Next Article in Journal
Amorphous Li-Al-Based Compounds: A Novel Approach for Designing High Performance Electrode Materials for Li-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

I62− Anion Composed of Two Asymmetric Triiodide Moieties: A Competition between Halogen and Hydrogen Bond

Institut für Anorganische Chemie und Strukturchemie, Lehrstuhl für Material- und Strukturforschung, Heinrich-Heine-Universität Düsseldorf, Universitätsstraße 1, 40225 Düsseldorf, Germany
*
Author to whom correspondence should be addressed.
Inorganics 2013, 1(1), 3-13; https://doi.org/10.3390/inorganics1010003
Submission received: 31 July 2013 / Revised: 25 September 2013 / Accepted: 21 October 2013 / Published: 31 October 2013

Abstract

:
The reaction of 1,8-diaminooctane with hydroiodic acid in the presence of iodine gave a new polyiodide-containing salt: 1,8-diaminiumoctane bis(triiodide), (H3N-(CH2)8-NH3)[I3]2. The title compound has been characterized by crystallographic and spectroscopic methods. The polyiodide ion is the first example of a hydrogen bonded I62− dianion consisting of two very asymmetric triiodide components with I−I distances of 2.7739(4) and 3.1778(4) Å interacting by a weak halogen bond (I···I: 3.5017(2) Å). The structural parameters of the triiodide anions, derived from X-ray crystallographic data, are in good agreement with the Raman and Far-IR spectroscopic results.

Graphical Abstract

1. Introduction

The term “polyiodide” defines an extended part of salt structures which fulfill the general formula I2m+nn (n = 1–4, m = integer) [1,2,3,4]. They are built from I-, I3- and I2 subunits and are known for their strong tendency to concatenate, forming extended motifs connected by halogen bonding [5,6,7,8,9]. Accurate relativistic correlated quantum chemical calculations of the halogen bond of the triiodide anion have been reported [10]. Some more complex polyiodides have been theoretically investigated in the recent years [11,12], and the general theory of halogen bonding has been improved [13,14,15]. There is also a general interest in diaminiumalkane iodide/triiodide systems derived from the knowledge that the choice of the diaminium cation significantly modifies the redox potential in dye-sensitized solar cells (DSSC) [16,17,18,19,20,21]. They play an important role in other electrochemical processes, e.g., lithium iodide batteries [22,23] or hydrogen production as well [24]. It is also well known that polyiodide chains, in the solid state, conduct the electric current by a Grotthuss-like mechanism [25,26]. Therefore the structures of chain-type polyiodides can be considered as a snapshot and model systems of this phenomenon. In the recent years several groups have synthesized different compounds using stick-shaped molecules and ions. There are several interesting classes of compounds in which the rod-shaped building blocks determine the shape and pore size of the framework [27,28]. The rod-shaped building blocks can be understood as tectons used for the crystal engineering of new functional materials [29,30]. It has been shown, especially for polyiodides, that tailored materials can be constructed, and that their structures fit with the lengths and shapes of the tectons used [31,32,33,34,35,36,37,38,39]. This selective and robust synthetic protocol for solid polyiodides using rod-shaped tectons or templates is now generally termed dimensional caging [31]. Salts constructed by α,ω-diaminiumalkanes represent an interesting class of organic-inorganic hybrid materials with a number of different structural arrangements including: layered materials [40]; hydrogen bonded frameworks as host systems for unusual species [41]; large-pore zeolites [42]; non-metal frameworks [43] and metal-frameworks [44,45]. We have already shown that the semi-flexible cationic α,ω-diaminiumalkanes are useful tectons for the construction of new polyiodides [46,47,48,49]. In this report we describe the structural and the spectroscopic properties of a halogen bonded I62− anion.

2. Results and Discussion

2.1. Crystal Structure

From the reaction of 1,8-diaminooctane with hydroiodic acid in the presence of a 1.5 molar excess of iodine dark-red crystals of 1,8-diaminiumoctane bis(triiodide) (18dbt) were obtained. A higher excess of iodine gave the title compound together with elemental iodine. Experiments with a lower excess of iodine gave a hitherto not completely characterized compound with the stoichiometry 1,8-diaminiumoctane/I = 1/3.5. The asymmetric unit of 18dbt consists of one half of a 1,8-diaminiumoctane dication located around a center of inversion and one triiodide anion in the centrosymmetric space group P-1. The cation shows an all-trans extended conformation with bond lengths and angles in the expected ranges (Table 1). In the 18dbt the rod-shaped 1,8-diaminiumoctane dications are connected to the polyiodide substructure via medium strong hydrogen bonds. Each hydrogen atom of the aminium group donates one hydrogen bond to one neighboring iodine atom. The N···I distances of 3.592(4), 3.653(5) and 3.688(4) Å indicate that the strengths of the bonds are similar. One iodine atom (I1) of the formal triiodide anion accepts three hydrogen bonds, whereas the two other iodine atoms are not involved in any hydrogen bonding (Figure 1; Table 2). The I–I distances of 2.7739(4) and 3.1778(4) Å in the I3 subunits are significantly more asymmetric than those in similar triiodide anions discussed so far [50,51], whereas the bond angle within these triiodide units shows typical 178°.
Figure 1. The I62− hydrogen bonded anion in the structure of 18dbt. (Symmetry codes: ' = −1 − x, 1 − y, −z; ’’ = 1 − x, 1 − y, 1 − z; ’’’ = −1 + x, y, z; displacement ellipsoids are drawn at the 70% probability level).
Figure 1. The I62− hydrogen bonded anion in the structure of 18dbt. (Symmetry codes: ' = −1 − x, 1 − y, −z; ’’ = 1 − x, 1 − y, 1 − z; ’’’ = −1 + x, y, z; displacement ellipsoids are drawn at the 70% probability level).
Inorganics 01 00003 g001
Table 1. Selected bond lengths [Å] and angles [°] for 18dbt.
Table 1. Selected bond lengths [Å] and angles [°] for 18dbt.
AtomsDistanceAtomsDistance
I1–I23.1778(4) I2–I32.7739(4)
I3–I3'3.5017(2)N1–C11.500(6)
C1–C21.516(6)C2–C31.537(7)
C3–C41.526(6)C4–C4''1.528(9)
I1–I2–I3178.34(2) I2–I3–I3' 163.36(2)
N1–C1–C2–C3−176.8(4)C1–C2–C3–C4179.0(4)
C2–C3–C4–C4''−177.5(5)--
' = −1−x, 1 − y, −z; '' = 2 − x, −y, 2 − z.
The asymmetry of this triatomic arrangement seems to be engendered in first instance by the electron drawing influence of the three hydrogen bonds accepted by the terminal iodine atom I1. Therefore, the structure could alternatively be understood as an inclusion compound with iodine molecules strongly coordinated to an electron-poor iodide anion. The distance between two neighboring triiodide anions in this structure is 3.5017(2) Å, which falls in the typical range of secondary I–I halogen interactions [5]. In contrast, the angle of 163° between the two units is unexpectedly acute. The arrangement found may be compared to that of two I3 ions (I–I distances within the I3 unit of 2.87 and 3.01 Å) trapped in a van der Waals hole of a hydrogen-bonded framework, which are separated by 3.63 Å [33]. Hence, the arrangement of two very asymmetric triiodide ions may be understood as the second, in this case hydrogen bonded, example of an I62− anion.
Table 2. Hydrogen bond parameters for 18dbt [Å and °].
Table 2. Hydrogen bond parameters for 18dbt [Å and °].
D―H∙∙∙AD―HH∙∙∙AD∙∙∙A< (DHA)
N1―H12∙∙∙I10.885(8)2.727(13)3.592(4)166(4)
N1―H11∙∙∙I1'0.893(8)2.810(12)3.688(4)168(4)
N1―H13∙∙∙I1''0.891(8)2.93(3)3.653(5)139(4)
' = 1 − x, 1 − y, 1− z; '' = 1 + x, y, z.
Figure 2. The staircase-shaped hydrogen bonded motif in 18dbt. Each staircase-shaped unit (shadowed) is constructed by two cyclic elements with the graph-set descriptor R24(8). The polyiodides and the alkan-diyl units connect these hydrogen-bonded strands to a three-dimensional framework (symmetry code: ' = 2 − x, −y, 2 − z).
Figure 2. The staircase-shaped hydrogen bonded motif in 18dbt. Each staircase-shaped unit (shadowed) is constructed by two cyclic elements with the graph-set descriptor R24(8). The polyiodides and the alkan-diyl units connect these hydrogen-bonded strands to a three-dimensional framework (symmetry code: ' = 2 − x, −y, 2 − z).
Inorganics 01 00003 g002
The hydrogen bonds between the dications and the polyiodide anions form a three-dimensional framework. Following the typical design of α,ω-diaminiumalkane polyiodide salts [46,47,48,49], the rod-shaped building blocks are arranged approximately parallel to each other. Each hydrogen atom of the aminium group donates one un-bifurcated hydrogen bond to a neighboring iodine center (Figure 2). The structure shows the eight membered hydrogen-bonded ring-motif R24(8) [52] which is well known for diaminiumalkyl polyiodides [46,49]. These ring motifs form staircase-shaped one-dimensional hydrogen-bonded units along [100] (Figure 2). The polyiodides and the alkan-diyl units connect the hydrogen-bonded strands to a three-dimensional framework (Figure 3).
Figure 3. A view along [100] shows the crosslinking between the one dimensional, staircase-shaped hydrogen-bonded strands.
Figure 3. A view along [100] shows the crosslinking between the one dimensional, staircase-shaped hydrogen-bonded strands.
Inorganics 01 00003 g003

2.2. Spectrosopy

The Raman spectroscopic results (Supplementary Materials, Figure S3) are in good agreement with those of the X-ray crystal structure analysis. The very strong band at 165 cm−1 and a weak one at ~100 cm−1 are typical for extremely asymmetric triiodide anions [26,53,54]. The high/low ratio of these bands has been discussed as a quantitative measure of the triiodide anion’s asymmetry. Typical values for asymmetrical and symmetrical triiodides are ~1.5 and ~1.2, respectively [26,55]. For 18dbt a ratio of 1.6 was found, which excellently reproduces the crystallographic results of a very asymmetric triiodide anion. The far-IR spectrum of 18dbt (Supplementary Materials, Figure S1) has the disadvantage that most bands are broad. But the spectrum independently verifies the presence and the position of a characteristic strong band at 165 cm−1.

3. Experimental

Synthesis: All chemicals were obtained from commercial sources and used as purchased. In a typical reaction 0.25 g (1.6 mmol) of 1,8-diaminooctane (Sigma-Alrich, 98%, Steinheim, Germany) and 0.61 g (2.4 mmol) of iodine (subl. ≥99%) were mixed, then stirred in 10 mL of aqueous hydroiodic acid (Sigma-Alrich, ≥57%, Steinheim, Germany), and heated up to ca. 370 K yielding a dark-red solution. Upon slow cooling to room temperature, dark-red, shiny crystals were formed at the bottom of the reaction vessel within one - two days.
General considerations: Mid-IR spectra of a powder sample were measured at room temperature on an Excalibur FTS 3500 spectrometer (Digilab, Germany) with an apodized resolution of 2 cm−1 using a MIRacle ATR unit (Pike Technologies, Madison, WI, USA) in the region of 4000–530 cm‑1. The far-IR spectra were measured on an IFS 113 V spectrometer (Bruker Optics, Leipzig, Germany) with an apodized resolution of 2 cm‑1 in the range of 500–25 cm−1 using polyethylene pellets. Raman spectra of powder samples were recorded on a MultiRam spectrometer (Bruker Optics, Leipzig, Germany) with an apodized resolution of 8 cm−1 equipped with a Nd-YAG laser: 1064 nm and a RT-InGaAs-detector in a backscattering geometry (4000–70 cm−1, 128 scans, 10 mW). Elemental analyses (C, H, N) were performed with an Euro EA3000 instrument (HEKA-Tech, Wegberg, Germany). Elemental analysis of iodine: A sample of 100 mg of 18dbt was dissolved in 15 mL of a water/acetone (10:1) mixture. After adding some drops of acetic acid and heating up to approximately 370 K zinc powder was added until the solution turned colorless. Filtering off the surplus of zinc yielded a clear solution which was analyzed by a classical precipitation titration (AgNO3 solution, 0.1 mol/L; potentiometric endpoint; Ag // Ag/AgCl electrodes) [56]. The melting point was determined using the Mettler-Toledo melting point system MP90 (range: 300–500 K, rate: 2 K/min).
Analytical data: Melting point: 446 K. Elemental analysis found: C, 10.7; H, 2.5; N, 3.4.; I, 83.3 Calc. for C8H22N2I6: C, 10.6; H, 2.4; N, 3.1; I, 83.9%.
Spectroscopic data: The Far-IR, Mid-IR and Raman spectra are shown in the Supplementary Materials section.
Crystal structure determination: The thin platelet crystals (plate face: {100}) were grown from the mother liquor of 18dbt. Several specimens of these crystals, that seemed to be suitable for X-ray diffraction, were harvested from the bulk samples. The crystals were mounted on an Oxford Xcalibur four-circle diffractometer [57] equipped with an EOS detector and a CryoJet cooling device (T = 100 K). In all cases the investigated specimens were non-merohedrally twinned. Data collection was performed with a crystal to detector distance of 45 mm and an omega-scan width of 1°/frame collecting a complete dataset up to 2θ = 60°. Indexing and determination of the twin law (−1.00, 0.00, 0.00, −0.24, 0.97, 0.07, −0.12, 0.97, −0.97 which is approximately a 180° rotation about −0.157, 0.981, 0.111 in direct space) followed the standard routine verified by a visual inspection via a reciprocal space viewer. Absorption correction was applied using indexed faces [57]. Structure solution by Direct Methods [58] using the data of one separated component always yielded all iodine atom positions. Secondary structure solution and simultaneous successive refinement of the primary structure using the SHELX program [58] yielded an almost complete model. As the refinement had reached an advanced stage, the method of Pratt, Coyle and Ibers [59] was used to process a dataset including both twin components (final twin ratio: 0.4755(6)/0.5245(6)). The hydrogen atoms were included at the latest stages of the refinement with restrained N-H and C-H distances. For the Uiso values of the NH3 group’s hydrogen atoms one common value was refined, whereas the Uiso values of hydrogen atoms attached to carbon atoms were calculated (Uiso = 1.2Ueq(C)). Selected crystallographic data are listed in Table 3.
Table 3. Crystallographic data for the title structure.
Table 3. Crystallographic data for the title structure.
Empirical formulaC8H22I6N2
Formula weight907.68
T/K 100(1)
Crystal system / Space grouptriclinic/P-1
a (Å)6.20008(14)
b (Å)7.0452(2)
c (Å)12.5770(3)
α (°)76.531(2)
β (°)78.842(2)
γ (°)84.574(2)
V3)/Z523.46(2)/1
μ (mm−1)8.888
Absorption correctionanalytical
Tmin./Tmax.0.102/0.850
Reflections*/reflections (I > 2σ(I))5489/5185
θ range (°)2.98–28.99
R(F)[I > 2σ(I)]/wR(F2) [all data]0.0341/0.0888
Parameters / restraints / completeness87/9/100%
Δρmax./Δρmin (e·Å−3 )1.530/−1.058
CCDC-number864870
* unmerged twin refinement.

4. Conclusions

This contribution is part of the recent efforts of several groups to synthesize new polyiodide containing hybrid materials and to understand the principles of caging small polyiodide anions [31,32,33,34,36,37,38] and iodine molecules [60,61,62]. The hydrogen bonded framework of 1,8-diaminiumoctane bis(triiodide), (H3N-(CH2)8-NH3)[I3]2, is constructed by one-dimensional staircase-shaped motifs which are connected with each other by polyiodide and alkan-diyl units. If we played the role of ‘devil’s advocate’ the structure could also be described as a hydrogen bonded framework of dications and iodide anions featuring oblong voids each filled with two iodine molecules. But as Lee et al. [63] have shown by an extensive database analysis, besides the stronger I−I bonds typical for I3- and I2 subunits also long-range interactions are reasonable. The polyiodide substructure may therefore be considered as a hydrogen bonded I62- anion consisting of two triiodide anions interacting by a weak halogen bond. The approximately linear triiodide subunit exhibits an extreme asymmetry. The structural properties of the polyiodide moiety have been verified by Raman and Far-IR spectroscopy.

Acknowledgments

We thank E. Hammes, P. Roloff and R. Linder for technical support. Furthermore, we would like to thank M. Bujak for useful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tebbe, K.-F. Polyhalogen Cations and Polyhalide Anions. In Homoatomic Rings, Chains and Macromolecules of Main-Group Elements, 1st ed.; Rheingold, A.L., Ed.; Elsevier: Amsterdam, The Netherlands, 1977; pp. 551–606. [Google Scholar]
  2. Coppens, P. Structural Aspects of Iodine-Containing Low-Dimensional Materials. In Extended Linear Chain Compounds; Miller, J.S., Ed.; Plenum Press: New York, NY, USA, 1982; Volume 1, pp. 333–356. [Google Scholar]
  3. Blake, A.J.; Devillanova, F.A.; Gould, R.O.; Li, W.S.; Lippolis, V.; Parsons, S.; Radek, C.; Schröder, M. Template self-assembly of polyiodide networks. Chem. Soc. Rev. 1998, 27, 195–205. [Google Scholar]
  4. Svensson, P.H.; Kloo, L. Synthesis, structure, and bonding in polyiodide and metal iodide-iodine systems. Chem. Rev. 2003, 103, 1649–1684. [Google Scholar] [CrossRef]
  5. Kloo, L.; Rosdahl, J.; Svensson, P.H. On the intra- and intermolecular bonding in polyiodides. Eur. J. Inorg. Chem. 2002, 1203–1209. [Google Scholar]
  6. Metrangolo, P.; Meyer, F.; Pilati, T.; Resnati, G.; Terraneo, G. Halogen bonding in supramolecular chemistry. Angew. Chem., Int. Ed. 2008, 47, 6114–6127. [Google Scholar]
  7. Resnati, G.; Metrangolo, P. Halogen Bonding. In Encyclopedia of Supramolecular Chemistry; Atwood, J.L., Steed, J.W., Eds.; Taylor & Francis: Boca Raton, FL, USA, 2007; Volume 1, pp. 628–635. [Google Scholar]
  8. Hasell, T.; Schmidtmann, M.; Cooper, A.I. Molecular doping of porous organic cages. J. Am. Chem. Soc. 2011, 133, 14920–14923. [Google Scholar]
  9. Yin, Z.; Wang, Q.-X.; Zeng, M.-H. Iodine release and recovery, influence of polyiodide anions on electrical conductivity and nonlinear optical activity in an interdigitated and interpenetrated bipillared-bilayer metal–organic framework. J. Am. Chem. Soc. 2012, 134, 4857–4863. [Google Scholar] [CrossRef]
  10. Gomes, A.S.P.; Visscher, L.; Bolvin, H.; Saue, T.; Knecht, S.; Fleig, T.; Eliav, E. The electronic structure of the triiodide ion from relativistic correlated calculations: A comparison of different methodologies. J. Chem. Phys. 2010, 133, 064305. [Google Scholar]
  11. Groenewald, F.; Esterhuysen, C.; Dillen, J. Extensive theoretical investigation: Influence of the electrostatic environment on the I3-···I3-–anion–anion interaction. Theor. Chem. Acc. 2012, 131, 1–12. [Google Scholar]
  12. Otsuka, M.; Mori, H.; Kikuchi, H.; Takano, K. Density functional theory calculations of iodine cluster anions: Structures, chemical bonding nature, and vibrational spectra. Comput. Theor. Chem. 2011, 973, 69–75. [Google Scholar] [CrossRef]
  13. Clark, T.; Hennemann, M.; Murray, J.; Politzer, P. Halogen bonding: The σ-hole. J. Mol. Model. 2007, 13, 291–296. [Google Scholar] [CrossRef]
  14. Politzer, P.; Murray, J.S.; Clark, T. Halogen bonding: An electrostatically-driven highly directional noncovalent interaction. Phys. Chem. Chem. Phys. 2010, 12, 7748–7757. [Google Scholar] [CrossRef]
  15. Politzer, P.; Murray, J.S. Halogen bonding: An interim discussion. ChemPhysChem 2013, 14, 278–294. [Google Scholar] [CrossRef]
  16. O´Regan, B.; Grätzel, M. A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 1991, 353, 737–740. [Google Scholar] [CrossRef]
  17. Gorlov, S.; Kloo, L. Ionic liquid electrolytes for dye-sensitized solar cells. Dalton Trans. 2008, 2655–2666. [Google Scholar] [CrossRef]
  18. Jeon, S.; Jo, Y.; Kim, K.-J.; Jun, Y.; Han, C.-H. High performance dye-sensitized solar cells with alkylpyridinium iodide salts in electrolytes. Appl. Mater. Interfaces 2011, 3, 512–516. [Google Scholar] [CrossRef]
  19. Rowley, J.G.; Meyer, G.J. Di- and tri-iodide reactivity at illuminated titanium dioxide interfaces. J. Phys. Chem. C 2011, 115, 6156–6161. [Google Scholar] [CrossRef]
  20. Yang, Y.; Sun, R.; Shi, C.; Wu, Y.; Xia, M. Application of diaminium iodides in binary ionic liquid electrolytes for dye-sensitized solar cells. Int. J. Photoenergy 2011, 986869:1–986869:5. [Google Scholar]
  21. Thorsmølle, V.K.; Topgaard, D.; Brauer, J.C.; Zakeeruddin, S.M.; Lindman, B.; Grätzel, M.; Moser, J.-E. Conduction through viscoelastic phase in a redox-active ionic liquid at reduced temperatures. Adv. Mater. 2012, 24, 781–784. [Google Scholar]
  22. Weinstein, L.; Yourey, W.; Gural, J.; Amatucci, G.G. Electrochemical impedance spectroscopy of electrochemically self-assembled lithium-iodine batteries. J. Electrochem. Soc. 2008, 155, A590–A598. [Google Scholar] [CrossRef]
  23. Yourey, W.; Weinstein, L.; Amatucci, G.G. Structure and charge transport of polyiodide networks for electrolytically in-situ formed batteries. Solid State Ionics 2011, 204, 80–86. [Google Scholar]
  24. Tyagi, D.; Varma, S.; Bhattacharya, K.; Jain, D.; Tripathi, A.K.; Pillai, C.G.S.; Bharadwaj, S.R. Iodine speciation studies on Bunsen reaction of S–I cycle using spectroscopic techniques. Int. J. Hydrogen Energy 2012, 37, 3621–3625. [Google Scholar]
  25. Alvarez, S.; Novoa, J.; Mota, F. The mechanism of electrical conductivity along polyhalide chains. Chem. Phys. Lett. 1986, 132, 531–534. [Google Scholar] [CrossRef]
  26. Marks, T.J.; Kalina, D.W. Highly Conductive Halogenated Low-Dimensional Materials. In Extended Linear Chain Compounds; Miller, J.S., Ed.; Plenum Press: New York, NY, USA, 1982; Volume 1, pp. 197–331. [Google Scholar]
  27. Hosseini, M.W. Molecular tectonics: From molecular recognition of anions to molecular networks. Coord. Chem. Rev. 2003, 240, 157–166. [Google Scholar] [CrossRef]
  28. Aakeröy, C.B.; Champness, N.R.; Janiak, C. Recent advances in crystal engineering. CrystEngComm 2010, 12, 22–43. [Google Scholar] [CrossRef]
  29. Brunet, P.; Simard, M.; Wuest, J.D. Molecular tectonics. Porous hydrogen-bonded networks with unprecedented structural integrity. J. Am. Chem. Soc. 1997, 119, 2737–2738. [Google Scholar] [CrossRef]
  30. Resnati, G.; Metrangolo, P. Tectons: Definition and Scope. In Encyclopedia of Supramolecular Chemistry; Atwood, J.L., Steed, J.W., Eds.; Taylor & Francis: Boca Raton, FL, USA, 2007; Volume 1, pp. 1484–1492. [Google Scholar]
  31. Svensson, P.H.; Gorlov, M.; Kloo, L. Dimensional caging of polyiodides. Inorg. Chem. 2008, 47, 11464–11466. [Google Scholar] [CrossRef]
  32. Abate, A.; Brischetto, M.; Cavallo, G.; Lahtinen, M.; Metrangolo, P.; Pilati, T.; Radice, S.; Resnati, G.; Rissanen, K.; Terraneo, G. Dimensional encapsulation of I-∙∙∙I2∙∙∙I- in an organic salt crystal matrix. Chem. Commun. 2010, 46, 2724–2726. [Google Scholar] [CrossRef]
  33. Meyer, M.K.; Graf, J.; Reiss, G.J. Dimer oder nicht dimer, das ist hier die Frage: Zwei benachbarte I3--Ionen eingeschlossen in Hohlräumen einer komplexen Wirtsstruktur. Z. Naturforsch. B 2010, 65, 1462–1466. (In German) [Google Scholar]
  34. García, M.D.; Martí-Rujas, J.; Metrangolo, P.; Peinador, C.; Pilati, T.; Resnati, G.; Terraneo, G.; Ursini, M. Dimensional caging of polyiodides: Cation-templated synthesis using bipyridinium salts. CrystEngComm 2011, 13, 4411–4416. [Google Scholar] [CrossRef] [Green Version]
  35. Reiss, G.J.; van Megen, M. Two new polyiodides in the 4,4’-bipyridinium diiodide/iodine system. Z. Naturforsch. B 2012, 67, 5–10. [Google Scholar]
  36. Van Megen, M.; Reiss, G.J. The pseudosymmetric structure of bis(pentane-1,5-diaminium) iodide tris(triiodide). Acta Cryst. E 2012, 68, o1331–o1332. [Google Scholar] [CrossRef]
  37. Müller, M.; Albrecht, M.; Gossen, V.; Peters, T.; Hoffmann, A.; Raabe, G.; Valkonen, A.; Rissanen, K. Anion–π interactions in salts with polyhalide anions: Trapping of I42−. Chem. Eur. J. 2010, 16, 12446–12453. [Google Scholar] [CrossRef]
  38. Garzón-Tovar, L.; Duarte-Ruiz, Á.; Wurst, K. Non-classical hydrogen bond (CH∙∙∙I) directed self-assembly formation of a novel 1D supramolecular polymer, based on a copper complex [Cu{(CH3)2SO}6]I4. Inorg. Chem. Commun. 2013, 32, 64–67. [Google Scholar] [CrossRef]
  39. Redel, E.; Röhr, C.; Janiak, C. An inorganic starch–iodine model. J. Chem. Soc. Chem. Commun. 2009, 2103–2105. [Google Scholar]
  40. Takeoka, Y.; Fukasawa, M.; Matsui, T.; Kikuchi, K.; Rikukawa, M.; Sanui, K. Intercalated formation of two-dimensional and multi-layered perovskites in organic thin films. Chem. Commun. 2005, 378–380. [Google Scholar]
  41. Frank, W.; Reiss, G.J. Bis(1,6-diammoniohexane) tetraaquahydrogen(1+) hexachlororhodate(III) dichloride, [H3N(CH2)6NH3]2[H9O4][RhCl6]Cl2: Chain-like [H9O4]+ ions enclosed in the cavities of a complex organic-inorganic framework. Inorg. Chem. 1997, 36, 4593–4595. [Google Scholar] [CrossRef]
  42. Jiang, J.; Yu, J.; Corma, A. Extra-large-pore zeolites: Bridging the gap between micro and mesoporous structures. Angew. Chem. Int. Ed. 2010, 49, 3120–3145. [Google Scholar] [CrossRef]
  43. Vizi, M.Z.; Knobler, C.B.; Owen, J.J.; Khan, M.I.; Schubert, D.M. Structures of self-assembled nonmetal borates derived from α,ω-diaminoalkanes. Cryst. Growth Des. 2006, 6, 538–545. [Google Scholar] [CrossRef]
  44. Seidlhofer, B.; Pienack, N.; Bensch, W. Synthesis of inorganic-organic hybrid thiometallate materials with a special focus on thioantimonates and thiostannates and in situ X-Ray Scattering Studies of their Formation. Z. Naturforsch. B 2010, 65, 937–975. [Google Scholar]
  45. Lu, J.Y.; Schauss, V. A novel nanostructured open-channel coordination polymer with an included fused-polyiodide ring. Eur. J. Inorg. Chem. 2002, 1945–1947. [Google Scholar]
  46. Reiss, G.J.; van Megen, M. Synthesis, structure and spectroscopy of a new polyiodide in the α,ω-diazaniumalkane iodide/iodine system. Z. Naturforsch. B 2012, 67, 447–451. [Google Scholar]
  47. Reiss, G.J. Personal communication to the Cambridge Structural Database. deposition number CCDC 786078. Cambridge, UK, 2010. [Google Scholar]
  48. Reiss, G.J.; Engel, J.S. Crystal Engineering of a new layered polyiodide using 1,9-diammoniononane as a flexible template cation. Z. Naturforsch. B 2004, 59, 1114–1117. [Google Scholar]
  49. Reiss, G.J.; Engel, J.S. Hydrogen bonded 1,10-diammoniodecane—An example of an organo-template for the crystal engineering of polymeric polyiodides. CrystEngComm 2002, 4, 155–161. [Google Scholar] [CrossRef]
  50. Rossi, M.; Marzilli, L.G.; Kistenmacher, T.J. Protonated 1-methylcytosine triiodide. Acta Cryst. B 1978, 34, 2030–2033. [Google Scholar] [CrossRef]
  51. Tebbe, K.-F.; Freckmann, B.; Horner, M.; Hiller, W.; Strähle, J. Verfeinerung der Kristallstruktur des Ammoniumtriiodids, NH4I3. Acta Cryst. C 1985, 41, 660–663. [Google Scholar] [CrossRef]
  52. Etter, M.C.; MacDonald, J.C.; Bernstein, J. Graph-set analysis of hydrogen-bond patterns in organic crystals. Acta Cryst. B 1990, 46, 256–262. [Google Scholar] [CrossRef]
  53. Deplano, P.; Ferraro, J.R.; Mercuri, M.L.; Trogu, E.F. Structural and Raman spectroscopic studies as complementary tools in elucidating the nature of the bonding in polyiodides and in donor-I2 adducts. Coord. Chem. Rev. 1999, 188, 71–95. [Google Scholar] [CrossRef]
  54. Ambrosetti, R.; Bellucci, G.; Bianchini, R.; Bigoli, F.; Deplano, P.; Pellinghelli, M.A.; Trogu, E.F. Kinetic and mechanistic study of the reaction between dithiomalonamides and diiodine. Crystal structure of the compounds 3,5-bis(ethylamino)-1,2-dithiolylium triiodide and 3,5-bis(morpholino)-1,2-dithiolylium iodide monohydrate. J. Chem. Soc., Perkin Trans. 1991, 2, 339–347. [Google Scholar]
  55. Kalina, D.W.; Lyding, J.W.; Ratajack, M.T.; Kannewurf, C.R.; Marks, T.J. Bromine as a partial oxidant. Oxidation state and charge transport in brominated nickel and palladium bis(diphenylglyoximates). A comparison with the iodinated materials and resonance Raman structure-spectra correlations for polybromides. J. Am. Chem. Soc. 1980, 102, 7854–7862. [Google Scholar]
  56. Egli, R.A. Neue Bestimmungsmethode für organisch gebundenes Iod, Brom und Chlor. Z. Anal. Chem. 1969, 247, 39–41. [Google Scholar] [CrossRef]
  57. CrysAlisPro, version 1.171.33.42; Oxford Diffraction Ltd.: Oxford, UK, 2009.
  58. Sheldrick, G.M. A short history of SHELX. Acta Cryst. A 2008, 64, 112–122. [Google Scholar] [CrossRef]
  59. Pratt, B.C.; Coyle, J.A.; Ibers, J. Redetermination of the structure of nitrosylpenta-amminecobalt(III) dichloride. J. Chem. Soc. A 1971, 2146–2151. [Google Scholar] [CrossRef]
  60. Peuronen, A.; Valkonen, A.; Kortelainen, M.; Rissanen, K.; Lahtinen, M. Halogen bonding-based “catch and release”: Reversible solid-state entrapment of elemental iodine with monoalkylated dabco salts. Cryst. Growth Des. 2012, 12, 4157–4169. [Google Scholar] [CrossRef]
  61. Zeng, M.-H.; Wang, Q.-X.; Tan, Y.-X.; Hu, S.; Zhao, H.-X.; Long, L.-S.; Kurmoo, M. Rigid pillars and double walls in a porous metal-organic framework: Single-crystal to single-crystal, controlled uptake and release of iodine and electrical conductivity. J. Am. Chem. Soc. 2010, 132, 2561–2563. [Google Scholar]
  62. Hu, J.; Wang, D.; Guo, W.; Du, S.; Tang, Z.K. Reversible control of the orientation of iodine molecules inside the AlPO4–11 crystals. J. Phys. Chem. C 2012, 116, 4423–4430. [Google Scholar]
  63. Lee, S.; Chen, B.; Fredrickson, D.C.; DiSalvo, F.J.; Lobkovsky, E.; Adams, J.A. Crystal structures of (pyrene)10(i3)4(i2)10 and [1,3,6,8-tetrakis(methylthio)pyrene]3(I3)3(I2)7: structural trends in fused aromatic polyiodides. Chem. Mater. 2003, 15, 1420–1433. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Van Megen, M.; Reiss, G.J. I62− Anion Composed of Two Asymmetric Triiodide Moieties: A Competition between Halogen and Hydrogen Bond. Inorganics 2013, 1, 3-13. https://doi.org/10.3390/inorganics1010003

AMA Style

Van Megen M, Reiss GJ. I62− Anion Composed of Two Asymmetric Triiodide Moieties: A Competition between Halogen and Hydrogen Bond. Inorganics. 2013; 1(1):3-13. https://doi.org/10.3390/inorganics1010003

Chicago/Turabian Style

Van Megen, Martin, and Guido J. Reiss. 2013. "I62− Anion Composed of Two Asymmetric Triiodide Moieties: A Competition between Halogen and Hydrogen Bond" Inorganics 1, no. 1: 3-13. https://doi.org/10.3390/inorganics1010003

Article Metrics

Back to TopTop