Next Article in Journal
Investigation of 1-Methylcytosine as a Ligand in Gold(III) Complexes: Synthesis and Protein Interactions
Previous Article in Journal
Imidazo-Phenanthroline Ligands as a Convenient Modular Platform for the Preparation of Heteroleptic Cu(I) Photosensitizers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Syntheses and Characterization of Two Dicyanamide Compounds Containing Monovalent Cations: Hg2[N(CN)2]2 and Tl[N(CN)2]

1
Chair of Solid-State and Quantum Chemistry, Institute of Inorganic Chemistry, RWTH Aachen University, 52056 Aachen, Germany
2
Max-Planck-Institut für Kohlenforschung, 45470 Mülheim an der Ruhr, Germany
*
Author to whom correspondence should be addressed.
Inorganics 2018, 6(4), 135; https://doi.org/10.3390/inorganics6040135
Submission received: 6 December 2018 / Revised: 12 December 2018 / Accepted: 13 December 2018 / Published: 18 December 2018
(This article belongs to the Section Inorganic Solid-State Chemistry)

Abstract

:
Crystals of Hg2[N(CN)2]2 were grown by a slow diffusion-reaction between aqueous Hg2(NO3)2·2H2O and Na[N(CN)2]. Hg2[N(CN)2]2 adopts the triclinic space group P 1 ¯ (no. 2) with a = 3.7089(5), b = 6.4098(6), c = 8.150(6) Å, α = 81.575(6)°, β = 80.379(7)°, γ = 80.195(7)°, and Z = 1. Crystals of Tl[N(CN)2] were obtained from the reaction of TlBr with Ag[N(CN)2] in water. Single-crystal structure analyses evidence that Tl[N(CN)2] is isotypic to α-K[N(CN)2] and adopts the orthorhombic space group Pbcm (no. 57) with a = 8.5770(17), b = 6.4756(13), c = 7.2306(14) Å, and Z = 4. Regarding volume chemistry, the dicyanamide anion occupies ca. 44 cm3·mol−1, and so it corresponds to a large pseudohalide. The IR spectra of both compounds exhibit vibrational modes that are characteristic of the dicyanamide anion.

Graphical Abstract

1. Introduction

Nitrogen-based solid-state materials have found extremely diverse applications, for example, as simple fertilizers, as high-performance steel coatings, as III-V optical semiconductors, and, quite recently, even as lithium- and sodium-ion battery materials. In particular, complex nitrogen containing compounds beyond the simple, yet fundamental nitrides, such as carbodiimides or guanidinates, are promising materials [1,2,3]. Similar to the latter in terms of chemical functionality, another interesting inorganic moiety is the boomerang-shaped dicyanamide anion [N(CN)2], which is often dubbed as [dca]. It is intriguing, since it possesses one doubly-coordinated and two singly-coordinated nitrogen atoms plus several electron lone pairs.
Generally speaking, the solubility of a metal dicyanamide in water is an important factor regarding the synthetic strategy of whatever binary or ternary phase is targeted. If the desired product is insoluble in water and can be filtered off after a metathesis reaction with Na[dca], therefore this method can be used to grow crystals via a diffusion reaction. If the target compound is water soluble, however, it can be synthesized employing a metathesis reaction with Ag[dca]. In this case, the driving force for the reaction is the subsequent formation of insoluble silver halides. After filtering to remove the silver halides, the product can then be crystallized by simply evaporating water. Binary dicyanamide compounds are known for ammonium [4], alkali metals [5,6,7,8] (except Fr), alkaline-earth metals [9] (except Be and Ra), transition metals (Cr–Zn, except Fe) [10,11,12,13,14,15,16], and rare-earth metals (La, Ce, Pr, Nd, Sm, Eu, Gd, Tb) [17,18,19]. Ternary compounds, such as KCs[dca]2 [20], LiK[dca]2 [21], LiRb[dca]2 [21], NaRb2[dca]3∙H2O [20], NaCs2[dca]3 [22], and LiCs2[dca]3 [23] are also known.
The synthesis of Hg2[dca]2 from Hg2(NO3)2 with Na[dca] has been reported almost a century ago [24], but no structural characterization was performed. Another attempt to characterize Hg2[dca]2 by Kuhn and Mecke [25] in 1961 was based on IR data on a number of different dicyanamides. These measurements provided valid proof that the shape of the dicyanamide anion is kinked (or boomerang-shaped) instead of linear. Even though a working synthesis was known, no further investigations were made regarding the crystal structure of Hg2[dca]2. A further dicyanamide compound containing a monovalent cation with unknown crystal structure is Tl[dca]. Because of its high solubility in water, Tl[dca] cannot be synthesized analogously to Hg2[dca]2. The use of Ag[dca] provides a convenient alternative route to the crystals of Tl[dca]. We here report the syntheses and single-crystal structure determinations of Hg2[dca]2 and Tl[dca].

2. Results and Discussion

2.1. Structural Description and Discussion

Hg2[N(CN)2]2 adopts the triclinic space group P 1 ¯ (no. 2) with a = 3.7089(5), b = 6.4098(6), c = 8.150(6) Å, α = 81.575(6)°, β = 80.379(7)°, γ = 80.195(7)°, and Z = 1. Similar to most Hg(I) compounds, crystalline Hg2[N(CN)2]2 contains Hg2 moieties with d(Hg–Hg) = 2.518(3) Å. The structure can be compared to that of calomel, Hg2Cl2, where d(Hg–Hg) = 2.53 Å [26]. In Hg2[dca]2, every Hg2 dumbbell is bonded to two [dca] anions forming a quasi-molecular Hg2[dca]2 unit (Figure 1). A distorted octahedral coordination results for each Hg(I) due to four more coordinating nitrogen atoms. Therefore, every Hg(I) is surrounded by a total of five nitrogen belonging to five different [dca] anions, and also by the neighboring Hg(I) ion. Three of these dicyanamide anions coordinate the mercury ion with their terminal nitrogen atoms (d(Hg–N1) = 2.166(10), d(Hg–N3) = 2.612(12), and d(Hg–N3) = 2.834(10) Å), whilst the others coordinate via their bridging nitrogen at a more distant d(Hg–N2) = 3.084(9) and 3.412(22) Å. Similar distorted coordination octahedra are also known for Hg2Cl2 [26] (Figure 2).
Tl[dca] crystallizes isotypically to α-K[dca] and α-Rb[dca] [7] in the orthorhombic space group Pbcm (no. 57) with a = 8.5770(17), b = 6.4756(13), c = 7.2306(14) Å, and Z = 4. This is not surprising due to the similar ionic radii of all three cations, Tl+, K+, and Rb+. The structure is built of layers of Tl+ and [dca], in which Tl+ is coordinated by six terminal nitrogen atoms with d(Tl–N1) = 3.082(4)/3.089(4) and d(Tl–N3) = 2.870(4) Å and two bridging nitrogen atoms with d(Tl–N2) = 3.053(5) Å of eight different [dca] moieties, generating a distorted square antiprism (Figure 3). These Tl–N distances are in good agreement with those that were reported for a quadratic antiprismatic coordination polyhedron in TlN3 [27] with d(Tl–N) = 3.03 Å.
Like before, the boomerang-shaped dicyanamide in Tl[dca] exhibits atomic distances and angles in the expected range: the bond length of d(C1–N1) = 1.148(8) and d(C2–N3) = 1.148(7) Å of the terminal C–N pairs indicate triple bonds, while the central C–N with d(C1–N2) = 1.316(8) and d(C2–N2) = 1.319(7) Å of the [dca] anion are also typically found for such a moiety, with ∡(N1–C1–N2) = 172.0(6)°, ∡(N2–C2–N3) = 172.6(6)°, and ∡(C1–N2–C2) = 120.6(6)° (Table 1).
Given the large number of structurally characterized dicyanamides, it is tempting to finally investigate their volume chemistry. Hence, we calculated the volume increment of the dicyanamide anion according to the method of Biltz, that is, taking into account the tabulated volume increments [28] of the monovalent cations M+, namely those of Li, Na, K, Rb, Cs, Cu, Ag, Hg, Tl, and NH4. As shown in Table 2, the dicyanamide anion has a volume of 44.3(17) cm3·mol−1 on average, so it is about 70% larger than the NCN2− cyanamide anion (26 cm3·mol−1) [29] That being said, the dicyanamide anion appears as a mid-heavy (66.04 u) but relatively spacious pseudohalide, and it finalizes the trend found [28] for the true halides (F: 9.5, Cl: 20, Br: 25, I: 34, [dca]: 44 cm3·mol−1). For completeness, we add that the calculated [dca] volume increments for Hg2[dca]2 and Tl[dca] (40.2 and 41.9 cm3·mol−1) appear as slightly smaller, which is probably due to the fact that these crystal structures were determined at a lower temperature (Table 2).

2.2. IR-Spectra

The vibrational frequencies, as obtained from the IR spectra of the title compounds, confirm the presence of the [dca] group (Table 3, Figure 4).

3. Materials and Methods

3.1. Syntheses

Crystalline powder of Hg2[dca]2 was synthesized by mixing stoichiometric amounts of Na[dca] and Hg2(NO3)2·2H2O in water, followed by precipitation, but the fine powder was unsuitable for single-crystal diffraction. Better crystals of Hg2[dca]2 were subsequently obtained by a diffusion reaction. An aqueous solution of Hg2(NO3)2·2H2O (5 mL, c = 0.05 M) was layered below, and another aqueous solution of Na[dca] (5 mL, c = 0.1 M) was layered above an aqueous solution of NaNO3 (5 mL, c = 0.22 M) in a regular test-tube. Colorless, transparent crystals of Hg2[dca]2 grew within two days in the region between the layers of Hg2(NO3)2 and NaNO3.
Tl[dca] was obtained by adding TlBr (184.8 mg, 0.65 mmol) to an aqueous Ag[dca] (120.5 mg, 0.69 mmol, 5 mL deionized H2O) suspension. The synthesis of Ag[dca] has been described in a previous paper [23]. The suspension was stirred for twelve hours under the exclusion of light. 1 mL of the silver bromide-free solution was taken, and the water was evaporated in a desiccator. A colorless, transparent, orthorhombic crystal of Tl[dca] suitable for single-crystal X-ray diffraction was selected for diffraction data collection.

3.2. Single Crystal Diffraction

A suitable single crystal of Hg2[dca]2 was mounted on a glass fiber. Intensity data were collected with a STOE STADIVARI Dectris Pilatus 200K detector (STOE & Cie GmbH, Darmstadt, Germany), equipped with a GeniX Mo High Flux source (Mo Kα radiation, λ = 0.71073 Å, multilayer optics). Temperature control was achieved using an Oxford Cryostream 800 (Oxford Cryosystems Ltd., Oxford, UK) at 100 K. Collected data were integrated with X-Area Integrate [30], and Gaussian-integration absorption corrections were applied with STOE X-Red [31]. The structure was solved by charge-flipping methods (Superflip [32]) and refined on F2, as implemented in Jana2006 [33]. More crystallographic details can be found in Table 4, Table 5 and Table 6. The goodness-of-fit is unusually large (3.60) and it probably goes back to the rather high absorption coefficient for Mo Kα radiation resulting in an imperfect absorption model.
A suitable single crystal of Tl[dca] was adhered to a 100 µm MiTeGen loop using perfluoropolyether PFO-XR75. Intensity data were collected on a FR 591 rotating anode that was equipped with an Incoatec Helios focusing multilayer optic (Mo Kα radiation, λ = 0.71073 Å) and a Bruker AXS Enraf-Nonius KappaCCD detector (Bruker AXS GmbH, Karlsruhe, Germany). The temperature of the crystal was maintained at 100 K using an Oxford Cryostream 700 (Oxford Cryosystems Ltd., Oxford, United Kingdom). Diffraction data were integrated with the program REVALCCD ver. 1.6, 2008 and a Gaussian integration absorption correction based on the crystal shape was applied using SADABS [34]. Data preparation and reciprocal space exploration were performed by XPREP [35]. The structure was solved with SHELXT [36] by a dual-space method and was refined on F2, as implemented in SHELXL [37]. Crystallographic details can be found in Table 4, Table 7 and Table 8.
Additional details concerning the structure determination are available in CIF format (see Supplementary Materials) and have been deposited under the CCDC entry numbers 1881933 for Tl[dca] and 1881934 for Hg2[dca]2. Copies of the data can be obtained free of charge from CCDC (http://www.ccdc.cam.ac.uk/conts/retrieving.html).

3.3. Infrared Spectra

The IR spectra were recorded using an ALPHA II FT-IR-spectrometer (Bruker Optik GmbH, Ettlingen, Germany), equipped with an ATR Platinum Diamond measuring cell. All measurements were undertaken within the range of 4000 to 400 cm−1.

4. Conclusions

The compounds Hg2[dca]2 and Tl[dca] were synthesized, their respective crystal structures were determined, and their IR spectra were measured. While the structure of Hg2[dca]2 shows similarities to the structure of Hg2Cl2, Tl[dca] is isostructural to α-K[dca] and α-Rb[dca]. The [dca] pseudohalide exceeds I in its volume increment. The acquired data of the IR spectra are similar to the data of the previously reported dicyanamides.

Supplementary Materials

The following are available online at https://www.mdpi.com/2304-6740/6/4/135/s1: CIF and CIFchecked files.

Author Contributions

M.M. and O.R. conceived and designed the experiment; M.M. performed the syntheses, ATR-IR experiments and the single-crystal X-ray diffraction (SXRD) of Hg2[dca]2; N.N. and R.G. performed the SXRD of Tl[dca]; results were discussed with all authors; M.M. wrote the paper in collaboration with all co-authors.

Funding

This research received no external funding.

Acknowledgments

Nils Nöthling and Richard Goddard would like to thank Martin Jäger and Joachim Horst, Labor für Instrumentelle und Umweltschutzanalytik, Hochschule Niederrhein, Krefeld, Germany, for helpful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Scholz, T.; Görne, A.L.; Dronskowski, R. Itinerant nitrides and salt-like guanidinates—The diversity of solid-state nitrogen chemistry. Prog. Solid State Chem. 2018, 51, 1–18. [Google Scholar] [CrossRef]
  2. Sougrati, M.T.; Arayamparambil, J.J.; Liu, X.; Mann, M.; Slabon, A.; Stievano, L.; Dronskowski, R. Carbodiimides as energy materials: Which directions for a reasonable future? J. Chem. Soc. Dalton Trans. 2018, 47, 10827–10832. [Google Scholar] [CrossRef] [PubMed]
  3. Sougrati, M.T.; Darwiche, A.; Liu, X.; Mahmoud, A.; Hermann, R.P.; Jouen, S.; Monconduit, L.; Dronskowski, R.; Stievano, L. Transition-Metal Carbodiimides as Molecular Negative Electrode Materials for Lithium- and Sodium-Ion Batteries with Excellent Cycling Properties. Angew. Chem. Int. Ed. 2016, 55, 5090–5095. [Google Scholar] [CrossRef] [PubMed]
  4. Jürgens, B.; Höppe, H.A.; Irran, E.; Schnick, W. Transformation of Ammonium Dicyanamide into Dicyandiamide in the Solid. Inorg. Chem. 2002, 41, 4849–4851. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Reckeweg, O.; DiSalvo, F.J.; Schulz, A.; Blaschkowski, B.; Jagiella, S.; Schleid, T. Synthesis, Crystal Structure, and Vibrational Spectra of the Anhydrous Lithium Dicyanamide Li[N(CN)2]. Z. Anorg. Allg. Chem. 2014, 640, 851–855. [Google Scholar] [CrossRef]
  6. Starynowicz, P. Structure of caesium dicyanamide. Acta Crystallogr. Sect. C Cryst. Struct. Commun. 1991, 47, 2198–2199. [Google Scholar] [CrossRef] [Green Version]
  7. Irran, E.; Jürgens, B.; Schnick, W. Trimerization of alkali dicyanamides M[N(CN)2] and formation of tricyanomelaminates M3[C6N9] (M = K, Rb) in the melt: Crystal structure determination of three polymorphs of K[N(CN)2], two of Rb[N(CN)2], and one of K3[C6N9] and Rb3[C6N9] from X-ray powder diffractometry. Chem. Eur. J. 2001, 7, 5372–5381. [Google Scholar] [PubMed]
  8. Jürgens, B.; Irran, E.; Schneider, J.; Schnick, W. Trimerization of NaC2N3 to Na3C6N9 in the Solid: Ab Initio Crystal Structure Determination of Two Polymorphs of NaC2N3 and of Na3C6N9 from X-ray Powder Diffractometry. Inorg. Chem. 2000, 39, 665–670. [Google Scholar] [CrossRef] [PubMed]
  9. Jürgens, B.; Irran, E.; Schnick, W. Syntheses, Vibrational Spectroscopy, and Crystal Structure Determination from X-Ray Powder Diffraction Data of Alkaline Earth Dicyanamides M[N(CN)2]2 with M = Mg, Ca, Sr, and Ba. J. Solid State Chem. 2001, 157, 241–249. [Google Scholar] [CrossRef]
  10. Manson, J.L.; Kmety, C.R.; Epstein, A.J.; Miller, J.S. Spontaneous Magnetization in the M[N(CN)2]2 (M = Cr, Mn) Weak Ferromagnets. Inorg. Chem. 1999, 38, 2552–2553. [Google Scholar] [CrossRef]
  11. Manson, J.L.; Kmety, C.R.; Huang, Q.-Z.; Lynn, J.W.; Bendele, G.M.; Pagola, S.; Stephens, P.W.; Liable-Sands, L.M.; Rheingold, A.L.; Epstein, A.J.; et al. Structure and Magnetic Ordering of MII[N(CN)2]2 (M = Co, Ni). Chem. Mater. 1998, 10, 2552–2560. [Google Scholar] [CrossRef]
  12. Reckeweg, O.; Dinnebier, R.E.; Schulz, A.; Blaschkowski, B.; Schneck, C.; Schleid, T. About the air- and water-stable copper(I) dicyanamide: Synthesis, crystal structure, vibrational spectra and DSC/TG analysis of Cu[N(CN)2]. Z. Naturforsch. B Chem. Sci. 2017, 72, 159–165. [Google Scholar] [CrossRef]
  13. Hodgson, S.A.; Hunt, S.J.; Sørensen, T.J.; Thompson, A.L.; Reynolds, E.M.; Faulkner, S.; Goodwin, A.L. Anomalous Thermal Expansion and Luminescence Thermochromism in Silver(I) Dicyanamide. Eur. J. Inorg. Chem. 2016, 2016, 4378–4381. [Google Scholar] [CrossRef]
  14. Reckeweg, O.; Schulz, A.; Schneck, C.; Lissner, F.; Schleid, T. Syntheses, single-crystal structures, vibrational spectra and DSC/TG analyses of orthorhombic and trigonal Ag[N(CN)2]. Z. Naturforsch. B Chem. Sci. 2016, 71, 827–834. [Google Scholar] [CrossRef]
  15. Manson, J.L.; Lee, D.W.; Rheingold, A.L.; Miller, J.S. Buckled-layered Structure of Zinc Dicyanamide, ZnII[N(CN)2]2. Inorg. Chem. 1998, 37, 5966–5967. [Google Scholar] [CrossRef] [PubMed]
  16. Jürgens, B.; Irran, E.; Höppe, H.A.; Schnick, W. Phase Transition of a Dicyanamide with Rutile-like Structure: Syntheses and Crystal Structures of α- and β-Cd[N(CN)2]2. Z. Anorg. Allg. Chem. 2004, 630, 219–223. [Google Scholar] [CrossRef]
  17. Jürgens, B.; Irran, E.; Schnick, W. Synthesis and characterization of the rare-earth dicyanamides Ln[N(CN)2]3 with Ln = La, Ce, Pr, Nd, Sm, and Eu. J. Solid State Chem. 2005, 178, 72–78. [Google Scholar] [CrossRef]
  18. Nag, A.; Schmidt, P.J.; Schnick, W. Synthesis and Characterization of Tb[N(CN)2]3·2H2O and Eu[N(CN)2]3·2H2O: Two New Luminescent Rare-Earth Dicyanamides. Chem. Mater. 2006, 18, 5738–5745. [Google Scholar] [CrossRef]
  19. Nag, A.; Schnick, W. Synthesis, Crystal Structure and Thermal Behavior of Gadolinium Dicyanamide Dihydrate Gd[N(CN)2]3·2H2O. Z. Anorg. Allg. Chem. 2006, 632, 609–614. [Google Scholar] [CrossRef]
  20. Reckeweg, O.; Wakabayashi, R.H.; DiSalvo, F.J.; Schulz, A.; Schneck, C.; Schleid, T. About alkali metal dicyanamides: Syntheses, single-crystal structure determination, DSC/TG and vibrational spectra of KCs[N(CN)2]2 and NaRb2[N(CN)2]3·H2O. Z. Naturforsch. B Chem. Sci. 2015, 70, 365–372. [Google Scholar] [CrossRef]
  21. Reckeweg, O.; DiSalvo, F.J. Synthesis and single-crystal structure of the pseudo-ternary compounds LiA[N(CN)2]2 (A = K or Rb). Z. Naturforsch. B Chem. Sci. 2016, 71, 157–160. [Google Scholar] [CrossRef]
  22. Jürgens, B.; Milius, W.; Morys, P.; Schnick, W. Trimerisierung von Dicyanamid-Ionen C2N3 im Festkörper—Synthesen, Kristallstrukturen und Eigenschaften von NaCs2(C2N3)3 und Na3C6N9·3H2O. Z. Anorg. Allg. Chem. 1998, 624, 91–97. [Google Scholar] [CrossRef]
  23. Mann, M.; Reckeweg, O.; Dronskowski, R. Synthesis and Characterization of the New Dicyanamide LiCs2[N(CN)2]3. Inorganics 2018, 6, 108. [Google Scholar] [CrossRef]
  24. Madelung, W.; Kern, E. Über Dicyanamid. Justus Liebigs Ann. Chem. 1922, 427, 1–26. [Google Scholar] [CrossRef]
  25. Kuhn, M.; Mecke, R. IR-Spektroskopische Untersuchungen am Dicyanamid-Anion, [N(CN)2]. Chem. Ber. 1961, 94, 3010–3015. [Google Scholar] [CrossRef]
  26. Dorm, E. Studies on the Crystal Chemistry of the Mercurous Ion and of Mercurous Salts; Stockholm Univ.: Stockholm, Sweden, 1970; pp. 1–25. [Google Scholar]
  27. Reckeweg, O.; Simon, A. Azide und Cyanamide—Ähnlich und doch anders/Azides and Cyanamides – Similar and Yet Different. Z. Naturforsch. B Chem. Sci. 2003, 58, 1097–1104. [Google Scholar] [CrossRef]
  28. Biltz, W. Raumchemie der festen Stoffe; Verlag von Leopold Voss: Leipzig, Germany, 1934. [Google Scholar]
  29. Stork, L.; Liu, X.; Fokwa, B.P.T.; Dronskowski, R. Crystal Structure Determination of Thallium Carbodiimide, Tl2NCN. Z. Anorg. Allg. Chem. 2007, 633, 1339–1342. [Google Scholar] [CrossRef]
  30. X-Area Integrate; 1.71.0.0; Stoe & Cie GmbH: Darmstadt, Germany, 2016.
  31. X-Area X-Red; 1.63.2.0; Stoe & Cie GmbH: Darmstadt, Germany, 2017.
  32. Palatinus, L.; Chapuis, G. SUPERFLIP– a computer program for the solution of crystal structures by charge flipping in arbitrary dimensions. J. Appl. Crystallogr. 2007, 40, 786–790. [Google Scholar] [CrossRef]
  33. Petříček, V.; Dušek, M.; Palatinus, L. Crystallographic Computing System JANA2006: General features. Z. Kristallogr. 2014, 229, 345–352. [Google Scholar] [CrossRef]
  34. SADABS; 2014/5; Bruker AXS Inc.: Madison, WI, USA, 2014.
  35. XPREP; 2014/5; Bruker AXS Inc.: Madison, WI, USA, 2014.
  36. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  37. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Cryst. Struct. Commun. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Crystal structure of Hg2[dca]2. The atomic displacement ellipsoids correspond to 90% probability using the refined anisotropic displacement parameters (ADPs).
Figure 1. Crystal structure of Hg2[dca]2. The atomic displacement ellipsoids correspond to 90% probability using the refined anisotropic displacement parameters (ADPs).
Inorganics 06 00135 g001
Figure 2. Comparison of the crystal structures of Hg2[dca]2 and calomel, Hg2Cl2. (a) Coordination sphere of Hg in Hg2[dca]2. The atomic displacement ellipsoids correspond to 90% probability using the refined ADPs. (b) Crystal structure of Hg2Cl2 [26].
Figure 2. Comparison of the crystal structures of Hg2[dca]2 and calomel, Hg2Cl2. (a) Coordination sphere of Hg in Hg2[dca]2. The atomic displacement ellipsoids correspond to 90% probability using the refined ADPs. (b) Crystal structure of Hg2Cl2 [26].
Inorganics 06 00135 g002
Figure 3. Crystal structure of Tl[dca] with (a) showing the view along c and (b) along b. The atomic displacement ellipsoids correspond to 90% probability using the refined ADPs.
Figure 3. Crystal structure of Tl[dca] with (a) showing the view along c and (b) along b. The atomic displacement ellipsoids correspond to 90% probability using the refined ADPs.
Inorganics 06 00135 g003
Figure 4. Infrared spectra of Hg2[dca]2 (left) and Tl[dca] (right).
Figure 4. Infrared spectra of Hg2[dca]2 (left) and Tl[dca] (right).
Inorganics 06 00135 g004
Table 1. Selected angles (°) and bond lengths (Å) in Hg2[dca]2 and Tl[dca].
Table 1. Selected angles (°) and bond lengths (Å) in Hg2[dca]2 and Tl[dca].
Hg2[dca]2(°)Tl[dca](°)
∡(N1–C1–N2)171.6(14)∡(N1–C1–N2)172.0(6)
∡(N2–C2–N3)173.2(13)∡(N2–C2–N3)172.6(6)
∡(C1–N2–C2)119.4(10)∡(C1–N2–C2)120.6(5)
(Å) (Å)
N1–C11.148(14)N1–C11.148(8)
C1–N21.325(13)C1–N21.316(8)
N2–C21.317(16)N2–C21.319(7)
C2–N31.108(16)C2–N31.148(7)
Hg–Hg2.518(3)Tl–Tl (2×)3.6153(7)
Hg–N12.166(10)Tl–Tl (2×)4.5129(7)
Hg–N23.412(11)Tl–N1 (2×)3.089(4)
Hg–N23.084(9)Tl–N1 (2×)3.082(4)
Hg–N32.612(12)Tl–N2 (2×)3.053(5)
Hg–N32.834(10)Tl–N3 (2×)2.870(4)
Table 2. Calculated molar volumes of [dca] based on M[dca] crystal structures containing monovalent cations.
Table 2. Calculated molar volumes of [dca] based on M[dca] crystal structures containing monovalent cations.
M[dca]Cell Volume (Å3), Z T (K)Molar Volume (cm3·mol−1)M+ Volume (cm3·mol−1) [28][dca] Volume (cm3·mol−1)
NH4[dca] [4]428.32, 420064.519.545.0
Li[dca] [5]313.84, 417047.21.545.7
Na[dca] [8]345.25, 429352.06.545.5
α-K[dca] [7]390.31, 429358.71642.7
β-K[dca] [7]416.78, 429362.71646.7
γ-K[dca] [7]415.56, 429362.51646.5
α-Rb[dca] [7]433.4, 429365.22045.2
β-Rb[dca] [7]1778.31, 1629366.92046.9
Cs[dca] [6]919.74, 829969.22643.2
Cu[dca] [12]332.21, 424850.0545.0
Ag[dca], o [14]349.24, 429352.6943.6
Ag[dca], tr [14]256.36, 329351.4942.4
Hg2[dca]2186.84, 1100112.41640.2
Tl[dca]401.60, 410060.418.541.9
Table 3. IR data of Hg2[dca]2 and Tl[dca]. All numbers are given in cm−1.
Table 3. IR data of Hg2[dca]2 and Tl[dca]. All numbers are given in cm−1.
Vibrationν(Hg2[dca]2)ν(Tl[dca])
σas(N–C≡N)493/515517
σs(N–C≡N)662656
νs(N–C)934905
νas(N–C)13541323
νas(N≡C)21712119
νas(N–C) + νs(N–C)22432191
νs(N≡C)23012247
νas(N≡C) + νs(N–C)30723010/2972
νs(N≡C) + νas(N–C)35773508
Table 4. Summary of single-crystal X-ray diffraction structure determination data of Tl[dca] and Hg2[dca]2.
Table 4. Summary of single-crystal X-ray diffraction structure determination data of Tl[dca] and Hg2[dca]2.
Chemical FormulaHg2[dca]2Tl[dca]
Formula weight (g∙mol−1)533.3270.42
Crystal systemtriclinicorthorhombic
Space groupP 1 ¯ (no. 2)Pbcm (no. 57)
Temperature (K)100(2)100(2)
a (Å)3.7089(5)8.5770(17)
b (Å)6.4098(6)6.4756(13)
c (Å)8.150(6)7.2306(14)
α (°)81.575(6)90
β (°)80.379(7)90
γ (°)80.195(7)90
V3)186.84(14)401.60(14)
Z14
Radiation, λ (Å)Mo Kα, 0.71073Mo Kα, 0.71073
μ (mm−1)41.0940.022
Crystal shape and colorColorless blockColorless block
Crystal size (mm3)0.17 × 0.08 × 0.020.05 × 0.04 × 0.02
ρcalcd (g∙cm−3)4.7394.473
DiffractometerSTOE STADIVARI with Hybrid Pixel Counting DetectorBruker AXS Enraf-Nonius with KappaCCD Detector
Absorption correctionGaussian Integration, STOE X-REDGaussian Integration, SADABS 2014/15
Tmin, Tmax0.0672, 0.67300.15236, 0.46282
No. of measured, independent and observed [I > 3σ(I)] reflections36718880
1263816
1154549
Robs4.681.87
wR2obs10.883.94
Rall4.923.25
wR2all10.914.46
GOFobs3.601.09
No. of parameters, restraints55, 036, 0
Table 5. Atomic coordinates (all on 2i) and equivalent isotropic displacement parameters Ueq2) of Hg2[dca]2.
Table 5. Atomic coordinates (all on 2i) and equivalent isotropic displacement parameters Ueq2) of Hg2[dca]2.
AtomxyzUeq
Hg0.96205(10)0.90300(6)0.14632(4)0.01463(12)
N10.923(3)0.805(2)0.4137(12)0.024(3)
N20.619(3)0.7427(15)0.7062(10)0.015(2)
N30.406(3)0.4138(17)0.8391(12)0.020(3)
C10.765(3)0.7693(19)0.5460(13)0.016(3)
C20.510(3)0.559(2)0.7710(14)0.018(3)
Table 6. Anisotropic displacement parameters Uij2) of Hg2[dca]2.
Table 6. Anisotropic displacement parameters Uij2) of Hg2[dca]2.
AtomU11U22U33U23U13U12
Hg0.01303(19)0.0173(2)0.01206(19)−0.00301(13)0.00010(12)0.00189(13)
N10.026(5)0.038(6)0.011(4)−0.014(4)0.005(3)−0.012(4)
N20.023(4)0.012(4)0.009(3)0.001(3)−0.001(3)−0.004(3)
N30.025(5)0.018(5)0.018(4)−0.005(4)−0.002(3)−0.006(3)
C10.008(4)0.022(5)0.018(4)−0.005(4)0.000(3)0.000(4)
C20.006(4)0.027(6)0.020(5)−0.002(4)0.004(3)−0.010(4)
Table 7. Atomic coordinates and equivalent isotropic displacement parameters Ueq2) of Tl[dca].
Table 7. Atomic coordinates and equivalent isotropic displacement parameters Ueq2) of Tl[dca].
AtomSitexyzUeq
Tl4c0.31674(2)¼½0.01287(6)
N14d0.5898(6)0.3835(8)¾0.0183(10)
N24d0.7848(7)0.1054(9)¾0.0218(12)
N34d0.0707(6)0.1391(8)¾0.0194(10)
C14d0.6880(5)0.2630(11)¾0.0144(10)
C24d0.9369(7)0.1356(9)¾0.0150(10)
Table 8. Anisotropic displacement parameters Uij2) of Tl[dca].
Table 8. Anisotropic displacement parameters Uij2) of Tl[dca].
AtomU11U22U33U23U13U12
Tl0.01236(9)0.01241(10)0.01384(9)−0.00016(13)00
N10.0111(19)0.013(2)0.031(3)00−0.002(2)
N20.012(2)0.010(2)0.043(4)00−0.003(2)
N30.015(2)0.012(2)0.031(3)000.0007(19)
C10.012(2)0.009(3)0.022(3)00−0.003(3)
C20.017(3)0.008(2)0.020(3)00−0.002(2)

Share and Cite

MDPI and ACS Style

Mann, M.; Reckeweg, O.; Nöthling, N.; Goddard, R.; Dronskowski, R. Syntheses and Characterization of Two Dicyanamide Compounds Containing Monovalent Cations: Hg2[N(CN)2]2 and Tl[N(CN)2]. Inorganics 2018, 6, 135. https://doi.org/10.3390/inorganics6040135

AMA Style

Mann M, Reckeweg O, Nöthling N, Goddard R, Dronskowski R. Syntheses and Characterization of Two Dicyanamide Compounds Containing Monovalent Cations: Hg2[N(CN)2]2 and Tl[N(CN)2]. Inorganics. 2018; 6(4):135. https://doi.org/10.3390/inorganics6040135

Chicago/Turabian Style

Mann, Markus, Olaf Reckeweg, Nils Nöthling, Richard Goddard, and Richard Dronskowski. 2018. "Syntheses and Characterization of Two Dicyanamide Compounds Containing Monovalent Cations: Hg2[N(CN)2]2 and Tl[N(CN)2]" Inorganics 6, no. 4: 135. https://doi.org/10.3390/inorganics6040135

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