Next Article in Journal
Metal–Dithiolene Bonding Contributions to Pyranopterin Molybdenum Enzyme Reactivity
Next Article in Special Issue
Rare Nuclearities in Ni(II) Cluster Chemistry: An Unprecedented {Ni12} Nanosized Cage from the Use of N-Naphthalidene-2-Amino-5-Chlorobenzoic Acid
Previous Article in Journal
A Review of the MSCA ITN ECOSTORE—Novel Complex Metal Hydrides for Efficient and Compact Storage of Renewable Energy as Hydrogen and Electricity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Copper(II) Halide Salts with 1-(4′-Pyridyl)-Pyridinediium

by
Jeffrey C. Monroe
1,*,
Christopher P. Landee
2,
Melanie Rademeyer
3 and
Mark M. Turnbull
1,*
1
Carlson School of Chemistry and Biochemistry, Clark University, 950 Main St., Worcester, MA 01610, USA
2
Department of Physics, Clark University, 950 Main St., Worcester, MA 01610, USA
3
Department of Chemistry, University of Pretoria, Private Bag X20, Hatfield 0028, South Africa
*
Authors to whom correspondence should be addressed.
Submission received: 10 December 2019 / Revised: 13 February 2020 / Accepted: 24 February 2020 / Published: 4 March 2020

Abstract

:
The compounds [1,4′-bipyridine]-1,1′-diium [CuCl4] (1) and [1,4′-bipyridine]-1,1′-diium [CuBr4] (2) were prepared and their crystal structures and magnetic properties are reported. The compounds are isomorphous and crystallize in the monoclinic space group C2/c. The cation crystallizes in a two-fold disordered fashion with the terminal nitrogen and carbon atoms exhibiting 50% occupancies. This results in a crystal packing arrangement with significant hydrogen bonding that is very similar to that observed in the corresponding 4,4′-bipyridinediium complexes. Temperature dependent magnetic susceptibility measurements and room temperature EPR spectroscopy indicate the presence of very weak antiferromagnetic exchange. The data were fit to the Curie–Weiss law and yielded Weiss constants of −0.26(5) K (1) and −1.0(1) K (2).

Graphical Abstract

1. Introduction

The importance of hydrogen bonding to crystal engineering has been noted in a number of recent studies [1,2,3]. The ability to predict or design an approximate packing structure using restricted avenues for strong hydrogen bonds is highly desirable in a variety of fields including crystal engineering [4], bioinorganic chemistry [5], and magnetism [6,7,8,9].
As described by Kumar et al. [10], the importance of the bifurcated CuCl2···H–N hydrogen bond is considered with dications such as 4,4′-bipyridinediium, which form ABAB ribbons where A is the CuCl4 dianion and B is the dication. A similar ribbon packing is observed for (4,4′-bipyridinediium)Cu2X6 (X = Cl, Br) [11]. The importance of the bifurcated CuCl2···H–N hydrogen bond was also shown by Yahsi et al. [12] with a variety of boronic acid cations for crystal engineering. Given our previous work with pyridinium salts of tetrahalidocuprate ions [13,14,15,16,17,18], we were intrigued with the potential for observing and using such hydrogen bonding in the formation of low-dimensional lattices for magnetic study. In many tetrahalidocuprate compounds [13,14,15,16,17,18,19,20,21,22,23,24], the general formula may be written as (BH)2[CuX4] where B is an organic base, most often an amine. The BH+ moiety provides charge balance and influences the structure of the crystal lattice through its steric and electronic properties. Several complexes have been prepared using dibasic organic species resulting in complexes of the general formula (BH2)[CuX4] [10,25,26,27].
Compounds with the doubly cationic [1,4′-bipyridine]-1,1′-diium are here described. While it was evident that the modification of the (BH)2[CuX4] model to the (BH)[CuX4] scaffold could result in less well-isolated layers, chains, ladders or honeycombs, due to the loss of one bulky cation per CuX42− unit, it was posited that some unique structures could be discovered, since the use of doubly cationic molecules is much less common. Thus, by exploring the use of [1,4′-bipyridine]-1,1′-diium (B1H), a sterically long, doubly cationic, and versatile counter-cation, it was proposed that design features stemming from the directionality of the hydrogen bond donors would present themselves. The compounds [1,4′-bipyridine]-1,1′-diium [CuCl4] (1), [1,4′-bipyridine]-1,1′-diium [CuBr4] (2) are described here regarding the effect of hydrogen bonding on crystal packing and the effect of B1H as a design agent in the synthesis of magnetic materials.

2. Results and Discussion

2.1. Synthesis

The synthesis of 1 and 2 may be achieved directly in the appropriate concentrated acid solution with stoichiometric ratios of the starting materials. The synthetic method is summarized in Scheme 1 for 1 and 2. Both compounds grow as flat plate crystals, with the expected colors for copper(II) halide salts (yellow, 1; purple, 2).

2.2. Structure

1 and 2 are isostructural with very similar lattice parameters, as shown in Table 1. Thus, in the following treatment, 2 is considered explicitly (Corresponding figures for 1 are shown in the Supplementary Materials as Figures S1 and S2).
The asymmetric unit of 2 contains half of a CuBr42− unit and half of a disordered B1H dication. The molecular unit of 2 is shown in Figure 1; the CuX4 distorted tetrahedra and full B1H dications are generated by 2-fold rotation axes running though Cu1 and the central bond of the B1H dication. The CuX42− unit in 2 may be comparably described as a flattened tetrahedron or a distorted square planar unit with a mean trans angle [28] in 1 = 144.9° and in 2 = 142.9° (the corresponding τ4 parameters are 0.53 (1) and 0.50 (2) [29]. The Cu–X bond lengths are those expected for a CuX42− unit [13,14,15,16,17,18,19,20,21,22,23,24]; selected bond lengths and angles are given in Table 2. The Jahn-Teller distortion characteristic of four-coordinate copper(II) compounds causes variation of the bond angles from purely tetrahedral. The very close similarity in the structures of 1 and 2 is easily seen in the overlay drawing (Figure S3).
The orientation of the organic group B1H is two-fold disordered in the lattice. Any given dication is hydrogen bonded to the CuBr42− unit via N11–H11. A superposition of the two orientations of the dication is shown in Figure 2 for 1. Atoms N11/C24 and N21/C14 occupy identical positions in the lattice and each have 0.5 occupancy as required by symmetry. Atoms C12, C13, C15 and C16 are not disordered and have full occupancies. Labels A, B and C indicate symmetry generated atoms. The fact that the organic moiety is in fact disordered in the lattice indicates that the orientation must indeed be random; if the orientation were to alternate is some fashion, the crystals could exhibit a different space group and a larger unit cell to accommodate the additional symmetry.
The N11–H11/C24–H24 hydrogen bond donors from B1H are involved in a bifurcated hydrogen bond to X1 and X2, with parameters shown below in Table 3 and shown in Figure 3. The symmetry equivalent disordered C24–H24 hydrogen bonding parameters are identical; however, the hydrogen bonding strength is presumed to be much weaker due to the lower polarity of the C–H bond. The angle between the normals to the symmetry equivalent pyridine rings in B1H is 45.9° (2) (48.2°, 1) as expected based on similar compounds [30,31,32]. Additional short C–H···Cl interactions are present in the structures (average dC···Cl~3.6 Å; average C–H···Cl~149° for 1). Complexes using B1H as the counterion have been previously reported including halido compounds of molybdenum and tungsten [31], the tetrachloridoplatinate and pentachloridostibnate salts [33] and a hexacyanidoferrate complex [34]. No structures involving tetrahedral metalate anions have been reported to our knowledge. The dihedral angles observed in the B1H cations are typically ~38° [30,31,33], significantly smaller than observed here, but in the ferrate species [34] that angle is 44°, comparable to what is observed for 1 and 2. None of the reported complexes exhibit the disorder observed for 1 and 2.
The crystal packing is such that the CuBr42− and B1H dications alternate in three dimensions as shown in Figure 4a (Figure S2 for 1). Due to the fact that there is only one B1H dication per CuBr42− unit, the CuBr42− units are close enough to each other to propagate weak magnetic interactions in three dimensions [1dCl···Cl = 4.36 Å, Cu-Cl···Cl = 119° and 109°, Cu–Cl···Cl–Cu = 71°; 2dBr···Br = 4.47 Å, Cu–Br···Br = 120° and 108°, Cu–Br···Br–Cu = 73°] [28]. It has been previously demonstrated in tetrahalidocuprate systems that the magnetic exchange is primarily dependent upon the halide···halide separation, rather than the distance between Cu(II) ions [35].
The crystal packing in 1 and 2 is very similar to that of (4,4′-bipyridinediium) [CuCl4] and 4,4′-(H2diazastilbene) [CuCl4] [10], each of which are composed of the ABAB ribbons with bifurcated hydrogen bonds controlling the packing pattern. The limited directionality of the hydrogen bond donors controls the packing of these compounds, thus enabling control of the 3D structure.

2.3. Magnetism

M(H) plots for 1 and 2 are shown in Figure 5. Comparison of the two plots indicates that the chloride analog exhibits weaker internal superexchange interactions as shown by the increased rate at which M(H) of 1 is nearing saturation at 50 kOe; 2 is still ~15% below saturation at that field. The internal interactions in compound 1 are sufficiently weak that the magnetization was ably fit to the Brillouin function, giving an average g-factor of 2.11.
The susceptibility versus temperature plots of 1 and 2 are void of structure and resemble the susceptibility of paramagnetic complexes. As demonstrated by the slight downturn in the χT(T) plots at low temperatures (Figure 6), there are very weak antiferromagnetic interactions. Given the weakness of the interactions, the 1/χ(T) data from 25 K to 310 K were fit to the Curie–Weiss model [36]. The fitted parameters for 1 and 2 are shown in Table 4 and indicate very weak coupling as expected. Weaker coupling in the chloride than the bromide complex is observed as expected of two-halide pathways of isostructural pairs of this type [28] and as expected from the M(H) data (vide supra).

2.4. Powder X-Band EPR

Despite the very weak nature of superexchange in these complexes, no hyperfine structure is observed in the powder EPR spectra. Both compounds 1 and 2 display axial g-tensors with largely Lorentzian line shapes as determined by fits using EasySpin [37]. The powder spectra are shown in Figure 7; the fitted g-values are g = 2.0594, g‖ = 2.3363 (gave = 2.1517) (1) and g = 2.0820, g‖ = 2.2361 (gave = 2.1334) (2). The broadening of g‖ is much larger in 1 (G-strain = 0.0523) than in 2 (G-strain = 0.0064), both because Δge is larger and because the interactions are much weaker, such that the effect of collapsed hyperfine coupling on the linewidth is larger. The average g-factor from EPR studies is considered more accurate than that from the fits to the magnetization and 1/χ(T).
The average g-factor found from the fit to the Brillouin function agrees within 2% of that found from EPR studies, and that from 1/χ(T) is within about 1% for 1. The average g-factor from the fit to 1/χ(T) for 2 is within 1.5% of that found from EPR studies. Similar to the findings of Farra et al. [38], g‖ for the bromide salt is much smaller than that of the chloride salt and the g is much larger for the bromide salt compared to the chloride salt, most likely due to the spin orbit coupling and ligand field effects of the bromide ion versus the chloride ion [39].

3. Materials and Methods

[1,4′-bipyridine]-1,1′-diium dichloride x-hydrate was purchased from TCI and was used as received except for the synthesis of 2 for which a halide exchange (vide infra) was performed. Copper(II) chloride dihydrate was purchased from Allied Chemical Corp. and copper(II) bromide was purchased from Alfa Aesar and used without further purification. Powder X-ray diffraction data were collected with a Bruker D8 Focus diffractometer. IR spectra were obtained with a Perkin Elmer Spectrum 100 FT-IR spectrometer via ATR. Elemental analyses were performed at the Marine Science Institute, University of California-Santa Barbara, CA, USA.

3.1. Synthesis

[1,4′-Bipyridine]-1,1′-diium tetrachloridocuprate (1): CuCl2⋅2H2O (0.341 g, 2.00 mmol) was dissolved in ca. 5 mL H2O, giving a blue solution. [1,4′-Bipyridine]-1,1′-diium dichloride x-hydrate (0.462 g, 2.02 mmol (calculation based on anhydrous form)) was dissolved in 5 mL H2O and while stirring 15 drops of concentrated HCl were added to this solution. The [1,4′-bipyridine]-1,1′-diium dichloride solution was combined with the CuCl2 solution giving a green solution. The solution was stirred for 5 more minutes and then left at room temperature, covered with filter paper. The following week, a mixture of blue and yellow crystals was isolated via vacuum filtration. The blue crystals were identified as [Cu(B1)2Cl2](Cl)2 (B1 = 1-(4′-pyridyl)pyridinium) [40]. The filtrate was acidified with additional drops of concentrated HCl. One month later, yellow crystals of 1 were isolated from the solution via vacuum filtration, washed with H2O and dried mechanically (0.3513 g, 48%). IR (cm−1) 3043 (mult, m), 1634 (w, sh), 1615 (s), 1528 (w), 1505 (w), 1488 (w), 1474 (s), 1355 (w), 1280 (w), 1002 (w), 806 (s), 771 (s), 760 (m, sh), 718 (s), 670 (vs), 550 (w). CHN Calcd.: C, 33.04; H, 2.77; N, 7.71; Found: C, 33.05; H, 2.79; N, 7.78.
[1,4′-Bipyridine]-1,1′-diium dibromide x-hydrate: [1,4′-bipyridine]-1,1′-diium dichloride x-hydrate (3.12 g) was added to a Schlenk flask with a stir bar and was dissolved in concentrated HBr (9M, 10 mL). A stream of air was passed over the solution until the solution evaporated to dryness. This process was repeated three times. The solid was recovered and stored in a desiccator. The crude product was used in the synthesis of 2 without further purification (Yield: 4.10 g, 94.8%) (anh). IR (cm−1): 3254 (br, m), 3049 (w), 3020 (m, mult), 2534 (v br, w), 1606 (m), 1505 (w), 1488 (w), 1468 (s), 1366 (w), 1239 (m), 1215 (m), 1168 (w), 1000 (m), 837 (s), 786 (s), 723 (m), 673 (s), 553 (w).
[1,4′-Bipyridine]-1,1′-diium tetrabromidocuprate (2): CuBr2 (0.110 g, 0.494 mmol) was dissolved in ca. 5 mL H2O. 1-(4′-pyridyl)-pyridinediium dibromide x-hydrate (0.162 g, 0.508 mmol) was dissolved in 3 mL 4.5 M HBr. The solutions were combined with stirring, giving a brownish solution. Two months later, purple crystals were isolated from solution via vacuum filtration, washed with methanol and dried mechanically (75 mg, 27%). IR (cm−1): 3220 (w), 3052 (med, mult), 1630 (w), 1611 (s), 1525 (w), 1503 (med), 1485 (med), 1472 (s), 1354 (w), 1311 (w), 1279 (w), 1197 (w), 950 (w), 796 (med), 763 (s), 715 (med), 668 (vs), 646 (med), 547 (med). CHN Calcd.: C, 22.19; H, 1.86; N, 5.17; Found: C, 22.21; H, 1.75; N, 5.06.

3.2. Magnetism

The magnetizations of 1 and 2 were collected on a Quantum Design MPMS SQUID magnetometer from 0 to 50 kOe at 1.8 K, with several data points collected from 50 kOe to 0 kOe to check for hysteresis. As expected of copper(II) complexes, none was observed. The magnetic susceptibilities of 1 and 2 were collected in a 1 kOe applied field from 1.8 K to 310 K. The data were corrected for the temperature independent paramagnetism of the Cu(II) ion and the diamagnetic moment of the constituent atoms as estimated from Pascals constants [41].

3.3. Crystal Structure Data Collection, Solution and Refinement

Single crystal X-ray diffraction data for 1 and 2 were collected at 150 K with Bruker APEX2 software (APEX2-2014, Bruker AXS Inc., Madison, WI, USA) employing φ and ω scans. The data reduction and refinement of the cell constants were performed with Bruker SAINT software [42]. Absorption corrections were made via SADABS [43]. The structures were solved using SHELXS-97 [44] and refined using SHELXL-2016 [45]. The disorder in the orientation of B1H in 1 and 2 was treated by constraining the anisotropic displacement parameters and positions of the disordered C and N atoms to be the same value/position (C14 and N21, and C24 and N11 occupy the same coordinates). The disorder occupancies were fixed at 50% to agree with the required symmetry in the final refinement; free refinement of the occupancies gave 53(2)% (1)/50(2)% (2). Details of the X-ray data collection and refinement are shown in Table 1. The data have been deposited with the CCDC: 1967063 (1), 1967066 (2).

4. Conclusions

The compounds (B1H) [CuX4] were synthesized and characterized via X-ray crystallography and magnetic studies, revealing the presence of very weak antiferromagnetic interactions in 3D between the CuX42− units which pack in an ABAB ribbon with the counter-dication B1H. Despite the loss of one strong N–H hydrogen bond donor in the ribbon, these structures resemble quite closely those of related compounds such as (4,4′-bipyridinediium) [CuCl4]. The disorder in the orientation of B1H may play an important role in the conservation of this ribbon structure. In future studies, the cation B1H can be modified to incorporate hydrogen bonding in different directions, modifying the crystal packing to achieve low dimensional magnetic systems with increased control. Further, the use of B1 as a monocationic ligand is under investigation.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2304-6740/8/3/18/s1, Structural drawings of compound 1 (molecular unit, Figure S1; layer structure, Figure S2), overlay structure of 1 and 2 (Figure S3).

Author Contributions

Investigation, J.C.M., C.P.L., M.R. and M.M.T.; Writing—original draft, J.C.M.; Writing—review and editing, M.M.T. All authors have read and agreed to the published version of the manuscript.

Funding

The APC was funded by the Carlson School of Chemistry and Biochemistry at Clark University, Worcester, MA, USA.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bu, R.; Xiong, Y.; Wei, X.; Li, H.; Zhang, C. Hydrogen Bonding in CHON-Containing Energetic Crystals: A Review. Cryst. Growth Des. 2019, 19, 5981–5997. [Google Scholar] [CrossRef]
  2. Saha, S.; Mishra, M.K.; Reddy, C.M.; Desiraju, G.R. From Molecules to Interactions to Crystal Engineering: Mechanical Properties of Organic Solids. Acc. Chem. Res. 2018, 51, 2957–2967. [Google Scholar] [CrossRef] [PubMed]
  3. Braga, D.; Grepioni, F.; Maini, L.; d’Agostino, S. From Solid-State Structure and Dynamics to Crystal Engineering. Eur. J. Inorg. Chem. 2018, 2018, 3597–3605. [Google Scholar] [CrossRef]
  4. Lago, A.B.N.; Carballo, R.; Garciía-Martiínez, E.; Vaázquez-Loópez, E.M. Metal to Ligand Interactions and Hydrogen Bonding in the Design of Metallosupramolecular Compounds: Effect of pH and Aprotic Solvents on the Nature of Materials Based on Bis(4-pyridylthio)methane and Copper(II) Chloride. Cryst. Growth Des. 2011, 11, 59–68. [Google Scholar] [CrossRef]
  5. Singh, Y.P.; Patel, R.N.; Singh, Y.; Choquesillo-Lazarte, D.; Butcher, R.J. Classical hydrogen bonding and stacking of chelate rings in new copper(II) complexes. Dalton Trans. 2017, 46, 2803–2820. [Google Scholar] [CrossRef] [PubMed]
  6. Okazawa, A.; Ishida, T. Super–superexchange coupling through a hydrogen bond in a linear copper(II) complex, [Cu(LH)(L)]·BF4·2H2O (LH = N-tert-butyl-N-2-pyridylhydroxylamine). Chem. Phys. Lett. 2009, 480, 198–202. [Google Scholar] [CrossRef]
  7. Manson, J.L.; Schlueter, J.A.; Funk, K.A.; Southerland, H.I.; Twamley, B.; Lancaster, T.; Blundell, S.J.; Baker, P.J.; Pratt, F.L.; Singleton, J.; et al. Strong H···F Hydrogen Bonds as Synthons in Polymeric Quantum Magnets: Structural, Magnetic, and Theoretical Characterization of [Cu(HF2)(pyrazine)2]SbF6, [Cu2F(HF)(HF2)(pyrazine)](SbF6)2, and [CuAg(H3F4)(pyrazine)5](SbF6)2. J. Am. Chem. Soc. 2009, 131, 6733–6747. [Google Scholar] [CrossRef]
  8. Tang, J.; Costa, J.S.; Golobic, A.; Kozlevcar, B.; Robertazzi, A.; Vargiu, A.V.; Gamez, P.; Reedijk, J. Magnetic coupling between copper(II) ions mediated by hydrogen-bonded (neutral) water molecules. Inorg. Chem. 2009, 48, 5473–5479. [Google Scholar] [CrossRef]
  9. Goodgame, D.M.L.; Grachvogel, D.A.; Williams, D.J. Combination of weak covalent, hydrogen bonding, and Ag⋯π interactions in the formation of a 3D porous network Ag(pypd)2(ClO4)·CH2Cl2 by the ambidentate ligand 1-(4′-pyridyl)pyridin-4-one. J. Chem. Soc. Dalton Trans. 2002, 11, 2259–2260. [Google Scholar] [CrossRef]
  10. Kumar, D.K.; Ballabh, A.; Jose, D.A.; Dastidar, P.; Das, A. How Robust Is the N–H···Cl2-Cu Synthon? Crystal Structures of Some Perchlorocuprates. Crys. Growth Des. 2005, 5, 651–660. [Google Scholar]
  11. Willett, R.D.; Butcher, R.E.; Landee, C.P.; Twamley, B. Two halide exchange in copper(II) halide dimers: (4,4′-Bipyridinium)Cu2Cl6xBrx. Polyhedron 2006, 25, 2093–2100. [Google Scholar] [CrossRef]
  12. Yahsi, Y.; Gungor, E.; Kara, H. Chlorometallate-Pyridinium Boronic Acid Salts for Crystal Engineering: Synthesis of One-, Two-, and Three-Dimensional Hydrogen Bond Networks. Cryst. Growth Des. 2015, 15, 2652–2660. [Google Scholar] [CrossRef]
  13. Dgachi, S.; Ben Salah, A.M.; Turnbull, M.M.; Bataille, T.; Naili, H. Investigations on (C6H9N2)2[MIIBr4] halogenometallate complexes with MII = Co, Cu and Zn: Crystal structure, thermal behavior and magnetic properties. J. Alloys Comps. 2017, 726, 315–322. [Google Scholar] [CrossRef]
  14. Solomon, B.L.; Landee, C.P.; Turnbull, M.M.; Wikaira, J.L. Copper(II) complexes of 2-amino-5-chloro-3-fluoropyridine: Syntheses and magnetic properties of (3,5-FCAP)2CuX2 and (3,5-FCAPH)2CuX4. J. Coord. Chem. 2014, 67, 3953–3971. [Google Scholar] [CrossRef]
  15. Hong, T.; Gvasaliya, S.N.; Herringer, S.N.; Turnbull, M.M.; Landee, C.P.; Regnault, L.-P.; Boehm, M.; Zheludev, A. Dynamics of the two-dimensional S = 1/2 dimer system (C5H6N2F)2CuCl4. Phy. Rev. B 2011, 83, 052401. [Google Scholar] [CrossRef] [Green Version]
  16. Tremelling, G.W.; Foxman, B.M.; Landee, C.P.; Turnbull, M.M.; Willett, R.D. Transition metal complexes of 2-amino-3,5-dihalopyridines: Syntheses, structures and magnetic properties of (3,5-diCAPH)2CuX4 and (3,5-diBAPH)2CuX4. Dalton Trans. 2009, 47, 10518–10526. [Google Scholar] [CrossRef]
  17. Deumal, M.; Giorgi, G.; Robb, M.A.; Turnbull, M.M.; Landee, C.P.; Novoa, J.J. The mechanism of magnetic interaction in spin-ladder molecular magnets: A first-principles, bottom-up, theoretical study of the magnetism in the two-legged spin-ladder bis(2-amino-5-nitropyridinium) tetrabromocuprate monohydrate. Eur. J. Inorg. Chem. 2005, 2005, 4697–4706. [Google Scholar] [CrossRef]
  18. Woodward, F.M.; Albrecht, A.S.; Wynn, C.M.; Landee, C.P.; Turnbull, M.M. Two-dimensional S = 1/2 Heisenberg antiferromagnets: Synthesis, structure, and magnetic properties. Phys. Rev. B 2002, 65, 144412. [Google Scholar] [CrossRef] [Green Version]
  19. Al Damen, M.A.; Haddad, S.F. The nonclassical noncovalent interactions control: A case study of the crystal structure of 3,5-dibromo-2-amino-4,6-dimethylpyridinium tetrahalocuprate [3,5-DBr-2-A-4,6-DMPH]2CuX4 (X = Cl, and Br). J. Mol. Struct. 2011, 985, 27–33. [Google Scholar] [CrossRef]
  20. Gong, J.; Chen, G.; Ni, S.F.; Zhang, Y.Y.; Wang, H.B. Bis(2-amino-6-methylpyridinium) tetrachloridocuprate(II). Acta Crystallogr. Sect. E 2009, 65, m1661–m1662. [Google Scholar] [CrossRef] [Green Version]
  21. Kovalchukova, O.V.; Palkina, K.K.; Strashnova, S.B.; Zaitsev, B.E. Synthesis, structure, geometrical, and spectral characteristics of the (HLn)2[CuCl4] complexes. Crystal and molecular structure of bis(2-methylimidazolium) tetrachlorocuprate(II). Russ. J. Coord. Chem. 2008, 34, 830–835. [Google Scholar] [CrossRef]
  22. Haddad, S.F.; AlDamen, M.A.; Willett, R.D. The role of non-classical supramolecular interactions in the structures of 2-amino-4,6-dimethylpyridinium tetrahalocuprate(II) salts. Inorg. Chim. Acta 2006, 359, 424–432. [Google Scholar] [CrossRef]
  23. Boutchard, C.L.; Hitchman, M.A.; Skelton, B.W.; White, A.H. Crystal structure and electronic spectrum of the distorted tetrahedral CuCl42− ion in bis(2-aminopyridinium) tetrachlorocuprate(II) and comparison with the tetragonally elongated octahedral complex bis(2-aminopyrimidinium)tetrachlorocopper(II). Aust. J. Chem. 1995, 48, 771–781. [Google Scholar] [CrossRef]
  24. Place, H.; Willett, R.D. Structure of catalytically related species involving copper(II) halides. III. 2-Amino-5-bromo-3-methylpyridinium 2-amino-3-methylpyridinium tetrabromocuprate(II). Acta Crystallogr. Sect. C 1987, 43, 1497–1500. [Google Scholar] [CrossRef] [Green Version]
  25. Walha, S.; Naili, H.; Yahyaoui, S.; Ali, B.F.; Turnbull, M.M.; Mhiri, T.; Ng, S.W. Three-Dimensional Network Polymeric Structure of (2-Amino-5-ammoniopyridinium)[CuCl4]: Crystal Supramolecularity and Magnetic Properties. J. Supercond. Nov. Magn. 2013, 26, 437–442. [Google Scholar] [CrossRef]
  26. Lee, Y.-M.; Park, S.-M.; Kang, S.K.; Kim, Y.-I.; Choi, S.-N. Crystal Structures and Magnetic Properties of Sparteinium Tetrahalocuprate Monohydrate Compounds. Bull. Kor. Chem. Soc. 2004, 25, 823–828. [Google Scholar]
  27. Wikaira, J.L.; Butcher, R.J.; Kersen, Ü.; Turnbull, M.M. Magnetic properties of 4,4′-bipiperidinediium tetrahalocuprates: Crystal structures of 4,4′-bipiperidinediium tetrabromocuprate(II) monohydrate and bis(4,4′-bipiperidinediium) hexabromodicuprate(I). J. Coord. Chem. 2015, 69, 57–71. [Google Scholar] [CrossRef]
  28. Turnbull, M.M.; Landee, C.P.; Wells, B.M. Magnetic exchange interactions in tetrabromocuprate compounds. Coord. Chem. Rev. 2005, 249, 2567–2576. [Google Scholar] [CrossRef]
  29. Yang, L.; Powell, D.R.; Houser, R.P. Structural variation in copper(I) complexes with pyridylmethylamide ligands: Structural analysis with a new four-coordinate geometry index, τ4. Dalton Trans. 2007, 9, 955–964. [Google Scholar] [CrossRef]
  30. Brenčič, J.V.; Čeh, B.; Leban, I. Structure of 1-(4-pyridyl)pyridinium trans-tetrachlorodi(pyridine)molybdate(III). Acta Crystallogr. Sect. C 1989, 45, 1144–1146. [Google Scholar] [CrossRef]
  31. Brenčič, J.V.; Čeh, B.; Leban, I. Preparation, Chemical, and Structural Characterization of Some Molybdenum(III) and Tungsten(III) Coordination Compounds Containing 1-(4-pyridyl)pyridinium Cation. Z. Anorg. Allg. Chem. 1988, 565, 163–170. [Google Scholar] [CrossRef]
  32. Goodgame, D.M.L.; Grachvogel, D.A.; White, A.J.P.; Williams, D.J. Diverse polymeric metal complexes formed by the ambidentate ligand 1-(4′-pyridyl)pyridin-4-one. Inorg. Chim. Acta 2003, 348, 187–193. [Google Scholar] [CrossRef]
  33. Angeloni, A.; Crawford, P.C.; Orpen, A.G.; Podesta, T.J.; Shore, B.J. Does Hydrogen Bonding Matter in Crystal Engineering? Crystal Structures of Salts of Isomeric Ions. Chem. Eur. J. 2004, 10, 3783–3791. [Google Scholar] [CrossRef] [PubMed]
  34. Gorelsky, S.I.; Ilyukhin, A.B.; Kholin, P.V.; Kotov, V.Y.; Lokshin, B.V.; Sapoletova, N.V. Dihydrohexacyanoferrates of N-heterocyclic cations. Inorg. Chim. Acta 2007, 360, 2573–2582. [Google Scholar] [CrossRef]
  35. Li, L.; Turnbull, M.M.; Landee, C.P.; Jornet, J.; Deumal, M.; Novoa, J.J.; Wikaira, J.L. Synthesis, structure and magnetic behavior of bis(2-amino-5-fluoropyridinium) tetrachlorocuprate(II). Inorg. Chem. 2007, 46, 11254–11265. [Google Scholar] [CrossRef]
  36. Landee, C.P.; Turnbull, M.M. A gentle introduction to magnetism: Units, fields, theory, and experiment. J. Coord. Chem. 2014, 67, 375–439. [Google Scholar] [CrossRef]
  37. Stoll, S.; Schweiger, A. EasySpin, a comprehensive software package for spectral simulation and analysis in EPR. J. Magn. Reson. 2006, 178, 42–55. [Google Scholar] [CrossRef]
  38. Farra, R.; Thiel, K.; Winter, A.; Klamroth, T.; Pöppl, A.; Kelling, A.; Schilde, U.; Taubert, A.; Strauch, P. Tetrahalidocuprates(II)—Structure and EPR spectroscopy. Part 1: Tetrabromidocuprates(II). New J. Chem. 2011, 35, 2793. [Google Scholar] [CrossRef] [Green Version]
  39. Willett, R.D.; Wong, R.J. Experimental Evidence for Phonon Modulation of Antisymmetric Exchange from the Temperature Dependence of EPR Linewidths in (RNH3)2CuX4 Salts. J. Magn. Reson. 1981, 42, 446–452. [Google Scholar] [CrossRef]
  40. Monroe, J.C.; Rademeyer, M.; Turnbull, M.M. Compounds of pyridylpyridones. In preparation.
  41. Carlin, R.L. Magnetochemistry; Springer: Berlin/Heidelberg, Germany, 1986. [Google Scholar]
  42. Bruker. SAINT-2012, Bruker AXS Inc.: Madison, Wisconsin, USA, 2012.
  43. Bruker SADABS. Version 2014/5, Bruker AXS Inc.: Madison, WI, USA.
  44. Sheldrick, G.M. A short history of SHELX. Acta Cryst. A 2008, 64, 112–122. [Google Scholar] [CrossRef] [Green Version]
  45. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, 71, 3–8. [Google Scholar]
Scheme 1. The preparation of 1 and 2 where B1H = [1,4′-bipyridine]-1,1′-diium and X = Cl, Br.
Scheme 1. The preparation of 1 and 2 where B1H = [1,4′-bipyridine]-1,1′-diium and X = Cl, Br.
Inorganics 08 00018 sch001
Figure 1. The molecular unit of 2, shown as 50% probability thermal ellipsoids; hydrogen atoms are shown as spheres of arbitrary size. Disordered positions (N11/C24 and N21/C14) are labelled with both atom names. Symm. Op. A = −x, y, −1/2 − z.
Figure 1. The molecular unit of 2, shown as 50% probability thermal ellipsoids; hydrogen atoms are shown as spheres of arbitrary size. Disordered positions (N11/C24 and N21/C14) are labelled with both atom names. Symm. Op. A = −x, y, −1/2 − z.
Inorganics 08 00018 g001
Figure 2. A superposition of the two disordered positions of the organic dication in 1.
Figure 2. A superposition of the two disordered positions of the organic dication in 1.
Inorganics 08 00018 g002
Figure 3. Br···H hydrogen bonding in 2. See Table 3 for hydrogen bonding data.
Figure 3. Br···H hydrogen bonding in 2. See Table 3 for hydrogen bonding data.
Inorganics 08 00018 g003
Figure 4. (a) Packing of B1H and CuBr4 units into a layer in 2, viewed parallel to the b-axis. (b) Packing of layers in 2, viewed parallel to the c-axis.
Figure 4. (a) Packing of B1H and CuBr4 units into a layer in 2, viewed parallel to the b-axis. (b) Packing of layers in 2, viewed parallel to the c-axis.
Inorganics 08 00018 g004
Figure 5. The magnetization of 1 (□) and 2 (○). The solid line is the fit to the S = ½ Brillouin function.
Figure 5. The magnetization of 1 (□) and 2 (○). The solid line is the fit to the S = ½ Brillouin function.
Inorganics 08 00018 g005
Figure 6. The magnetic susceptibility plotted as χT(T) (left axis) and 1/χ(T) (right axis) of 1 (top) and 2 (bottom).
Figure 6. The magnetic susceptibility plotted as χT(T) (left axis) and 1/χ(T) (right axis) of 1 (top) and 2 (bottom).
Inorganics 08 00018 g006
Figure 7. The X-band powder EPR spectra of 1 (a) and 2 (b) at room temperature with fits shown displaced from the data by −5 intensity units in red.
Figure 7. The X-band powder EPR spectra of 1 (a) and 2 (b) at room temperature with fits shown displaced from the data by −5 intensity units in red.
Inorganics 08 00018 g007
Table 1. Crystallographic information for 1 and 2.
Table 1. Crystallographic information for 1 and 2.
12
FormulaC10H10N2Cl4CuC10H10N2Br4Cu
Molecular Weight363.54541.38
Crystal Systemmonoclinicmonoclinic
Space GroupC2/cC2/c
a(Å)15.6790(11)16.0672(9)
b(Å)7.4315(5)7.5861(4)
c(Å)12.0742(8)12.5969(6)
α(°)9090
β(°)108.270(3)108.126(2)
γ(°)9090
V(Å3)1335.95(16)1459.21(13)
Z44
T(K)150(2)150(2)
ρcalc (g cm−1)1.8032.464
μ(mm−1)2.41112.429
λ(Å)0.710730.71073
Index Ranges−28 ≤ h ≤ 28−22 ≤ h ≤ 21
−13 ≤ k ≤ 13−10 ≤ k ≤ 10
−21 ≤ l ≤ 21−16 ≤ l ≤ 16
Indep. Reflections [I > 2σ(I)]30601473
Obs. reflections40221856
Parameters9495
Goodness of fit1.0211.063
R [I > 2σ(I)]0.04320.0363
Rw [I > 2σ(I)]0.08930.0795
R (all reflections)0.06410.0530
Rw (all reflections)0.09830.0850
Table 2. Selected bond lengths (Å) and angles (°) in 1 and 2. (Symm. Op. a = −x, y, ½ − z).
Table 2. Selected bond lengths (Å) and angles (°) in 1 and 2. (Symm. Op. a = −x, y, ½ − z).
12
Cu1–X12.2481(5)2.3816(5)
Cu1–X22.2515(5)2.3919(5)
X1–Cu1–X1a140.63(3)143.03(4)
X2–Cu1–X2a145.24(3)146.71(4)
X1–Cu1–X295.49(2)95.231(19)
X1–Cu1–X2a96.060(19)95.190(18)
X1a–Cu1–X296.060(18)95.189(18)
X1a–Cu1–X2a95.49(2)95.230(19)
Table 3. Hydrogen bonding parameters in 1 and 2.
Table 3. Hydrogen bonding parameters in 1 and 2.
DHA DH (Å)H···A (Å)D···A (Å) DHA (°)
N11–H11···X110.91(6)2.47(6)3.2699(18)147(5)
20.843(14)2.66(3)3.429(5)152(4)
N11–H11···X210.91(6)2.79(6)3.464(2)132(4)
20.843(14)2.99(4)3.599(5)131(4)
Table 4. Fitted parameters to the Curie–Weiss Law for 1 and 2. gave was calculated from the fitted value of the Curie constant (CC).
Table 4. Fitted parameters to the Curie–Weiss Law for 1 and 2. gave was calculated from the fitted value of the Curie constant (CC).
CC (emu·K/mol·Oe)θ (K)gaveR2
10.4406(1)−0.26(5)2.1680.99999
20.4157(4)−1.0(1)2.1050.99994

Share and Cite

MDPI and ACS Style

Monroe, J.C.; Landee, C.P.; Rademeyer, M.; Turnbull, M.M. Copper(II) Halide Salts with 1-(4′-Pyridyl)-Pyridinediium. Inorganics 2020, 8, 18. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8030018

AMA Style

Monroe JC, Landee CP, Rademeyer M, Turnbull MM. Copper(II) Halide Salts with 1-(4′-Pyridyl)-Pyridinediium. Inorganics. 2020; 8(3):18. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8030018

Chicago/Turabian Style

Monroe, Jeffrey C., Christopher P. Landee, Melanie Rademeyer, and Mark M. Turnbull. 2020. "Copper(II) Halide Salts with 1-(4′-Pyridyl)-Pyridinediium" Inorganics 8, no. 3: 18. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8030018

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop