Raman spectroscopy study of new thia - South Ural State University

0 downloads 0 Views 2MB Size Report
Feb 27, 2015 - crystalline state.1 Diiodine molecules in the I2 crystal are not isolated and form a series of intermolecular interactions with different nature.27 In ...
NJC PAPER

Cite this: New J. Chem., 2015, 39, 6163

Raman spectroscopy study of new thia- and oxazinoquinolinium triodides Irina D. Yushina,*a Boris A. Kolesovb and Ekaterina V. Bartashevicha A number of new solid polyiodides of thia- and oxa-zinoquinolinium derivatives have been characterized using a combined approach of Raman spectroscopy and periodic 3D calculations of the Raman intensities. Various cation–anion ratios, including tri- and polyiodides of the complex structure [I3  I2], have been

Received (in Montpellier, France) 27th February 2015, Accepted 25th May 2015

studied. For the oriented single crystal containing the [I3  I2] fragments, the polarized Raman spectra and

DOI: 10.1039/c5nj00497g

have allowed us to get new information on the spatial organization of the intermolecular interactions of

the angular dependency of the band intensities on crystal rotation have been obtained. These techniques iodine in polyiodide-anion chains. They have also shown perspectives on the geometrical characterization

www.rsc.org/njc

of complex polyiodides using Raman spectroscopy.

Introduction Organic and inorganic polyiodides with anion stoichiometry Inx (the n value varies in the range 1 o n o 29)1 exhibit a wide spectrum of practically important properties. They are known as low-temperature conductors,2,3 ionic liquids,4,5 water disinfectants,6 radioactive markers7 and anti-thyroid drugs.8 Such physicochemical and biological properties warm up the investigations in the field of modern methods of synthesis,9 crystallization10 and structural characterization of polyiodides.11 A number of polyiodides with various organic cations are presented in the crystallographic database CSD,12 thus it is possible to analyze the crystalline packing features and estimate the various stoichiometry ratios of iodine atoms in polyiodide anions.13,14 The structural diversity of polyiodides remains a subject of constant interest.1,15 Only few discreet subunits (such as I, bound molecular iodine I2, triiodide anion I3) can form a wide variety of 2D and 3D spatial organizations in polyiodides. It becomes possible due to the ability of iodine to form different types of non-covalent interactions inside one anion chain, with neighboring anions and atoms of a heterocyclic cation via I  I, I  S, I  O, I  N and I  H interactions. Among such interactions, two of the most frequent cases are the directed and electrostatically driven16,17 halogen bonds that co-exist with undirected and weak van der Waals interactions.18 Halogen bonding according to IUPAC recommendation is defined as an interaction, where ‘‘there is an evidence of a net attractive

a

Chemistry Department, South Ural State University (National Research University), Lenin av., 76, 454080, Chelyabinsk, Russia. E-mail: [email protected], [email protected] b Institute of Inorganic Chemistry SB RAS, Lavrentiev av., 3, 630090, Novosibirsk, Russia. E-mail: [email protected]

interaction between an electrophilic region associated with a halogen atom in a molecular entity and a nucleophilic region in another, or the same, molecular entity’’.19 The reason of halogen bond formation is the anisotropy of electrostatic potential distribution on the van der Waals surface of a molecule containing a covalently bonded halogen atom. It is known17 that there are electron-deficient regions, located on the extension of the covalent I–I bond in molecular I2 and the electron concentration region in the equatorial part of iodine atom. Halogen bond formation is only possible if iodine and triiodide species are orientated inter-consistently to each other. Such features result in the formation of complex polyiodide chains, ribbons, channels or layers within the crystal. It is known20 that many practically important properties depend on the different types of interactions with iodine participation. Thus, the study of intermolecular interactions can give a clue not only to predict the physicochemical properties2,21 but also can give an insight into structural engineering and crystal design.22 Spectral methods are widely used in the structural characterization of polyiodide spatial organization and intermolecular interaction features.23,24 Raman spectroscopy is well-known as an effective tool for compounds with I–I bonds.25 The limitations of IR spectroscopy originate from the low-frequency location of the I–I vibration bands so that they are difficult to observe.1 Furthermore, symmetry selection rules make some of the vibrations of iodine bonds inactive. Raman spectroscopy gives an opportunity to identify the typical species in polyiodides such as triiodide, pentaiodide and bound molecular iodine.25 Moreover, this method has several advantages, such as an ability to obtain spectra using a small amount of a substance without its destruction, to acquire spectra in all states of matter, to visually control changes in the external properties and surface features.26

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015

New J. Chem., 2015, 39, 6163--6170 | 6163

Paper

NJC

The latter is very important for samples with an inhomogeneous composition or for non-stable substances in order to control the laser-induced degradation processes from taking place.1,25 A typical Raman spectrum of polyiodide can be represented as a number of bands in several spectral ranges. The band for the I–I stretching vibration in molecular iodine is located near 213 cm1 for the gaseous state, 194 cm1 for the melt and 180 cm1 for the crystalline state.1 Diiodine molecules in the I2 crystal are not isolated and form a series of intermolecular interactions with different nature.27 In the case of the solid-state iodine-containing molecular complexes, this band shifts to lower wavenumber of 140–180 cm1 because of iodine halogen bond formation1,28 or other non-covalent interactions. Location of this band is thoroughly investigated for a wide variety of cases, especially for complexes with N  I2 and S  I2 or Se  I2 interactions.29–32 It is known that shortening of the I  X distances (where X = S, N or another electron-donor atom) leads to the corresponding elongation of the I–I bond.25 According to such correlations in bond lengths, the molecular iodine complexes are divided into two groups: weak adducts, where d(I–I) is less than 2.86 Å and strong ones, where d(I–I) is more than 2.86 Å. Such division is proved by the different spectral behavior of these two types of adducts. For weak and medium-weak adducts, the linear dependency between the elongation of the I–I bond and a decrease in the corresponding wavenumber for the I–I band was observed.28 The changes in the spectral properties of iodine and its complexes at different temperatures and under pressure have also been in the focus of attention.33,34 The range of wavenumber for solid iodine is limited by 180 cm1 on the one side and 140–150 cm1 on the other side for 2.85–2.86 Å I–I bond lengths.28 For strong adducts, there is no such strict dependency, as in this case, specific interactions take place and they can crucially change the spectral properties. In some cases, even the ionization of the S  I bond takes place and instead of S  I2  I2 complex a triiodide-like structure S–I+  I3 is formed.25 In this limiting case, the band for coordinated I2 disappears and the bands for nsym(S–I–I) and nasym(S–I–I) are observed instead. The spectral behavior in this case is also similar to typical triiodide-anion [I–I–I] as all three iodine atoms oscillate consistently.25 The next frequently observed polyiodide subunit is triiodide I3. A symmetric triiodide anion has one Raman active mode: the symmetrical stretching vibration n1 and two IR active bands: bending n2 and asymmetric stretching vibration n3.1 If two I–I bonds in triiodide have different lengths, the symmetry decreases and all bands become IR and Raman active. The spectral ranges for these bands are as follows: 50–70 cm1 for n2 or dI–I, 100–120 cm1 for n1 or nsym, 130–140 cm1 for n3 or nasym.1 The asymmetry of the I–I bonds in triiodide can be the result of non-equivalent interactions of the terminal iodine atoms in the anion.35 The influence of I  H interactions between the anion and cation, I  I and other non-covalent contacts on the bond length and polarization leads to the emergence of nasym asymmetric vibration band. It has also been mentioned that the existence of polyvalent cations may increase the asymmetry of the triiodide anion.36,37 The asymmetry of the bonds can reach a

6164 | New J. Chem., 2015, 39, 6163--6170

limit after which only the band corresponding to molecular iodine with an elongated bond is observed, but not the bands for I3, though the stoichiometry of this anion still includes three iodine atoms.38 Such types of triiodide can be presented as a I  I–I unit. Spectral examples for the latter structure are known for CsI3 and 3,5-bis(ethylamino)-1,2-dithiolylium I3. Similar organization is observed for I42 (I  I–I  I) anions.39,40 Triiodides can be isolated and do not form strong interactions with neighboring anions or the atoms of cations. However, sometimes they can form more complicated spatial organizations via I3  I3 or I3  I2 motifs. Asymmetric triiodides are inclined to such interactions more frequently as the I(1)  I(2)–I(3) motif has an ability to form directed halogen bonds or van der Waals side interactions between the I(3) atom of one anion and the I(1) 0 atom of the neighboring anion. In the simplest case, such organization results in the formation of a I3  I3 (I62) structure.38,41 Such a motif has the typical bands of asymmetric triiodide: an intense band near 165 cm1 and a weak band around 100 cm1.1,38 The abovementioned experimental facts reveal the problem of band assignment and anion composition characterization. Though there is no shortage of experimental data concerning the observed bands and corresponding bond lengths in anions, there is still a problem as in many studies, only bonds inside the anions are taken into account. However, various anion–anion and anion–cation interactions can crucially influence the experimental spectra and lead to difficulties in interpreting the anion stoichiometry on the basis of spectral data only. Spectroscopic assignment of the bands for samples without X-ray diffraction data is usually not a simple task. In the case of isolated triiodides, it is possible to come to unambiguous decision, but in more complex structures uncertainty is still present. On one hand, it is always necessary to compare the spectral properties with X-ray diffraction data for correct band assignment. On the other hand, the analysis of bond lengths from X-ray diffraction experiments alone is not sufficient to characterize the type of bonding and anion asymmetry. A combination of these two approaches can give a clue to the multi-faceted investigation of intermolecular interactions in crystals and their corresponding spectral properties. Polarized Raman spectra of oriented single crystals are widely used towards the determination of the principle directions of linear or chain-like structures inside the crystal bulk.42 In the case of polyiodides, this technique can be useful in determining the direction of triiodide anions and their mutual orientations. Moreover, if we deal with polyiodides of I3  I2 type, it will help us to locate both iodine subunits according to the maximum intensity of the corresponding band: nsym for triiodide and nI–I for bound I2.43 It is known that high-pressure deformations differ for the directions collinear to chain-like anions and perpendicular to them.44 Moreover, recent discoveries have discussed the elastic component of deformations in the direction of a halogen bond.45 Therefore, it is very important to locate the specific direction of polyiodide subunits in a crystal. High-pressure studies are always related to the morphology of the crystal in

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015

NJC

each particular case, so the polarized Raman spectroscopy technique is an effective preliminary stage to establish the most suitable crystal orientation under an incident laser beam. Furthermore, the changes in the spectral properties of the highly intense and characteristic polyiodide bands can be a reliable criterion for the changes in the intermolecular interactions and can prove the ongoing phase transitions under high-pressure conditions. Therefore, the study of polyiodide subunit orientation with respect to the morphological crystal axis, used in this work, is more informative for further high-pressure polyiodide studies than the same analysis of orientations in terms of the crystallographic axis. A morphological approach deals with the real crystal and takes into account its irregular shape, growth features and even the visual defects on a crystals surface. Calculations of the geometric and vibrational characteristics of triiodide anions were previously demonstrated mainly as their isolated form in a gaseous state,46–48 solution41,49 or in a model crystalline state with analytical correction in order to take into account the neighboring anions in the solid state. The fully periodic calculation of triiodide combined with an organic cation is a clue towards a deeper understanding of the noncovalent interactions in the solid state.50,51 The spectroscopic behavior of triiodide was investigated as a part of calculations for a wide series of possible polyiodide spatial organizations in several theoretical approaches.52 The main disadvantage of such techniques is the inability to take into consideration the asymmetry of triiodide as a result of the non-equivalent noncovalent interactions of the two terminal iodine atoms in the anion.35,53 The various thia- and oxazinoquinolinium cations are of practical interest as promising bacteriostatic agents, disinfectants, iodine carriers, anti-inflammatory supplements and anti-thyroid drugs.7,8 Furthermore, they are in close touch with the fundamental questions of intermolecular interactions in solids and the problem of obtaining a number of crystalline polyiodides with one organic cation and anions of varying stoichiometric composition.13 A combination of experimental polarized Raman data and theoretically calculated wavenumber and intensities in the crystalline approximation can give a clue towards a deeper understanding of the polyiodide units location in the crystal bulk and I–I/I  I intermolecular interactions. Therefore, the aim of the present study is to analyze the features of the crystal structures for different types of triiodides on the basis of a combined approach, including experimental crystal structure analysis, ab initio quantum chemical calculations, experimental spectral properties and polarized Raman spectra of oriented crystals.

Experimental Spectroscopic experiments Raman spectra were obtained using a Triplemate, SPEX spectrometer with a CCD detector, LN-1340PB, from Princeton Instruments. The 632.3 nm line of a He–Ne laser was used for spectral excitation. The samples were placed in the focal plane of a

Paper

LD-EPIPLAN, 40/0.60 Pol., Zeiss objective with 2 mm working distance and aperture 0.6, which was used for laser beam focusing and collection of the scattered light. The diameter of the laser beam on the samples surface was 2 mm. The spectra were obtained in a 1801 backscattering collection geometry with a Raman microscope. All measurements were performed with a spectral resolution of 2.0 cm1. Raman spectra with the 488 nm excitation line of an Ar+ laser (35LAP431 from ‘‘Melles Griot’’ company, USA) were acquired using a LabRAM Horiba spectrometer with a CCD detector Symphony from Jobin Yvon. The laser line was used for wavenumber correction of the spectra. The laser output power was about 20 mW and the power on the sample surface was about 1 mW or less. The stability of the laser power was better than 0.5%. The incident light was focused and the scattered light was collected using a ‘‘Olympus’’ objective MPlan N with 10-fold enhancement and numerical aperture A = 0.25. The polarization leakage L was no more than 0.03%. All measurements were performed with a spectral resolution of 0.7 cm1. Raman spectra for single crystals were analyzed with 633 nm wavelength of incident light with various relative directions in the polarization vector of the incident and scattered light and the morphological crystal directions. The directions of the polarization vectors of incident and scattered light were fixed, while single crystals were rotated in specified planes. The polarized spectra were used not only to distinguish basic polyiodide species but also to obtain information on the local interactions with their neighbors and their orientation in the unit cell. For the latter, the angular dependency of the Raman intensity was measured. The Raman tensor of the n1 mode had only one non-zero diagonal ll-component, wherein l coincided with the molecular I2, I3 axis. When the angle y between the molecular axis and the direction of the electric field of the incident and scattered light changed, the ll-component of the Raman tensor changed too as cos2 y. The Raman intensity was proportional to the square of the Raman tensor component. Thus, the Raman intensity of the n1 changed as cos4 y and the angular dependency of n1 intensity provided the true direction of the linear I2 and I3 molecules in the crystal lattice to be measured. For triiodide 1 and complex polyiodide 4, single crystal rotations with a step of 20 degrees were performed. The obtained results were presented as the angular dependency of peak intensity for the nsym I3 vibration band on the degree of crystal rotation. In the starting point of rotation, the directions of the polarization vectors of incident and scattered light were collinear with the longest direction of the needle-like crystal 1. For crystal 4, the polarized spectra were presented as the percentage, corresponding to the nsym and nI–I peak area in total intensity for each crystal orientation. As polyiodides, especially with a high iodine ratio, are rather unstable compounds and can decompose under heating,54 it is very important to thoroughly control the surface of a sample after the spectrum registration. The thermal stability of polyiodides usually decreases as the iodine content increases.1,15 The possibility of undergoing degradation processes under laser exposure not only complicates the experimental procedure,

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015

New J. Chem., 2015, 39, 6163--6170 | 6165

Paper

NJC

but also influences the interpretation of obtained results, as there are several examples in the literature, when a sample was characterized incorrectly because the obtained spectrum referred to decomposition products (gaseous iodine, triiodide) but not to the initial complex polyiodide.25,55 This problem leads to the necessity for controlling the laser power, changing the time of exposure and observing the samples surface after laser exposure. In the present study, color filters with 2-, 10-, 20-, 100-, 200-fold weakening have been used. Calculation details Periodic 3D calculations for crystalline triiodide C12H11INS+I3 1 were performed using the Kohn–Sham method with the B3LYP functional.56,57 The CRYSTAL1458 program package was used. The modified DZVP basis set for iodine atoms59 and basis set for C, H, N, O and S atoms from60 was used. The crystal geometry from X-ray diffraction experiments13 was used as the starting point in our calculations. Optimization of the atomic positions was carried out for all atoms in the irreducible part of the crystallographic cell with fixed cell parameters and space group. The scaling factor in the Pack–Monkhorst scheme was set at 8 in all crystal cell directions. More details on the results of the optimization process and electron density properties can be found elsewhere.51 The vibrational frequencies were computed via a coupled perturbed Hartree–Fock analytical approach at the G point of the Brillouin zone in the harmonic approximation.61–64 The dynamical matrix of the second derivatives was computed by numerical evaluation of the first derivatives of the analytical atomic gradients. For an oriented single crystal, the Raman intensity was associated with a specific mode of vibration. The intensities of the two polarized components in the spectra – parallel and perpendicular – were calculated for the corresponding powder structure, and the resulting total powder intensities were evaluated.65,66 The calculated Raman spectrum was visualized using the transverse optical (TO) modes presented in the pseudoVoigt functional form.

Results and discussion The substances in our study were obtained using a halocyclization reaction.67 Compounds 1–4 were prepared according to the method, described in previous work.68 Crystal structures were proven by X-ray diffraction experiments and described earlier.13,69–71 Herein, we can see various types of polyiodide organization: isolated slightly asymmetric triiodide 1 and asymmetric triiodide 2, which interact with the organic cation, and symmetric triiodide 3 and triiodide 4 as a part of the polyiodide formed as a result of Z-shaped I3  I2 interactions. These types of triiodide anions demonstrate the different crystal packing due to specific non-covalent interactions, thus we can presume their different spectral properties. Such structural-spectral correlations for different types of triiodides with thia- and oxazinoquinolinium cations are the subject of this study. The spectral

6166 | New J. Chem., 2015, 39, 6163--6170

Table 1 The spectral and structural data for the single crystal samples, the intense bands are shown in bold

Anion structure

Observed bands, cm1

DI–I, (I3) Å

1

I3

52, 116, 137

2.902 2.933

2

I3

66, 111, 127

2.892 2.949

3

I3

115, 131

2.909 2.910

4

I3  I2 D(I3  I2): 3.445 Å

62, 113, 134, 172

2.875 2.957 DI2: 2.747 Å

Number

Cation

and geometrical structural data for the new single crystals 1–4 are presented in Table 1. Crystal 1 is a representative of a slightly asymmetric triiodide; the difference in bond lengths is 0.031 Å, so an asymmetric band is observed (137 cm1), but it has a relatively low intensity. Triiodides in this crystal are isolated from each other as there are no I  I short contacts between the neighboring I3 anions, the shortest distance d(I3  I3) is 4.9 Å. In the low-frequency region, six bands have been observed and three of them were found in the boundaries typical for n2 vibrations: 52, 61 (observed only for spectra with the 633 nm laser line, very low intensity) and 67 cm1. In comparison with other experimental spectra, the bands near 61 or 67 cm1 can be assigned preferentially to the bending mode. A comparison with the calculated Raman spectra (see below) shows that the 52 cm1 band is more likely to be a dI–I band. Moreover, in the polarized spectra of the oriented crystal, the band near 52 cm1 becomes very intense in such orientations, when the polarization vector direction is nearly perpendicular to the triiodide anion, so it can be better attributed to the bending vibration. Other low-frequency bands in this region can be matched with various anion–anion, anion–cation or cation– cation vibrations. The cation–anion interactions in this case are held mainly via I  H contacts, though most of them have rather long distances (see Fig. 1). In the spectrum of this crystal, we can also observe strong C–I stretching vibration bands (498, 513 and 523 cm1) and this is in agreement with literature data: typical range of 490–620 cm1 for vibrations of –CH2I group, two or more strong bands.72 Triiodide in crystal 2 is more asymmetric, the difference in I–I bond lengths in triiodide is 0.058 Å. This triiodide is not isolated: the crystal structure is stabilized mainly via I  H

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015

NJC

Fig. 1

Paper

Geometry of the cation–anion interactions in crystal 1.

interactions and multiple short contacts between the iodine atoms in the anion and cation: I3  I–C(cat) with a distance of 3.495 Å, C(cat)–I  I–C(cat) with a distance of 3.872 Å. Such C(cat)– I  I–C(cat) contacts can significantly influence the crystal packing and direction of the intermolecular interactions.73 Triiodide 3 can be classified as a practically symmetric triiodide.74 The difference in bond lengths is 0.001 Å, which is comparable to the experimental error. The n3 mode (E1u,Pu) is inactive in Raman spectra, nevertheless the band still exists in the spectrum of triiodide 3, though it is of weak intensity. Such a fact is known mainly for triiodides in solution.1,75,76 For example, the selection rules can be broken due to anion–anion interactions, the existence of complexes with a solvent or the relative instability of symmetric triiodide, as for Me4NI3 solutions in nitrobenzene and ethanol.1,77 The existence of the n3 band in crystal 3 can be explained by changing the polarization of the I–I bonds due to the non-equivalent interaction of one terminal atom in the triiodide anion with an iodine atom in the cation. Their mutual orientation enables us to presume the formation of a halogen bond between them.17 In this way, the experimental registration of the changes in the number of observed spectral bands for symmetric triiodide can be useful in characterizing the non-covalent interactions in polyiodide systems. The complex structure of polyiodide 4 (Fig. 2) can be presented as an I3  I2 unit; it demonstrates the typical spectral behavior of this group of polyiodides. For example, the observed spectrum is analogous to the one reported for morpholinium I5, Et4NI725 or (N-n-pentylurotropinium)2I8.78 Its anions forms a Z-like structure and appears to be a widely known polyiodide structure motif.1,13,25 Together with the typical triiodide bands: 62 cm1 – dI–I, 114 cm1 – nsym, 134 cm1 – nasym, a band

Fig. 2 Geometrical characteristics of the complex polyiodide structure in crystal 4.

corresponding to bound iodine – 172 cm1 – is also observed in the experimental spectrum. It can be noted that in this case, the relative intensity of the nasym band is higher compared to simple triiodides. A shift in the band of the interacting iodine molecule to lower wavenumber is 8 cm1 in comparison with crystalline iodine itself.1,79 Together with a slight elongation of the I–I bond length (far less than 2.86 Å), it allows us to characterize the I3  I2 interactions in complex polyiodides as rather weak.28 A deeper understanding of the non-covalent interactions existing in the crystalline state can be achieved using the polarized Raman technique applied to the single crystal samples.80 The polarized Raman spectrum of single crystal 1 in specified orientations is presented in Fig. 3. According to the angular dependency of the nsym I3 vibration band, we can presume that the iodine atoms are located within the crystal bulk as shown above for the 01 and 801 positions, as there is only one type of triiodide anion in the unit cell. At the 80  101 orientation corresponding to the maximum intensity of the nsym I3 band, the position of the triiodide anion is almost collinear to both the directions of polarization vectors of incident and scattered light. The polarized Raman spectra of crystal 4 at various orientations are shown in Fig. 4a. The polarization vectors of both incident and scattered light are directed along the y axis. The maximum intensity is observed when the laser beam direction coincides with the direction of arrow 1. The observed overall intensities are much lower in directions 2 and 3. The spectrum in direction 2 reveals triiodide signals and only a bound iodine band along direction 3. Such spectral data allows us to presume that in our case, the Z-shaped polyiodide is planar and is located on the plane perpendicular to arrow 1. There is a slight shift in the peak position in directions 2 and 3 for the three

Fig. 3 Crystal 1 under 10-fold microscope magnification and the experimental dependency of the nsym I3 vibration band intensity on the rotation angle. Orientations of the crystal and the triiodide anions inside it are shown for the positions of the minimum and maximum of the curve. Polarization vectors for both incident and scattered light are directed along the y axis and are constant for all measurements.

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015

New J. Chem., 2015, 39, 6163--6170 | 6167

Paper

NJC

Fig. 4 (a) Polarized Raman spectra of the oriented single crystal 4 in three planes. (b) Dependency of the partial intensity (d, %) of the characteristic bands: n(I2) and nsym(I3) on the rotation angle. Crystal orientations corresponding to the maxima of the nsym(I3) and n(I2) bands are shown above.

bands for nsym, nasym and nI–I vibrations to higher wavenumber, but this is within the limits of experimental error. Crystal rotation in the plane that is perpendicular to arrow 1 allows us to obtain the angular dependency of the percentage intensity for each polyiodide band in order to obtain the geometrical characteristics of Z-shaped polyiodide. From the obtained data, it can be seen that the maximum intensity of the triiodide nsym band corresponds to the minimum intensity of the bound iodine nI–I band and the difference between the maxima of both bands is approximately 70  101 (Fig. 4b). The errors in rotation and the angle step is rather big (201) and did not allow us to achieve a good agreement with the experimental results (861), but this technique still allows to obtain some geometrical information on complex polyiodides. The underestimation of the I–I  I angle in the I3  I2 chain may also be due to its slight deviation out of the plane XY in terms of Fig. 4b. In the orientation corresponding to an angle of 01, the polarized spectra in parallel and perpendicular polarizations have been obtained and are shown in Fig. 4b. According to these data, the depolarization ratio81 have been computed for each band. For bending vibration, r = Iperpendicular/Iparallel equals 0.36; for the symmetric triiodide band, it is 0.1; for the asymmetric, it is 0.12; and for the I–I vibration of bound molecular iodine, the ratio equals 0.18. Though for all I–I experimental bands, these values are below 0.75, therefore they can all be attributed to polarized bands, the triiodide stretching bands can be characterized as the most polarized.82 Calculation of the solid state Raman intensities allows us to obtain the polycrystalline isotropic and single crystal directional intensities for the I–I and C–I vibrations. All I–I vibrations in crystal 1 have been found to be fully symmetric and refer to the Ag type of symmetry, which corresponds to the A1g(Sg+) type for a linear isolated molecule. The calculated vibration bands for triiodide anions in crystal 1 in the crystalline approximation has resulted in the values of 112 cm1 for nsym and 140 cm1 for nasym. In both cases, there is a slight 3–4 cm1 decrease in the calculated data when compared with the experimental values (Fig. 5).

6168 | New J. Chem., 2015, 39, 6163--6170

Fig. 5 Comparison of the experimental and calculated Raman spectra in the interval of the I–I vibrations. Curves are normalized to the experimental intensity of the highest peak; the units are arbitrary.

C–I vibrations are observed at 408, 502 and 510 cm1, the wavenumber in this case are also lower than the experimental ones, the difference here is within 13 cm1. These three calculated modes are slightly closer to each other than in the experiment: the distance between the first and the second maxima is 4 cm1, between the second and the third equals 8 cm1. In experiment these differences are 15 and 10 cm1.

Conclusions Spectral data for new thiazolo- or thia(oxy)azino-quinolinium crystalline polyiodides have been obtained. The asymmetric vibration band in crystal 3: (E)-3-(1-iodoethylidene)-2,3-dihydro-[1,4]thiazino[2,3,4-ij]quinolin-4-ium triiodide, C13H11I4NS with symmetric triiodide anion can be used as an evidence of halogen bond formation that can be presumed from the geometrical criterion. Complex polyiodide 4: (E)-3-(iodomethylene)-2,3-dihydro[1,4]oxazino-[2,3,4-ij]quinolin-4-ium triiodide–iodine (2 : 1) is described by the Z-like I3  I2 type, with non-covalent interactions inside the anion that are interpreted as weak or medium weak halogen bonds according to the shift in the n(I–I) vibration band to lower wavenumber. The angular dependency of the band intensities in the polarized spectra of crystal 4 (E)-3-(iodomethylene)-2,3-dihydro[1,4]oxazino-[2,3,4-ij]quinolin-4-ium triiodide–iodine (2 : 1) on crystal rotation on the plane of maximum spectral intensity allows us to obtain an estimation of the geometrical parameters of the zigzag-like structure. The selected level of theoretical approximation using periodic boundary conditions in the solid state reproduces the geometrical characteristics of the I–I bonds in crystal 1. The wavenumber corresponding to the basic I–I vibration modes and their Raman intensities have been obtained. The calculated values for the nsym, nasym and dI–I vibration modes are in good agreement with the experimental results.

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015

NJC

Acknowledgements This study was supported by the grant of Russian Ministry for Education and Science GZ729.

Notes and references 1 P. H. Svensson and L. Kloo, Chem. Rev., 2003, 103, 1649–1684. 2 S. Alvarez, J. Novoa and F. Mota, Chem. Phys. Lett., 1986, 132(6), 531–534. 3 R. P. Shibaeva and E. B. Yagubskii, Chem. Rev., 2004, 104, 5347–5378. 4 H. Stegemann, A. Rhode, A. Reiche, A. Shnittke and H. Fullbier, Electrochim. Acta, 1992, 37, 379–383. 5 I. Jerman, V. Jovanovski, A. Surca Vuk, S. B. Hocevar, M. Gaberscek, A. Jesih and B. Orel, Electrochim. Acta, 2008, 53, 2281–2288. 6 S. Punyani, P. Narayana, H. Singh and P. Vasudevan, J. Sci. Ind. Res., 2006, 165, 116–120. 7 P. V. Kulkarni, V. Arora and A. S. Rooney, Nuclear instruments and methods in physical research B, 2005, 241, 676–680. 8 M. C. Aragoni, M. Arca, F. Demartin, F. A. Devillanova, A. Garau, F. Isaia, V. Lippolis and G. Verani, J. Am. Chem. Soc., 2002, 124, 4538–4539. 9 A. J. Blake, W.-S. Li, V. Lippolis, S. Parsons and M. Schroder, Acta Crystallogr., Sect. B: Struct. Sci., 2007, 63, 81–92. 10 P. Auffinger, F. A. Hays, E. Westhof and P. S. Ho, Proc. Natl. Acad. Sci. U. S. A., 2004, 101, 16789–16794. 11 L. Kloo, J. Rosdahl and P. H. Svensson, Eur. J. Inorg. Chem., 2002, 1203–1209. 12 Cambridge Structural Database. Version 5.33. University of Cambridge, UK. 13 E. V. Bartashevich, I. D. Yushina, E. A. Vershinina, P. A. Slepukhin and D. G. Kim, J. Struct. Chem., 2014, 55, 112–119. 14 A. Abate, M. Brischetto, G. Cavallo, M. Lahtinen, P. Metrangolo, T. Pilati, S. Radice, G. Resnati, K. Rissanene and G. Terraneo, Chem. Commun., 2010, 46, 2724–2726. 15 F. C. Kupper, M. C. Feiters, B. Olofsson, T. Kaiho, S. Yanagida, M. B. Zimmermann, L. J. Carpenter, G. W. Luther III, Z. Lu, M. Jonsson and L. Kloo, Angew. Chem., Int. Ed., 2011, 50, 11598–11620. 16 T. Clark, M. Hennemann, J. S. Murray and P. Politzer, J. Mol. Modell., 2007, 13(2), 291–296. 17 P. Politzer, J. S. Murray and T. Clark, Phys. Chem. Chem. Phys., 2010, 12(28), 7748–7757. 18 E. V. Bartashevich and V. G. Tsirelson, Russ. Chem. Rev., 2014, 83, 1181–1203. 19 G. R. Desiraju, P. S. Ho, L. Kloo, A. C. Legon, R. Marquardt, P. Metrangolo, P. Politzer, G. Resnati and K. Rissanen, Pure Appl. Chem., 2013, 85, 1711–1713. 20 C. D. Antoniadis, G. J. Corban, S. K. Hadjikakou, N. Hadjiliadis, M. Kubicki, S. Warner and I. S. Butler, Eur. J. Inorg. Chem., 2003, 1635–1640. 21 Z. Yin, Q.-X. Wang and M.-H. Zeng, J. Am. Chem. Soc., 2012, 134, 4857–4863.

Paper

22 P. Metrangolo, F. Meyer, T. Pilati, G. Resnati and G. Terraneo, Angew. Chem., Int. Ed., 2008, 47, 6114–6127. 23 D. Tyagi, S. Varma, K. Bhattacharya, D. Jain, A. K. Tripathi, C. G. S. Pillai and S. R. Bharadwaj, Int. J. Hydrogen Energy, 2012, 37, 3621–3625. 24 A. J. Blake, F. A. Devillanova, R. O. Gould, W.-S. Li, V. Lippolis, S. Parsons, C. Radek and M. Schroder, Chem. Soc. Rev., 1998, 27, 195–205. 25 P. Deplano, J. R. Ferraro, M. L. Mercuri and E. F. Trogu, Coord. Chem. Rev., 1999, 188, 71–95. 26 27 F. Bertolotti, A. V. Shishkina, A. Forni, G. Gervasio, A. I. Stash and V. G. Tsirelson, Cryst. Growth Des., 2014, 14, 3587–3595. 28 M. Arca, M. C. Aragoni, F. A. Devillanova, A. Garau, F. Isaia, V. Lippolis, A. Mancini and G. Verani, Bioinorg. Chem. Appl., 2006, 1–12. 29 V. Daga, S. K. Hadjikakou, N. Hadjiliadis, M. Kubicki, J. H. Z. dos Santos and I. S. Butler, Eur. J. Inorg. Chem., 2002, 1718–1728. 30 M. C. Aragoni, M. Arca, C. Caltagirone, C. Castellano, F. Demartin, A. Garau, F. Isaia, V. Lippolis, R. Montis and A. Pintus, CrystEngComm, 2012, 14, 5809–5823. 31 F. Isaia, M. C. Aragoni, M. Arca, F. Demartin, F. A. Devillanova, G. Ennas, A. Garau, V. Lippolis, A. Mancini and G. Verani, Eur. J. Inorg. Chem., 2009, 3667–3672. 32 R. D. Bailey, G. W. Drake, M. Grabarczyk, T. W. Hanks, L. L. Hook and W. T. Pennington, J. Chem. Soc., Perkin Trans. 2, 1997, 2773–2780. 33 R. Wojciechowski, K. Yamamoto and K. Yakushi, Phys. Rev. B: Condens. Matter Mater. Phys., 2003, 67, 224105. 34 R. Swietlik, D. Schweitzer and H. J. Keller, Phys. Rev. B: Condens. Matter Mater. Phys., 1987, 36, 6681–6688. 35 G. Manca, A. Ienco and C. Mealli, Cryst. Growth Des., 2012, 12(4), 1762–1771. 36 P. K. Bakshi, M. A. James, T. S. Cameron and O. Knop, Can. J. Chem., 1996, 74, 559–573. 37 C. Horn, M. Scudder and I. Dance, CrystEngComm, 2001, 2, 1–6. 38 M. van Megen and G. J. Reiss, Inorganics, 2013, 1, 3–13. 39 M. Muller, M. Albrecht, V. Gossen, T. Peters, A. Hoffmann, G. Raabe, A. Valkonen and K. Rissanen, Chem. – Eur. J., 2010, 16, 12446–12453. 40 M. Giese, M. Albrecht, C. Bohnen, T. Repenko, A. Valkonen and K. Rissanen, Dalton Trans., 2014, 43, 1873. 41 F. Groenewald, C. Esterhuysen and J. Dillen, Theor. Chem. Acc., 2012, 131, 1281. 42 X. Chen, M. A. Leugers, T. Kirch and J. Stanley, J. Spectrosc., 2015, 2015, 518054, DOI: 10.1155/2015/518054. 43 J. Hu, D. Wang, W. Guo, S. Du and Z. K. Tang, J. Phys. Chem. C, 2012, 116, 4423–4430. 44 M. C. C. Ribeiro, J. Phys. Chem. B, 2012, 116, 7281–7290. 45 A. Mukherjee and G. R. Desiraju, IUCrJ, 2014, 1, 49–60. 46 G. A. Landrum, N. Goldberg and R. HoVmann, J. Chem. Soc., Dalton Trans., 1997, 3605–3613.

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015

New J. Chem., 2015, 39, 6163--6170 | 6169

Paper

NJC

47 N. A. Al-Hashimi and Y. H. A. Hussein, Spectrochim. Acta, Part A, 2010, 75, 198–202. 48 M. C. Aragoni, M. Arca, F. A. Devillanova, M. l. B. Hursthouse, S. L. Huth, F. Isaia, V. Lippolis and A. Mancini, CrystEngComm, 2004, 6(87), 540–542. 49 V. D. Filimonov, E. A. Krasnokutskaya, O. Kh. Poleshchuk, Yu. A. Lesina and V. K. Chaikovskii, Russian Chemical Bulletin, International Edition, 2006, 55, 1328–1336. 50 E. B. Starikov, Int. J. Quantum Chem., 1997, 64, 473–479. 51 E. V. Bartashevich, I. D. Yushina, A. I. Stash and V. G. Tsirelson, Cryst. Growth Des., 2014, 14(11), 5674–5684. 52 M. Otsuka, H. Mori, H. Kikuchi and K. Takano, Comput. Theor. Chem., 2011, 973, 69–75. 53 K. N. Robertson, T. S. Cameron and O. Knop, Can. J. Chem., 1996, 74, 1572. 54 I. Yushina, B. Rudakov, I. Krivtsov and E. Bartashevich, J. Therm. Anal. Calorim., 2014, 118, 425–429. 55 P. Deplano, F. A. Devillanova, J. R. Ferraro, M. L. Mercuri, V. Lippolis and E. F. Trogu, Appl. Spectrosc., 1994, 48, 1236–1241. 56 C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785–789. 57 A. D. Becke, Phys. Rev., 1988, 38, 3098–3100. 58 R. Dovesi, R. Orlando, B. Civalleri, C. Roetti, V. R. Saunders and C. M. Zicovich-Wilson, Z. Kristallogr., 2005, 220, 571. 59 Iodine Basis set. URL: http://www.tcm.phy.cam.ac.uk/ Bmdt26/basis_sets/I_basis.txt. 60 C. Gatti, V. R. Saunders and C. Roetti, J. Chem. Phys., 1994, 101, 10686–10696. 61 F. Pascale, C. M. Zicovich-Wilson, F. Lopez, B. Civalleri, R. Orlando and R. Dovesi, J. Comput. Chem., 2004, 25, 888–897. 62 C. M. Zicovich-Wilson, F. Pascale, C. Roetti, V. R. Saunders, R. Orlando and R. Dovesi, J. Comput. Chem., 2004, 25, 1873–1881. ´rat, R. Orlando and R. Dovesi, J. Chem. 63 M. Ferrero, M. Re Phys., 2008, 128, 014100. ´rat, R. Orlando and R. Dovesi, J. Comput. 64 M. Ferrero, M. Re Chem., 2008, 29, 1450–1459.

6170 | New J. Chem., 2015, 39, 6163--6170

65 L. Maschio, B. Kirtman, M. Rerat, R. Orlando and R. Dovesi, J. Chem. Phys., 2013, 139, 164101. 66 L. Maschio, B. Kirtman, M. Rerat, R. Orlando and R. Dovesi, J. Chem. Phys., 2013, 139, 164102. 67 K. E. Harding and T. H. Tiner, in Comprehensive Organic Synthesis, ed. B. M. Trost and I. Fleming, Pergamon Press, New York, 1991, vol. 4, p. 363. 68 D. G. Kim, Chem. Heterocycl. Compd., 2008, 11, 1664–1668. 69 V. I. Batalov, A. Dikhtiarenko, I. D. Yushina, E. V. Bartashevich, D. G. Kim and S. Garcı´a-Granda, Z. Kristallogr. – New Cryst. Struct., 2014, 229, 213–214. 70 V. I. Batalov, D. G. Kim, A. Dikhtiarenko, Z. Amghouz, E. V. Bartashevich and S. Garcı´a-Granda, Z. Kristallogr. – New Cryst. Struct., 2014, 229, 211–212. 71 V. I. Batalov, A. Dikhtiarenko, I. D. Yushina, E. V. Bartashevich, D. G. Kim and S. Garcı´a-Granda, Z. Kristallogr. – New Cryst. Struct., 2014, 229, 195–196. 72 G. Socrates, Infrared and Raman Characteristic Group Frequencies: Tables and Charts, John Wiley & Sons, Ltd, Chichester, England, 3rd edn, 2001, p. 347. ´. Garcı´a, P. Cabildo, R. M. Claramunt, E. Pinilla, 73 M. A M. R. Torres, I. Alkorta and J. Elguero, Inorg. Chim. Acta, 2010, 363, 1332–1342. 74 P. H. Svensson and L. Kloo, J. Chem. Soc., Dalton Trans., 2000, 2449–2455. 75 K. R. Loos and A. C. Jones, J. Phys. Chem., 1974, 78(22), 2306–2307. 76 A. E. Johnson and A. B. Myers, J. Phys. Chem., 1996, 100(19), 7778–7788. 77 J. Breckenridge, W. Warzecha and T. Surles, Inorg. Nucl. Chem. Lett., 1973, 9, 437–442. 78 H. Mittag, H. Stegemann and H. Fullbier, J. Raman Spectrosc., 1989, 29, 251–255. 79 F. Vovelle and G. G. Dumas, Mol. Phys., 1977, 34, 1661–1673. 80 A. Congeduti, M. Nardone and P. Postorino, Chem. Phys., 2000, 256, 117–123. 81 C. D. Allemand, Appl. Spectrosc., 1970, 24(3), 348–353. 82 D. A. Long, Proc. R. Soc. London, Ser. A, 1953, 217, 203–221.

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2015