files (CIF) are available for compounds 1, 1·0.5DCE, 1·MeCN. This material is available free of charge via the Internet at http://pubs.acs.org. Crystallographic information files are also available from the Cambridge Crystallographic Data Center (CCDC) upon request (http://www.ccdc.cam.ac.uk, deposition numbers 997122−997124). 4447

dx.doi.org/10.1021/cg500533f | Cryst. Growth Des. 2014, 14, 4439−4449

Crystal Growth & Design



Article

I. L.; Kuzmina, N. E. Mendeleev Commun. 2001, 112. (i) Sheremetev, A. B.; Shatunova, E. V.; Averkiev, B. B.; Dmitriev, D. E.; Petukhov, V. A.; Antipin, M. Yu. Heteroatom Chem. 2004, 15, 131. (j) Averkiev, B. B.; Timofeeva, T. V.; Sheremetev, A. B.; Shatunova, E. V.; Antipin, M. Yu. Acta Crystallogr. 2004, C60, 520. (k) Sheremetev, A. B.; Aleksandrova, N. S.; Dmitriev, D. E.; Averkiev, B. B.; Antipin, M. Yu. J. Heterocycl. Chem. 2005, 42, 519. (l) Sheremetev, A. B.; Ivanova, E. A.; Dmitriev, D. E.; Kulagina, V. O.; Averkiev, B. B.; Antipin, M. Yu. J. Heterocycl. Chem. 2005, 42, 803. (m) Epishina, M. A.; Kulikov, A. S.; Makhova, N. N. Russ. Chem. Bull. 2008, 57, 644; Translation Izv. Akad. Nauk, Ser. Khim. 2008, 631]. (n) Chavez, D. E.; Parrish, D. A.; Leonard, P. Synlett 2012, 23, 2126. (7) (a) Pivina, T. S. International Annual Conference of ICT, 1993, 24th (Energetic Materials: Insensitivity and Environmental Awareness), 91-1/91-16. (b) Pivina, T. S.; Sukhachev, D. V.; Evtushenko, A. V.; Khmelnitskii, L. I. Propellants, Explos., Pyrotechnics 1995, 20, 5. (c) Sheremetev, A. B.; Kulagina, V. O.; Batog, L. V.; Lebedev, O. V.; Yudin, I. L.; Pivina, T. S. Proceedings of the International Pyrotechnics Seminar, 1996, 22nd; 377. (d) Batog, L. V.; Konstantinova, L. S.; Rozhkov, V. Yu.; Lebedev, O. V.; Epishina, M. A.; Makhova, N. N.; Ovchinnikov, I. V.; Khmel’nitskii, L. I. International Annual Conference of ICT (1998), 29th (Energetic Materials), 55-1/55-10. (e) Liu, J.; Tian, D.; Hong, W.; Huang, Q. Huaxue Wuli Xuebao 2002, 15, 351. (f) Zhang, C.; Shu, Y.; Zhao, X.; Dong, H.; Wang, X. J. Mol. Struct.: THEOCHEM 2005, 728, 129. (g) Li, Z.-X.; Tang, S.-Q.; Wang, W.-J. Hanneng Caillao (Chin. J. Energy Mater.) 2007, 15, 6. (h) Li, X.; Zhang, R.; Zhang, X. Struct. Chem. 2013, 24, 1193. (8) Kots, A. Ya.; Batog, L. V.; Betin, V. L.; Rozhkov, V. Yu.; Ukraintsev, K. E.; Khropov, Yu. V.; Epishina, M. A.; Sheremetev, A. B.; Makhova, N. N.; Bulargina, T. V. Patent RU 2151799, 1999. (9) (a) Eman, V. E.; Sukhanov, M. S.; Lebedev, O. V.; Batog, L. V.; Konstantinova, L. S.; Rozhkov, V. Yu.; Dekaprilevich, M. O.; Struchkov, Yu. T.; Kmel’nitskii, L. I. Mendeleev Commun. 1997, 5. (b) Li, W.-W.; Li, Z.-X.; Wang, W.-J. Hanneng Caillao (Chin. J. Energy Mater.) 2009, 17, 11. (c) Veauthier, J. M.; Chavez, D. E.; Tappan, B. C.; Parrish, D. A. J. Energy Mater. 2010, 28, 229. (10) Sheremetev, A. B.; Kharitonova, O. V.; Mantseva, E. V.; Kulagina, V. O.; Shatunova, E. V.; Aleksandrova, N. S.; Melnikova, T. M.; Ivanova, E. A.; Dmitriev, D. E.; Eman, V. A.; Yudin, I. L.; Kuzmin, V. S.; Strelenko, Yu. A.; Novikova, T. S.; Lebedev, O. V.; Khmelnitskii, L. I. Russ. J. Org. Chem. 1999, 35, 1525. (11) Sheremetev, A. B.; Betin, V. L.; Yudin, I. L.; Kulagina, V. O.; Khropov, Yu. V.; Bulargina, T. V.; Kots, A. Ya.; Aherementev, A. V. Patent RU 2167161, 2001. (12) Braga, D.; d’Agostino, S.; Grepioni, F. Cryst. Growth Des. 2012, 12, 4880. (13) (a) Vishweshwar, P.; McMahon, J. A.; Bis, J. A.; Zaworotko, M. J. J. Pharm. Sci. 2006, 95, 499. (b) Shan, N.; Zaworotko, M. J. Drug Discovery Today 2008, 13, 440. (14) (a) Almarsson, O.; Zaworotko, M. J. Chem. Commun. 2004, 1889. (b) Bolton, O.; Matzger, A. J. Angew. Chem., Int. Ed. 2011, 50, 8960. (c) Childs, S. L.; Zaworotko, M. J. Cryst. Growth Des. 2009, 9, 4208. (15) Aakeröy, C. B.; Desper, J.; Fasulo, M.; Hussain, I.; Levin, B.; Schultheiss, N. CrystEngComm 2008, 10, 1816. (16) (a) Aakeröy, C. B.; Chopade, P. D. In Supramolecular Chemistry: From Molecules to Nanomaterials; Gale, Ph.; Steed, J., Eds.; John Wiley & Sons: Chicester, U.K., 2012; 2975. (b) Aakeröy, C. B.; Champness, N. R.; Janiak, C. CrystEngComm 2010, 12, 22. (c) Shattock, T. R.; Arora, K. K.; Vishweshwar, P.; Zaworotko, M. J. Crystal Growth Des. 2008, 8, 4533. (17) (a) Selig, W. Explosivstoffe 1967, 15, 76. (b) Landenberger, K. B.; Matzger, A. J. Cryst. Growth Des. 2012, 12, 3603. (c) Lin, H.; Zhu, S.-G.; Zhang, L.; Peng, X.-H.; Li, H.-Z. J. Energy Mater. 2012, 31, 261. (18) Bolton, O.; Simke, L. R.; Pagoria, P. F.; Matzger, A. J. Cryst. Growth Des. 2012, 12, 4311. (19) APEX2 and SAINT; Bruker AXS Inc.: Madison, Wisconsin, USA, 2005. (20) Sheldrick, G. M. Acta Crystallogr. 2008, A64, 112.

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]; fax: +7 499 135 5085; phone: +7 499 135 9214. Notes

The authors declare no competing financial interest.

■ ■

ACKNOWLEDGMENTS The authors gratefully acknowledge support of the Russian Foundation for Basic Research (Project No. 13-03-12197). DEDICATION The authors dedicate this work to the memory of our teacher and colleague Prof. Mikhail Yu. Antipin. K.Yu.S., K.A.L., and I.V.A. worked in the lab of Professor Antipin for many years, and it was certainly his devotion that stimulated our interest in structural chemistry.



REFERENCES

(1) (a) Host Guest Complex Chemistry: Macrocycles. Synthesis, Structures, Applications; Vogtle, F.; Weber, E., Eds.; Springer-Verlag: Berlin Heidelberg, 1985. (b) The Chemistry of Heterocyclic Compounds: Aza-crown macrocycles; Bradshaw, J. S.; Krakowiak, K. E.; Izatt, R. M., Eds.; John Wiley & Sons, Inc: Hoboken, NJ, USA, 1993; Vol. 51. (c) Kulikov, O. V.; Pavlovsky, V. I.; Andronati, S. A. Chem. Heterocycl. Compd. 2005, 41, 1447. (d) Modern Supramolecular Chemistry: Strategies for Macrocycle Synthesis; Diederich, F.; Stang, P. J.; Tykwinski, R. R., Eds.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, 2008. (e) Ushakov, E. N.; Alfimov, M. V.; Gromov, S. P. Russ. Chem. Rev. 2008, 77, 39. (f) Rusalov, M. V.; Uzhinov, B. M.; Alfimov, M. V.; Gromov, S. P. Russ. Chem. Rev. 2010, 79, 1099. (g) Buzykin, B. I. Chem. Heterocycl. Compd. 2010, 46, 1043. (h) Davis, F.; Higson, S. Macrocycles: Construction, Chemistry and Nanotechnology Applications; John Wiley & Sons, Ltd.: Chichester, UK, 2011. (i) Knyazeva, R.; Burilov, A. R.; Pudovik, M. A.; Habicher, W. D. Russ. Chem. Rev. 2013, 82, 150. (j) Borovika, D.; Bertrand, P.; Trapencieris, P. Chem. Heterocycl. Compd. 2013, 49, 1560. (2) Reuter, R.; Wegner, H. A. Chem. Commun. 2011, 47, 12267. (3) (a) Balzani, V.; Venturi, M.; Credi, A. Molecular Devices and Machines; Wiley-VCH: Weinheim, 2003. (b) Hoeben, F. J. M.; Jonkheijm, P.; Meijer, E. W.; Schenning, A. P. H. J. Chem. Rev. 2005, 105, 1491. (c) Caruso, M. M.; Davis, D. A.; Shen, Q.; Odom, S. A.; Sottos, N. R.; White, S. R.; Moore, J. S. Chem. Rev. 2009, 109, 5755. (4) (a) Dziomko, V. M. Khim. Geterotsikl. Soedin. 1982, 3; Chem. Heterocycl. Compd. 1982, 18, 1 (Engl. Transl.)]. (b) Dziomko, V. M.; Ruchkina, G. N.; Ryabokobylko, Y. S.; Orlova, E. E.; Poponova, R. V. Khim. Geterotsikl. Soedin. 1991, 360; Chem. Heterocycl. Compd. 1991, 27, 292 (Engl. Transl.)]. (5) (a) Elwahy, A. H. M. J. Heterocycl. Chem. 2003, 40, 1. (b) Borisova, N. E.; Reshetova, M. D.; Ustynyuk, Y. A. Russ. Chem. Rev. 2007, 76, 785; Usp. Khim. 2007, 76, 843 (Engl. Transl.)]. (6) (a) Sheremetev, A. B.; Kharitonova, O. V. Mendeleev Commun. 1992, 157. (b) Chertanova, L.; Pascard, C.; Sheremetev, A. Supramol. Chem. 1993, 3, 71. (c) Eman, V. E.; Sukhanov, M. S.; Lebedev, O. V.; Batog, L. V.; Konstantinova, L. S.; Rozhkov, V. Yu.; Kmel’nitskii, L. I. Mendeleev Commun. 1996, 66. (d) Batog, L. V.; Konstantinova, L. S.; Lebedev, O. V.; Kmel’nitskii, L. I. Mendeleev Commun. 1996, 193. (e) Sheremetev, A. B.; Kulagina, V. O.; Ivanova, E. A. J. Org. Chem. 1996, 61, 1510. (f) Batog, L. V.; Konstantinova, L. S.; Eman, V. E.; Sukhanov, M. S.; Batsanov, A. S.; Struchkov, Y. T.; Lebedev, O. V.; Kmel’nitskii, L. I. Khim. Geterotsikl. Soedin. 1996, 406; Chem. Heterocycl. Compd. 1996, 32, 352 (Engl. Transl.)]. (g) Batog, L. V.; Rozhkov, V. Y.; Konstantinova, L. S.; Eman, V. E.; Dekaprilevich, M. O.; Struchkov, Y. T.; Semenov, S. E.; Lebedev, O. V.; Kmel’nitskii, L. I. Izv. Akad. Nauk, Ser. Khim. 1996, 1250; Russ. Chem. Bull. 1996, 45, 1189 (Engl. Transl.)]. (h) Sheremetev, A. B.; Kulagina, V. O.; Yudin, 4448

dx.doi.org/10.1021/cg500533f | Cryst. Growth Des. 2014, 14, 4439−4449

Crystal Growth & Design

Article

(21) Hansen, N. K.; Coppens, P. Acta Crystallogr. 1978, A34, 909. (22) Volkov, A.; Macchi, P.; Farrugia, L. J.; Gatti, C.; Mallinson, P.; Richter, T.; Koritsanszky, T. XD2006 - A Computer Program Package for Multipole Refinement, Topological Analysis of Charge Densities and Evaluation of Intermolecular Energies from Experimental and Theoretical Structure Factors, 2006. (23) Hirshfeld, F. L. Acta Crystallogr. 1976, A32, 239. (24) Kirzhnits, D. A. Sov. Phys. JETP 1957, 5, 54. (25) Tsirelson, V. G. Acta Crystallogr. 2002, 58, 632. (26) Lyssenko, K. A. Mendeleev Commun. 2012, 22, 1. (27) Espinosa, E.; Alkorta, I.; Rozas, I.; Elguero, J.; Molins, E. Chem. Phys. Lett. 2001, 336, 457. (28) Stash, A.; Tsirelson, V. J. Appl. Crystallogr. 2002, 35, 371. (29) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B. et al. Gaussian 03, Revision E.01; Gaussian, Inc.: Wallingford, CT, 2004. (30) Zhao, Y.; Schultz, N. E.; Truhlar, D. G. J. Chem. Theory Comput. 2006, 2, 364. (31) Zhao, Y.; Truhlar, D. G. J. Chem. Theory Comput. 2007, 3, 289. (32) Leverentz, H. R.; Truhlar, D. G. J. Phys. Chem. A 2008, 112, 6009. (33) Sponer, J.; Riley, K. E.; Hobza, P. Phys. Chem. Chem. Phys. 2008, 10, 2595. (34) Suponitsky, K. Y.; Masunov, A. E.; Antipin, M. Y. Mendeleev Commun. 2009, 311. (35) Suponitsky, K. Y.; Lyssenko, K. A.; Antipin, M. Y.; Aleksandrova, N. S.; Sheremetev, A. B.; Novikova, T. S. Russ. Chem. Bull. 2009, 58, 2129. (36) Sheremetev, A. B.; Yudin, I. L.; Palysaeva, N. V.; Suponitsky, K. Y. J. Heterocycl. Chem. 2012, 49, 394. (37) Suponitsky, K. Y.; Masunov, A. E. J. Chem. Phys. 2013, 139, 094310. (38) Allen, F. H. Acta Crystallogr. 2002, B58, 380 Cambridge Crystallographic Database, version 5.34, 2012. (39) (a) Panunto, T. W.; Urbanczyk-Lipkowska, Z.; Johnson, R.; Etter, M. C. J. Am. Chem. Soc. 1987, 109, 7786. (b) Etter, M. C. Acc. Chem. Res. 1990, 23, 120. (c) Etter, M. C. J. Phys. Chem. 1991, 95, 4601. (40) (a) Turi, L.; Dannenberg, J. J. J. Phys. Chem. 1992, 96, 5819. (b) Turi, L.; Dannenberg, J. J. Chem. Mater. 1994, 6, 1313. (c) Turi, L.; Dannenberg, J. J. J. Am. Chem. Soc. 1994, 116, 8714. (d) Turi, L.; Dannenberg, J. J. J. Phys. Chem. 1996, 100, 9638. (e) Masunov, A.; Dannenberg, J. J. J. Phys. Chem. B 2000, 104, 806. (41) Suponitsky, K. Yu.; Timofeeva, T. V.; Antipin, M. Yu. Usp. Khim. 2006, 75, 515. (42) Sheremetev, A. B.; Palysaeva, N. V.; Struchkova, M. I.; Suponitsky, K. Yu. Mendeleev. Commun. 2012, 22, 302. (43) Sheremetev, A. B.; Yudin, I. L.; Suponitsky, K. Yu. Mendeleev Commun. 2006, 16, 264. (44) Gobel, M.; Klapotke, T. M. Adv. Funct. Mater. 2009, 19, 347. (45) Sheremetev, A. B.; Aleksandrova, N. S.; Palysaeva, N. V.; Struchkova, M. I.; Tartakovsky, V. A.; Suponitsky, K. Yu. Chem.Eur. J. 2013, 19, 12446. (46) Palysaeva, N. V.; Kumpan, K. P.; Struchkova, M. I.; Dalinger, I. L.; Kormanov, A. V.; Aleksandrova, N. S.; Chernyshev, V. M.; Pyreu, D. F.; Suponitsky, K. Yu.; Sheremetev, A. B. Org. Lett. 2014, 16, 406. (47) (a) Kitaigorodsky, A. I. Organic Chemical Crystallography; Consultants Bureau: New York, 1961. (b) Kitaigorodsky, A. I. Molecular Crystals and Molecules; Academic Press: Orlando, FL, 1973. (48) Nelyubina, Yu. V.; Glukhov, I. V.; Antipin, M. Yu.; Lyssenko, K. A. Chem. Commun. 2010, 46, 3469. (49) (a) Zorkii, P. M.; Bel’skii, V. K.; Lazareva, S. G.; Porai-Koshits, M. A. J. Struct. Chem. 1967, 8, 267. (b) Zorky, P. M.; Zorkaya, O. N. Adv. Mol. Struct. Res. 1997, 3, 147. (c) Grineva, O. V.; Zorkii, P. M. Russ. J. Phys. Chem. 1998, 72, 714. (50) Espinosa, E.; Molins, E.; Lecomte, C. Chem. Phys. Lett. 1998, 285, 170. (51) Mata, I.; Alkorta, I.; Espinosa, E.; Molins, E. Chem. Phys. Lett. 2011, 507, 185.

(52) Shishkin, O. V.; Dyakonenko, V. V.; Maleev, A. V. CrystEngComm 2012, 14, 1795. (53) Rowland, R. S.; Taylor, R. J. Phys. Chem. 1996, 100, 7384. (54) Gavezzotti, A. Acc. Chem. Res. 1994, 27, 309. (55) Dunitz, J. D. Chem. Commun. 2003, 545.

4449

dx.doi.org/10.1021/cg500533f | Cryst. Growth Des. 2014, 14, 4439−4449