High Pressure Effects on Zwitterionic and Thione Mesomeric

Jul 19, 2017 - ... Adam Mickiewicz University, Umultowska 89 b, 61-614 Poznań, Poland ... Direct and Inverse Relations between Temperature and Pressur...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/JPCC

High Pressure Effects on Zwitterionic and Thione Mesomeric Contributions in 2‑Benzimidazole-2-Thione Hanna Tomkowiak and Andrzej Katrusiak* Faculty of Chemistry, Adam Mickiewicz University, Umultowska 89 b, 61-614 Poznań, Poland S Supporting Information *

ABSTRACT: High pressure reduces the zwitterionic mesomeric contribution and increases the thione contribution in 2-benzimidazole-2-thione. These mesomeric changes are manifested in the shortened bond S−C and elongated bond C−N in the S−C−N moiety. These transformations are consistent with the le Chatelier law, as they counteract the increase of electrostatic interactions when the intermolecular distances between electronegative sulfur atoms and arene π-electrons are compressed. The changing interactions affect the crystal strain and its structural transformations. Consequently, the crystal compression and thermal expansion initially, until about 1.0 GPa, are inconsistent with the inverse relationship rule of pressure and temperature effects. Some anomalous features of the thermal expansion can be associated with isostructural transformations of the crystal.



INTRODUCTION Transformation of molecules of chemical compounds and the intermolecular interactions in crystals can be efficiently studied by high-pressure structural determinations.1−8 It was established that high pressure affects the conformation of organic compounds, for example in the crystal 1,1,2-trichloroethane9 and 1,1,2,2-tetrachloroethane10 molecules transform between gauche and s-transoid conformations, and pyrrolidine between C-off and N-off envelopes.11 Valuable information was obtained about the pressure effects on intermolecular interactions, hydrogen bonds OH···O,12−16 NH···O,15−22 NH···N,21−24 weak CH···O,25−28 and CH···π bonds,22,29 halogen interactions,30,31 and the competitions between different types of cohesion forces in crystals.32,33 However, strikingly little is known about mesomeric transformations under pressure. To our knowledge there are only a few reports on the effects of pressure on the bond orders and electronic structure of organic molecules; the carbonyl CO bond lengths were studied for compressed CO···H−O−C bonds,34 mesomeric forms were considered for compressed urea21 and the aromaticity of syn1,6:8,13-biscarbonyl[14]annulene (BCA) was investigated recently.35 This limited number of such studies contrasts with the intensive research on pressure-induced polymerization,36−39 for which the information about the electronic structure of molecules in the confined environment is essential. So far only one determination of bond length in CS2 as a function of pressure prior to the pressure-induced polymerization of this compound was reported.40 In order to investigate the pressure effects on the molecular structure, we have chosen a compound capable of transforming between different mesomeric forms. 2Benzimidazole-2-thione (denoted BzImS) can be present in the thione, thiol, and zwitterionic forms (Figure 1)41 and the thione and zwitterionic forms can ideally bind both to the neutral ( S−M) and cationic (−S−−M+) metal atoms. These features of BzImS and its derivatives are applied in anticorrosion paints, © 2017 American Chemical Society

Figure 1. Molecular diagrams of (1) thione and (2) zwitterionic mesomers as well as (3) the thiol H-tautomer of 2-benzimidazole-2thione (BzImS). The atomic labels applied in this study (half of the molecule is independent due to mirror plane passing through atoms S1 and C1) are marked in formula (1).

owing to their capability of coordinating metal atoms and forming a protective layer on the metal surface. Local strains on the metal surfaces can be mimicked by high-pressure conditions and studies, which can provide new structural information about the protective performance of BzImS. The small contribution of the thione form compared to the zwitterionic form would be consistent with the acidity of thioalcohols. It was established that in the crystalline state BzImS is present in the form of mixed thione-zwitterion mesomers and the BzImS molecules are NH···S hydrogen bonded into chains.41,42 The main aim of our present study was to investigate the effects of high pressure on the molecular structure and hydrogenbonding pattern in the BzImS crystal as well as the comparison of the compression and thermal expansion of this highly anisotropic crystal. Received: June 21, 2017 Revised: July 14, 2017 Published: July 19, 2017 18830

DOI: 10.1021/acs.jpcc.7b06083 J. Phys. Chem. C 2017, 121, 18830−18836

Article

The Journal of Physical Chemistry C



EXPERIMENTAL SECTION High-pressure measurements were performed in a MerrillBassett diamond anvil-cell (DAC) modified by mounting the diamond anvils on steel supports with conical windows.43 Pressure in the DAC chamber was calibrated by the rubyfluorescence method, 44,45 with a Photon Control Inc. spectrometer of enhanced resolution, affording an accuracy of 0.02 GPa; the calibration was repeated before and after each diffraction measurement. Single crystals of BzImS were grown in situ in isochoric conditions (Figure 2) from the solution in

the most accurate ones have been selected for plotting the crystal compression (finally all the pressure range was treated as one phase of BzImS, as explained in Results and Discussion). The CrysAlis software46 was used for controlling the highpressure diffraction experminents47 and preliminary data reduction. Reflections have been corrected for the effects of the DAC absorption, sample shadowing by the gasket, the sample absorption,46 and the reflections overlapping with diamond reflections were eliminated. All structures could be solved straightforwardly by direct methods, but the same model was refined by full-matrix least-squares.48 For low- and hightemperature data anisotropic temperature factors were applied for all atoms except H atoms, while for high-pressure data anisotropic temperature factors were applied for all N and S atoms, while atoms C and H were refined with isotropic parameters. Selected crystal and experimental data for highpressure structures are listed in Table 1, while the detailed information is given in Tables S1−S3 in the Supporting Information and have been deposited in the CIF format at the Cambridge Structural Database as Supplementary publications CCDC 1555097−1555120. Their copies can be obtained free of charge from Web site http://www.ccdc.cam.ac.uk/conts/ retrieving.html.



RESULTS AND DISCUSSION

The isochoric crystallizations of BzImS from methanol and aqueous solutions yield the crystal of compressed ambientpressure phase,41,42 shown in Figure 2 and Figure 3. No clear indications of phase transitions nor BzImS solvates have been found up to 2.6 GPa. The crystals are monoclinic, of space group P21/m, Z = 2. The crystal compression and thermal expansion are compared in Figure 4. The BzImS crystal structure is mainly governed by NH···S hydrogen bonds linking the molecules into ribbons extending along direction [y]. Atoms S1 and C1 of the BzImS molecule are located on the crystallographic mirror plane, and each of two symmetry-dependent NH groups forms one NH···S bond. Thus, each BzImS molecule forms four NH···S hydrogen bonds, two to each of its neighboring molecules in the ribbon (Figure 3). There are only weak van der Waals contacts between the ribbons. All ribbons are approximately parallel to crystal plane (102). The crystal compression corresponds to the intermolecular interactions: the least compressed is the [y] direction along the double NH···S bonded ribbons (Figure 4). Initially the most compressed is the unit-cell parameter a, however its compression is strongly nonlinear and about 2.0 GPa it becomes “harder” than parameter b, along the ribbons; then the

Figure 2. Stages of isochoric growth of a single crystal of the solution of BzImS dissolved in methanol in the DAC chamber: (a) one seed at 423 K, (b) 393 K, (c) 323 K, and (d) 296 K/1.20 GPa. One bigger and several smaller ruby chips lie on the left and upper sides of the chamber.

water, methanol, as well as methanol/ethanol/water (16:3:1 vol.) mixture, and the X-ray diffraction data were measured with a KUMA KM4-CCD diffractometer, with graphitemonochromated λ(MoKα) = 0.71073 Å from a sealed tube. Some of the measurements were performed for the crystals compressed in the hydrostatic fluid (solution) without recrystallization. We were particularly concerned about the pressure region around 0.5 GPa, because it appeared that there the crystal compression is anomalous. About 25 full diffraction data sets were measured for several sample crystals, and then

Table 1. Selected Crystal Data for High-Pressure BzImS Structures pressure (GPa) crystal system space group unit cell:

a (Å) b (Å) c (Å) β (°)

volume (Å3) Dcalc (g/cm3) Rint final R1/wR2 (all data)

0.18(2) monoclinic P21/m 4.8528(4) 8.5459(7) 8.246(18) 91.94(3) 341.8(8) 1.460 0.0328 0.0503/0.0871

0.48(2) monoclinic P21/m 4.8045(3) 8.5412(4) 8.132(15) 92.40(2) 333.4(6) 1.496 0.0306 0.0457/0.0886 18831

0.90(2) monoclinic P21/m 4.7172(6) 8.4993(9) 8.03(3) 93.12(5) 321.7(12) 1.551 0.0292 0.0378/0.0959

1.68(2) monoclinic P21/m 4.6546(4) 8.4427(5) 7.855(16) 94.11(3) 307.9(6) 1.620 0.0354 0.0419/0.0946

2.58(2) monoclinic P21/m 4.6355(6) 8.4117(9) 7.635(19) 95.29(5) 296.4(7) 1.683 0.0389 0.0561/0.1259

DOI: 10.1021/acs.jpcc.7b06083 J. Phys. Chem. C 2017, 121, 18830−18836

Article

The Journal of Physical Chemistry C

Figure 3. (a) Autostereograms of two overlapping NH···S bonded ribbons in the BzImS structure viewed perpendicular to the molecules and the voids within these ribbons at 0.1 MPa and 2.58 GPa. The voids, marked in yellow, have been calculated with the probe radius 0.87 Å and grid spacing 0.1 Å.49 (b) Projection of BzImS along the [010] direction at 0.1 MPa. The hydrogen bonds have been marked by cyan lines. Greek letter Δ denotes the distance between least-squares planes fitted to the atoms of NH···S bonded molecules, while Δ′ and Δ″ are the next shortest interplanar distances (Δ″ = Δ + Δ′).

Figure 4. (a) Compression at 296 K (full circles) and thermal expansion at 0.1 MPa (open circles) of the unit-cell dimensions of BzImS, related to the 0.1 MPa/296 K values (cf. Tables S1−S3). The lines joining the points are for guiding the eye only. The inset enhances the temperature dependence of parameter b. (b) Graphical representation of the BzImS crystal strain tensors due to the hydrostatic compression at 0.1 MPa, at 2.58 GPa and due to the thermal expansion at 296 K.60

“softest” becomes parameter c that initially was intermediate between a and b. The nonlinear compression of a contrasts with nearly linear compression of b, c, and with the linear changes of angle β as a function of pressure (Figure S1). It is remarkable that the crystal is least compressed in the [y] direction along the NH···S bonds. They are compressed nearly linearly in all the pressure range to 2.58 GPa (Figure 5), from 3.378 to 3.293 Å, that is, by 0.08 Å. The NH···S bonds are

nearly parallel to the [y] direction and it well agrees with the shortening of b by 0.163 Å (two N···S distances per one b period, cf. Figure 3). It initially appeared from the compression measurements that the crystal displays an anomalous behavior around 0.5 GPa, which could be associated with an isostructural phase transition28,50−53 (see Experimental Section). Although some anomalous changes were found in the BzImS structure, too, they take place in a broad pressure region to about 1.5 18832

DOI: 10.1021/acs.jpcc.7b06083 J. Phys. Chem. C 2017, 121, 18830−18836

Article

The Journal of Physical Chemistry C

parameter b, along the ribbons, which is consistent (according to the inverse rule) with the initial small compression along this direction. The strongest thermal expansion of c is consistent with its strongest compression above 1.0 GPa; however, the initial (between 0.1 MPa and 1.0 GPa) compression of parameters a and c and their thermal expansion are inconsistent with the inverse rule. For this pressure range parameter c is less compressed than parameter a, while it is the c parameter that expands strongest. Interestingly, the strongly nonlinear compression of a is not reflected in its thermal expansion, either. It can be also noted that parameter b displays a subtle, but clear, negative linear thermal expansion (NLE) between 120 and 140 K, above which it resumes the positive expansion until about 300 K; between 300 and 343 K the NLE of b is observed again. Nonetheless, at 120 K, parameter b is marginally (within one ESD) longer than at 300 K. The NLE of parameter b can be due to the vibrations off-plane of the molecule, along its axis of the smallest inertia, that is, along the axis along bond CS. The NH···S bonded molecules viewed down this smallest-inertia axis (Figure S3) illustrate that the atomic thermal ellipsoids are elongated perpendicular to the molecular plane. The librational motion of molecules moves the average atomic positions closer to the axis of vibrations. When the amplitudes of the librations decrease, the molecules and NH···S bonds become better aligned (in terms of their smaller atomic displacements) along [y], which generates the NLE of the crystal. The off-plane vibrations hardly affect the length of NH···S bond, measured as the distance between average atomic positions (Figure 5). The strongly nonlinear compression of parameter a can be connected with the interplane distances between the molecules within one ribbon and between the neighboring ribbons. These interplane distances have been denoted as Δ, Δ′, and Δ″ and indicated in Figure 3. The intraribbon distance between the planes of NH···S bonded molecules initially increases with pressure from 0.875 Å at 0.1 MPa to 0.96 Å about 1.5 GPa (Figure 6). The next shortest interplanar distance Δ′ is strongly

Figure 5. Shortest intermolecular distances C···S, N···S, H···S, and H··· H as a function of pressure (full circles) and temperature (open circles) in BzImS. The lines joining the points are to guide the eye only. The ESDs are smaller than the plotted symbols. Symmetry codes: (i) 1 + x, y, z; (ii) −x, 0.5 + y, 1 − z; (iii) 2 − x, 1 − y, −z.

GPa, and therefore, all the investigated pressure region to 2.6 GPa has been treated as one phase of monotonous compression for the further discussion in this paper. The crystal compression is often compared to the thermal expansion with a reference to the rule of inverse pressure and temperature effects.54−57 It is believed that most of crystals comply with this rule and that its exempts are only the crystals of wurcite and few related minerals. However, it was shown recently that the inverse rule cannot be applied to a metal− organic framework,58 and to a molecular crystal of orthoxylene.59 In the BzImS crystal (Figure 4), the least expanding is

Figure 6. (a) Distance Δ between planes fitted to NH···S bonded molecules and the shortest inter-ribbon plane-to-plane distances Δ′ and Δ″ (cf. Figure 3) as a function of pressure (full circles) and temperature (open circles). (b) Angle φ between atom S, the midpoint of bond C2−C2′ and atom Si from the close molecule of the neighboring ribbon as a function of pressure (full circles) and temperature (open circles), as indicated in the inset. Symmetry code: (i) 1 + x, y, z. 18833

DOI: 10.1021/acs.jpcc.7b06083 J. Phys. Chem. C 2017, 121, 18830−18836

Article

The Journal of Physical Chemistry C nonlinear, too, in such a way that the nonlinearity of Δ, including its NLC and positive compression regions, is nearly fully compensated. This can be seen in the nearly linear compression of Δ″ (Figure 6a), because Δ″ = Δ + Δ′(Figure 3). The strongly nonlinear compression of interplanar distances Δ and Δ′ have little effect on b, which is parallel to the molecular planes (and nearly parallel to the NH···S bonds). On the other hand, the magnitudes of Δ and Δ′ are connected with the unit-cell parameter a and its nonlinear compression clearly corresponds to the pressure dependence of Δ (Figure 6a). Another structural transformation contributing to the nonlinear compression of parameters a and c is the shift of the neighboring ribbons one with respect to the other, as illustrated in Figure 6b and measured by angle φ. The origin of this shift can be associated with the compression of the “softest” H···H contacts (Figure 5). The magnitudes of changes in molecular dimensions are commensurate with the estimated standard deviations (ESDs) and therefore of much smaller statistical significance than the changes in intermolecular contacts. Nonetheless, clear trends are observed in the length of bonds C−S, shortening by 0.06 Å, and C−N, lengthening by 0.03 Å (Figure 7). These changes are

shows clearly that they can be divided into compounds where the thione group is located in moieties of saturated C−C bonds, for example Csp3−C(S)−Csp3,63−67 and a group where the thione bond is conjugated. 68−72 The effect of the zwitterionic contribution for the CS bond length is clearly observed in the crystal structure of 1,3-dibenzyl-2-(4′,4′dimethyl-2′,6′-dithioxocyclohexylidene) hexahydropyrimidine (refcode LEVRIN), where CS distances become 1.667 and 1.641 Å.73 On the other hand, according to a survey of the CSD, the average length of double bond CS is 1.681 Å,74 nearly exactly as this in BzImS at 0.1 MPa and considerably longer than that in CS2 − 1.546(1) Å at 5.3 K/0.1 MPa75 and 1.550(2) Å at 295K/3.7 GPa.40 This may indicate that most of CS bonds in the CSD have a considerable contribution of C−S− mesomer.



CONCLUSIONS High-pressure crystallizations of BzImS of methanol and aqueous solution yield only unsolvated crystals up to at least 2.6 GPa. This behavior is different from that observed in the thiourea where the high-pressure crystallization promote the formation of hydrates.76 The persistent crystallization of anhydrous and unsolvated BzImS is connected to its NH···S hydrogen bonds, enhanced by high pressure. Small systematic changes in the bond lengths of moiety N−CS are consistent with the increased contribution of the thione mesomer and reduced contribution of the zwitterionic mesomer, as well as with the reduced electrostatic interactions between molecules in the compressed crystal. The changes of the BzImS crystal compression and thermal expansion can be due to supercritical effects or a subtle isostructural solid−solid phase transition.



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.7b06083. Crystal data and structure-refinements details of 2benzimidazole-2-thione structures (Tables S1−S3); unit-cell dimensions of 2-benzimidazole-2-thione as a function of pressure and temperature (Figure S1); the crystal structure of BzImS along [100] and [001] direction (Figure S2); atomic thermal ellipsoids at 120, 195, and 296 K (Figure S3); the molecular voids volume as a function of pressure (Figure S4); the intermolecular S1···H3 and C···C distances and angle N−H···S as a function of pressure (Figure S5; PDF).

Figure 7. Bond lengths S−C1, N−C1, and, for reference, the bond N−C2 (green crosses), as well as the average length of C−C bonds in the benzene ring (black crosses) as a function of pressure (the reference values are presented without their ESDs for clarity). The lines joining the points are for guiding the eye only.

consistent with the increased contribution of thione mesomer (form 1 in Figure 1) and reduced zwitterionic form 2. According to Allmann,61 the bond between atoms N1 and C1 in BzImS correspond to the bond order 1.4 at 0.1 MPa and 1.2 at 2.58 GPa. This change can originate from compressed intermolecular contacts in the structure and stronger interactions between the electronegative sulfur atom and π electrons of the close benzene ring. The electrostatic component of these repulsive interactions can be reduced by decreased negative net charge on the S atom. According to Kitajgorodski,62 the Csp3−Csp3 bond length change of 0.05 Å corresponds to the energy of about 4 kJ mol−1. The bond length changes in BzImS are of about this order of magnitudes, and these changes are compatible with the work energy of about 12 kJ mol−1 performed by the pressure of 2.0 GPa on the BzImS crystal. Our inspection of the CS bond in structures deposited in the Cambridge Crystallographic Database (CSD)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Andrzej Katrusiak: 0000-0002-1439-7278 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We are grateful to the Wielkopolskie Centrum Zaawansowanych Technologii for the experimental support. 18834

DOI: 10.1021/acs.jpcc.7b06083 J. Phys. Chem. C 2017, 121, 18830−18836

Article

The Journal of Physical Chemistry C



(21) Roszak, K.; Katrusiak, A. Giant Anomalous Strain Between High-Pressure Phases and the Mesomers of Urea. J. Phys. Chem. C 2017, 121, 778−784. (22) Lee, R.; Howard, J. A.; Probert, M. R.; Steed, J. W. Structure of Organic Solids at Low Temperature and High Pressure. Chem. Soc. Rev. 2014, 43, 4300−4311. (23) Olejniczak, A.; Anioła, M.; Szafrański, M.; Budzianowski, A.; Katrusiak, A. New Polar Phases of 1,4-Diazabicyclo[2.2.2]Octane Perchlorate, an NH+···N Hydrogen-Bonded Ferroelectric. Cryst. Growth Des. 2013, 13, 2872−2879. (24) Zieliński, W.; Katrusiak, A. Hydrogen Bonds NH···N in Compressed Benzimidazole Polymorphs. Cryst. Growth Des. 2013, 13, 696−700. (25) Patyk, E.; Marciniak, J.; Tomkowiak, H.; Katrusiak, A.; Merz, K. Isothermal and Isochoric Crystallization of Highly Hygroscopic Pyridine N-Oxide of Aqueous Solution. Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater. 2014, 70, 487−491. (26) Gajda, R.; Katrusiak, A. Pressure-Promoted CH···O Hydrogen Bonds in Formamide Aggregates. Cryst. Growth Des. 2011, 11, 4768− 4774. (27) Dziubek, K. F.; Jęczmiński, D.; Katrusiak, A. Pressure-Generated Hydrogen Bonds and the Role of Subtle Molecular Features in Tetrahydrofuran. J. Phys. Chem. Lett. 2010, 1, 844−849. (28) Patyk, E.; Skumiel, J.; Podsiadło, M.; Katrusiak, A. HighPressure (+)-Sucrose Polymorph. Angew. Chem., Int. Ed. 2012, 51, 2146−2150. (29) Katrusiak, A.; Podsiadło, M.; Budzianowski, A. Association CH··· π and No Van Der Waals Contacts at the Lowest Limits of Crystalline Benzene I and II Stability Regions. Cryst. Growth Des. 2010, 10, 3461−3465. (30) Rajewski, K. W.; Andrzejewski, M.; Katrusiak, A. Competition Between Halogen and Hydrogen Bonds in Triiodoimidazole Polymorphs. Cryst. Growth Des. 2016, 16, 3869−3874. (31) Wang, K.; Duan, D.; Zhou, M.; Li, S.; Cui, T.; Liu, B.; Liu, J.; Zou, B.; Zou, G. Structural Properties and Halogen Bonds of Cyanuric Chloride under High Pressure. J. Phys. Chem. B 2011, 115, 4639− 4644. (32) Podsiadło, M.; Olejniczak, A.; Katrusiak, A. Halogen···Halogen Contra C-H···Halogen Interactions. CrystEngComm 2014, 16, 8279− 8285. (33) Destro, R.; Sartirana, E.; Loconte, L.; Soave, R.; Colombo, P.; Destro, C.; Lo Presti, L. Competing CO···CO, C−H···O, Cl···O, and Cl···Cl Interactions Governing the Structural Phase Transition of 2,6-Dichloro-p-Benzoquinone at Tc = 122.6 K. Cryst. Growth Des. 2013, 13, 4571−4582. (34) Katrusiak, A. High-Pressure X-Ray Diffraction Studies on Organic Crystals. Cryst. Res. Technol. 1991, 26, 523−531. (35) Casati, N.; Kleppe, A.; Jephcoat, A. P.; Macchi, P. Putting Pressure on Aromaticity Along with In Situ Experimental Electron Density of a Molecular. Nat. Commun. 2016, 7, 10901. (36) Wang, Y.; Wang, L.; Zheng, H.; Li, K.; Andrzejewski, M.; Hattori, T.; Sano-Furukawa, A.; Katrusiak, A.; Meng, Y.; Liao, F.; Hong, F.; Mao, H.-k. Phase Transitions and Polymerization of C6H6C6F6 Cocrystal under Extreme Conditions. J. Phys. Chem. C 2016, 120, 29510−29519. (37) Pravica, M.; Sneed, D.; Smith, Q.; Billinghurst, B.; May, T.; White, M.; Dziubek, K. A Novel Synthesis of Polymeric CO via Useful Hard X-Ray Photochemistry. Cogent Phys. 2016, 3, 1169880. (38) Santoro, M.; Scelta, D.; Dziubek, K.; Ceppatelli, M.; Gorelli, F. A.; Bini, R.; Garbarino, G.; Thibaud, J.-M.; Di Renzo, F.; Cambon, O.; Hermet, P.; Rouquette, J.; van der Lee, A.; Haines, J. Synthesis of 1D Polymer/Zeolite Nanocomposites under High Pressure. Chem. Mater. 2016, 28, 4065−4071. (39) Scelta, D.; Ceppatelli, M.; Bini, R. Pressure-Induced Polymerization of Fluid Ethylene. J. Chem. Phys. 2016, 145, 164504−1− 164504−7. (40) Dziubek, K.; Katrusiak, A. Compression of Intermolecular Interactions in CS2 Crystal. J. Phys. Chem. B 2004, 108, 19089−19092.

REFERENCES

(1) Fabbiani, F. P. A.; Pulham, C. R.; Warren, J. E. A High-Pressure Polymorph of Propionamide from In Situ High-Pressure Crystallisation from Solution. Z. Kristallogr. - Cryst. Mater. 2014, 229, 667− 675. (2) Neumann, M. A.; Van De Streek, J.; Fabbiani, F. P. A.; Hidber, P.; Grassmann, O. Combined Crystal Structure Prediction and HighPressure Crystallization in Rational Pharmaceutical Polymorph Screening. Nat. Commun. 2015, 6, 7793. (3) Roszak, K.; Katrusiak, A.; Katrusiak, A. High-Pressure Preference for the Low Z’ Polymorph of a Molecular Crystal. Cryst. Growth Des. 2016, 16, 3947−3953. (4) Boldyreva, E. V.; Shakhtshneider, T. P.; Ahsbahs, H.; Sowa, H.; Uchtmann, H. Effect of High Pressure on the Polymorphs of Paracetamol. J. Therm. Anal. Calorim. 2002, 68, 437−452. (5) Pravica, M.; Liu, Y.; Robinson, J.; Velisavljevic, N.; Liu, Z.; Galley, M. A High-Pressure Far- and Mid-Infrared Study of 1,1-Diamino-2,2Dinitroethylene. J. Appl. Phys. 2012, 111, 103534−103534. (6) Tschauner, O.; Kiefer, B.; Lee, Y.; Pravica, M.; Nicol, M.; Kim, E. Structural Transition of PETN-I to Ferroelastic Orthorhombic Phase PETN-III at Elevated Pressures. J. Chem. Phys. 2007, 127, 094502− 094502. (7) Anioła, M.; Katrusiak, A. Pressure-Preferred Symmetric Reactions of 4,4′-Bipyridine Hydrobromide. CrystEngComm 2016, 18, 3223− 3228. (8) Cai, W.; Katrusiak, A. Conformationally Assisted Negative Area Compression in Methyl Benzoate. J. Phys. Chem. C 2013, 117, 21460− 21465. (9) Bujak, M.; Podsiadło, M.; Katrusiak, A. Energetics of Conformational Conversion Between 1,1,2-Trichloroethane Polymorphs. Chem. Commun. 2008, 37, 4439−4441. (10) Bujak, M.; Blaser, D.; Katrusiak, A.; Boese, R. Conformational Polymorphs of 1,1,2,2-Tetrachloroethane: Pressure vs. Temperature. Chem. Commun. 2011, 47, 8769−8771. (11) Dziubek, K.; Katrusiak, A. Pressure-Induced Pseudorotation in Crystalline Pyrrolidine. Phys. Chem. Chem. Phys. 2011, 13, 15428− 15431. (12) Fabbiani, F. P. A.; Pulham, C. R. High-Pressure Studies of Pharmaceutical Compounds and Energetic Materials. Chem. Soc. Rev. 2006, 35, 932−942. (13) Ridout, J.; Probert, M. R. Low-Temperature and High-Pressure Polymorphs of Isopropyl Alcohol. CrystEngComm 2014, 16, 7397− 7400. (14) Patyk-Kaźmierczak, E.; Warren, M. R.; Allan, D. R.; Katrusiak, A. Intermolecular Contacts in Compressed α-d-Mannose. Cryst. Growth Des. 2016, 16, 6885−6890. (15) Boldyreva, E. V. High-Pressure Studies of the Anisotropy of Structural Distortion of Molecular Crystals. J. Mol. Struct. 2003, 647, 159−179. (16) Fabbiani, F. P. A.; Allan, D. R.; Dawson, A.; David, W. I. F.; McGregor, P. A.; Oswald, I. D. H.; Parsons, S.; Pulham, C. R. PressureInduced Formation of a Solvate of Paracetamol. Chem. Commun. 2003, 24, 3004−3005. (17) Fabbiani, F. P. A.; Buth, G.; Levendis, D. C.; Cruz-Cabeza, A. J. Pharmaceutical Hydrates under Ambient Conditions from HighPressure Seeds: a Case Study of GABA Monohydrate. Chem. Commun. 2014, 50, 1817−1819. (18) Zakharov, B. A.; Tumanov, N. A.; Boldyreva, E. V. β-Alanine under Pressure: Towards Understanding the Nature of Phase Transitions. CrystEngComm 2015, 17, 2074−2079. (19) Zakharov, B. A.; Seryotkin, Y. V.; Tumanov, N. A.; Paliwoda, D.; Hanfland, M.; Kumosov, A. V.; Boldyreva, E. V. The Role of Fluids in High-Pressure Polymorphism of Drugs: Different Behaviour of βChlorpropamide in Different Inert Gas and Liquid Media. RSC Adv. 2016, 6, 92629−92637. (20) Fedorov, A. Y.; Rychkov, D. A.; Losev, E. A.; Zakharov, B. A.; Stare, J.; Boldyreva, E. V. Effect of Pressure on Two Polymorphs of Tolazamide: Why No Interconversion? CrystEngComm 2017, 19, 2243−2252. 18835

DOI: 10.1021/acs.jpcc.7b06083 J. Phys. Chem. C 2017, 121, 18830−18836

Article

The Journal of Physical Chemistry C (41) Form, G. R.; Raper, E. S.; Downie, T. C. The Crystal and Molecular Structure of 2-Mercaptobenzimidazole. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1976, 32, 345−348. (42) Ravikumar, K.; Chandra Mohan, K.; Bidyasagar, M.; Swamy, G. Y. S. K. Crystal Structure of 2-Mercaptobenzimidazole and Bis[2Mercaptobenzimidazole]Dichlorocobalt(II). J. Chem. Crystallogr. 1995, 25, 325−329. (43) Merrill, L.; Bassett, W. A. Miniature Diamond Anvil Pressure Cell for Single Crystal X-Ray Diffraction Studies. Rev. Sci. Instrum. 1974, 45, 290−294. (44) Piermarini, G. J.; Block, S.; Barnett, J. D.; Forman, R. A. Calibration of the Pressure Dependence of the R1 Ruby Fluorescence Line to 195 kbar. J. Appl. Phys. 1975, 46, 2774−2780. (45) Mao, H.-K.; Xu, J.; Bell, P. M. Calibration of the Ruby Pressure Gauge to 800 kbar under Quasi-Hydrostatic Conditions. J. Geophys. Res. 1986, 91, 4673−4676. (46) Data collection and data reduction GUI. CrysAlisPro, version 1.171.33.61; Oxford Diffraction Ltd.: Wroclaw, Poland, 2014. (47) Budzianowski, A.; Katrusiak, A. High-Pressure Crystallographic Experiments with a CCD Detector. In High-Pressure Crystallography; Katrusiak, A., McMillan, P. F., Eds.; Kluwer: Dordrecht, 2004; pp 101−112. (48) Sheldrick, G. M. A Short History of SHELX. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112−122. (49) Macrae, C. F.; Bruno, I. J.; Chisholm, J. A.; Edgington, P. R.; McCabe, P.; Pidcock, E.; Rodriguez-Monge, L.; Taylor, R.; van de Streek, J.; Wood, P. A. Mercury CSD 2.0 - New Features for the Visualization and Investigation of Crystal Structures. J. Appl. Crystallogr. 2008, 41, 466−470. (50) Marciniak, J.; Andrzejewski, M.; Cai, W.; Katrusiak, A. Wallach’s Rule Enforced by Pressure in Mandelic Acid. J. Phys. Chem. C 2014, 118, 4309−4313. (51) Cai, W.; Marciniak, J.; Andrzejewski, M.; Katrusiak, A. Pressure Effect on DL-Mandelic Acid Racemate Crystallization. J. Phys. Chem. C 2013, 117, 7279−7285. (52) Patyk, E.; Katrusiak, A. Transformable H-bonds and Conformation in Compressed Glucose. Chem. Sci. 2015, 6, 1991− 1995. (53) Patyk, E.; Jenczak, A.; Katrusiak, A. Giant Strain Geared to Transformable H-bonded Network in Compressed β-D-Mannose. Phys. Chem. Chem. Phys. 2016, 18, 11474−11479. (54) Hazen, R. M.; Finger, L. W. Comparative Crystal Chemistry; John Wiley & Sons: New York, 1982; pp 180−186. (55) Hazen, R. M.; Finger, L. W. The Crystal Structures and Compressibilities of Layer Minerals at High Pressure. II. Phlogopite and Chlorite. Am. Mineral. 1978, 63, 293−296. (56) Levien, L.; Prewitt, C. T. High-Pressure Structural Study of Diopside. Am. Mineral. 1981, 66, 315−323. (57) Hazen, R. M.; Finger, L. W. Crystal Structures and Compressibilities of Pyrope and Grossular to 60 kbar. Am. Mineral. 1978, 63, 297−303. (58) Cai, W.; Gładysiak, A.; Anioła, M.; Smith, V. J.; Barbour, L. J.; Katrusiak, A. Giant Negative Area Compressibility Tunable in a Soft Porous Framework Material. J. Am. Chem. Soc. 2015, 137, 9296−9301. (59) Marciniak, J.; Katrusiak, A. Direct and Inverse Relations Between Temperature and Pressure Effects in Crystals: a Case Study on o-Xylene. J. Phys. Chem. C 2017, submitted for publication. (60) Cliffe, M. J.; Goodwin, A. L PASCal: a Principal Axis Strain Calculator for Thermal Expansion and Compressibility Determination. J. Appl. Crystallogr. 2012, 45, 1321−1329. (61) Allmann, R. In Homoatomic Rings, Chains and Macromolecules of the Main Group Elements; Rheingold, A., Ed.; Elsevier: Amsterdam, 1977; pp 25−58. (62) Kitajgorodski, A. I. Kryształy Molekularne; PWN: Warsaw, 1976; pp 168−171. (63) Kozakiewicz, A.; Ullrich, M.; Wełniak, M.; Wojtczak, A. Synthesis, Structure and Activity of Sulfonamides Derived from (+)-Camphor in the Enantioselective Addition of Diethylzinc to Benzaldehyde. J. Mol. Catal. A: Chem. 2010, 326, 128−140.

(64) Petzold, H.; Görls, H.; Weigand, W.; Romanski, J.; Mloston, G. Complexation of Cage Thiones with Bisphosphine Platinum(0) Complexes. Heteroat. Chem. 2007, 18, 584−590. (65) Brunelli, M.; Fitch, A. N.; Mora, A. J. Crystal Structures of RThiocamphor. Z. Kristallogr. - Cryst. Mater. 2002, 217, 83−87. (66) Shimada, K.; Nanae, T.; Aoyagi, S.; Takikawa, Y.; Kabuto, C. Regioselective Monohalogenation of 3,3-Disubstitued Bornane-2Thiones via Thione-Dihalogen Complexes. Tetrahedron Lett. 2001, 42, 6167−6169. (67) Kwiatkowki, W.; Cameron, T. S.; Salama, P.; Poirier, M. (1R,1R′)-2-Exo-Mercapto-2′-Thioxo-3-Exo,3′-Exo-Bibornane, 2-Dehydro-2,2′-Exo-Epidithio-3,3′-Bibornane and 2-Endo,2′-Exo-Epidithio3,3′-Bibornanylidene. Potential Antiviral Agents. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1997, 53, 387−391. (68) Rosenau, T.; Mereiter, K.; Jäger, C.; Schmid, P.; Kosma, P. Sulfonium Ylides Derived from 2-Hydroxy-Benzoquinones: Crystal and Molecular Structure and Their One-Step Conversion into Mannich Bases by Amine N-Oxides. Tetrahedron 2004, 60, 5719− 5723. (69) Schroth, W.; Hintzsche, E.; Spitzner, R.; Ströhl, D.; Schmeiβ, K.; Sieler, J. 3-Exo,3′-Exo-(1R,1′R)-Bithiocamphor − a Versatile Source for Functionally Different 3,3′-Bibornane Derivatives, II. 1 An Access to 3-Exo,3′-Exo-(1r,1′r)-Bicamphor and Related Compounds. Tetrahedron 1995, 51, 13261−13270. (70) Schroth, W.; Hintzsche, E.; Spitzner, R.; Ströhl, D.; Sieler, J. 3Exo,3′-Exo-(1R,1′R)-Bithiocamphor − a Versatile Source for Functionally Different 3,3′-Bibornane Derivatives − I. Ring-Closure Reactions and Prototropic Rearrangements. Tetrahedron 1995, 51, 13247−13260. (71) Sletten, J. The Crystal Structure of 2-(5′-Phenyl-1′,2′-Dithiole3′-Ylidene)-6-(5″-t-Butyl-1″,2″-Dithiole-3″-Ylidene)-Cyclohexanethione. An Extended No-bond Resonance System Comprising Five Sulphur Atoms. Acta Chem. Scand. 1970, 24, 1464−1466. (72) Sletten, J. Structures of Linear Multisulfur Systems. IX. The Crystal and Molecular Structure of 2,6-Bis(p-Methoxy-Phenyl-1,2Dithiole-3-Ylidene)Cyclohexanethione Carbon Disulfide Solvate, C26H22O2S5.1/2CS2. Acta Chem. Scand. 1975, 29, 436−442. (73) Khan, A. Z-Q.; Liao, F.-L.; Sandström, J.; Wang, S.-L. Twist Angles and Bond Lengths in Three Twisted Push-Pull Ethylenes. Interplay Between Steric and Electronic Effects. J. Chem. Soc., Perkin Trans. 2 1994, 1569−1573. (74) Allen, F. H.; Kennard, O.; Watson, D. G.; Brammer, L.; Orpen, G.; Taylor, R. Tables of Bond Lengths Determined by X-Ray and Neutron Diffraction. Part 1. Bond lengths in Organic Compounds. J. Chem. Soc., Perkin Trans. 2 1987, S1−S19. (75) Powell, B. M.; Dolling, G. Structure of Solid Carbon Disulphide Between 5 and 150 K. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1982, 38, 28−32. (76) Tomkowiak, H.; Olejniczak, A.; Katrusiak, A. PressureDependent Formation and Decomposition of Thiourea Hydrates. Cryst. Growth Des. 2013, 13, 121−125.

18836

DOI: 10.1021/acs.jpcc.7b06083 J. Phys. Chem. C 2017, 121, 18830−18836