Skip to content
BY 4.0 license Open Access Published by De Gruyter March 21, 2019

Serendipitous formation and characterization of K2[Pd(NO3)4]·2HNO3

  • Michael Zoller , Jörn Bruns , Gunter Heymann , Klaus Wurst and Hubert Huppertz EMAIL logo

Abstract

A potassium tetranitratopalladate(II) with the composition K2[Pd(NO3)4] · 2HNO3 was synthesized by a simple solvothermal process in a glass ampoule. The new compound crystallizes in the monoclinic space group P21/c (no. 14) with the lattice parameters a = 1017.15(4), b = 892.94(3), c = 880.55(3) Å, and β = 98.13(1)° (Z = 2). The crystal structure of K2[Pd(NO3)4] · 2HNO3 reveals isolated complex [Pd(NO3)4]2− anions, which are surrounded by eight potassium cations and four HNO3 molecules. The complex anions and the cations are associated in layers which are separated by HNO3 molecules. K2[Pd(NO3)4] · 2HNO3 can thus be regarded as a HNO3 intercalation variant of β-K2[Pd(NO3)4]. The characterization is based on single-crystal X-ray and powder X-ray diffraction.

1 Introduction

Synthesis and characterization of new binary and ternary borates is an intensively investigated field of research. Through different reaction conditions not only BO3 but also BO4 groups can be formed [1]. Thus, the structural diversity is immense and properties like second harmonic generation are accessible, especially for compounds with different anionic substructures. Further promising candidates are heteroleptic compounds with different anionic species like Sn3[B3O7]F [2] and Pb3B6O11F2 [3]. With the synthesis of K3Na[B6O9(OH)3]NO3 [4], Lu2B2O5(NO3)2·2H2O [5], and Pr[B5O8(OH)(H2O)0.87]NO3·2H2O [6], we contributed first results on the synthesis of borate nitrates. These compounds were obtained from a low acidic medium, essentially via hydrothermal synthesis. To avoid the implementation of water, we now investigated reactions of suitable metal precursors and borate species with fuming nitric acid (N2O4 in HNO3), in which N2O4 will capture remaining molecules of water under formation of a HNO3-HNO2 mixture.

A serendipitous result of the reaction between elemental palladium, K2B4O7·4 H2O, and fuming nitric acid is the formation of K2[Pd(NO3)4]·2HNO3. Tetranitratopalladates are well known and comprise several compounds like α-K2[Pd(NO3)4] [7], which is similar to K2[Pt(NO3)4]·½H2O [8]. α-K2[Pd(NO3)4] was already published by Elding, Norén, and Oskarsson in 1986. In both compounds, the four squarely coordinated nitrate groups were found to be situated on the same side of the coordination plane (Fig. 1, left).

Fig. 1: Comparison of the complex anions [Pd(NO3)4]2− in α-K2[Pd (NO3)4] (left) and β-K2[Pd(NO3)4] (right). In α-K2[Pd(NO3)4], the nitrate groups are coordinated on one side of the coordination plane, whereas in β-K2[Pd(NO3)4] the nitrate groups are turned in pairs.
Fig. 1:

Comparison of the complex anions [Pd(NO3)4]2− in α-K2[Pd (NO3)4] (left) and β-K2[Pd(NO3)4] (right). In α-K2[Pd(NO3)4], the nitrate groups are coordinated on one side of the coordination plane, whereas in β-K2[Pd(NO3)4] the nitrate groups are turned in pairs.

In the year 2000, the sodium tetranitratopalladate Na2[Pd(NO3)4] [9] was synthesized, and in 2005, the compounds Rb2[Pd(NO3)4] [10] and Cs2[Pd(NO3)4] [10] were discovered. All three structures consist of [Pd(NO3)4]2− anions with monodentate nitrate groups turned in pairs in one direction, just like in β-K2[Pd(NO3)4] [11], which was synthesized in 2009. A comparison of the complex anions in α-K2[Pd(NO3)4] (left) and β-K2[Pd(NO3)4] (right) is shown in Fig. 1. The tetranitratopalladate system has now been complemented by the synthesis of the novel compound K2[Pd(NO3)4]·2HNO3 described in the following.

2 Experimental section

2.1 Synthesis

The reaction was performed in an evacuated, torch sealed borosilicate glass ampoule (l=300 mm, Ø=16 mm, thickness of the tube wall=1.8 mm). A mixture of 50 mg (0.47 mmol) Pd and 100 mg (0.33 mmol) K2B4O7·4 H2O in 1 mL fuming nitric acid was heated up to a temperature of 353 K and kept there for 24 h. Afterwards, the ampoule was cooled with a rate of 6 K h−1 down to room temperature. Red platelets were obtained on the ampoule wall. The mother liquor was separated from the crystals by decantation, the acid containing side was cooled with liquid nitrogen and after cracking of the ampoule, several crystals were transferred directly into inert oil. The remaining bulk material was dried under reduced pressure in a Schlenk flask and transferred into the glovebox for further treatment. The handling was carried out strictly under inert conditions, as the compound appears to be sensitive to moisture.

2.2 X-ray structure determination

A STOE Stadi P diffractometer with curved Ge(111)-monochromatized Mo1 radiation (λ=0.7093 Å) was used to characterize a polycrystalline sample by X-ray powder diffraction. Diffraction intensities were measured with a Mythen2 1 K detector (Dectris, Switzerland) microstrip detector with 1280 strips.

The Rietveld refinements were accomplished with the software package Diffracplus-Topas® 4.2 (Bruker AXS GmbH, Karlsruhe, Germany) [12] based on single-crystal data for β-K2[Pd(NO3)4] [11] (collection code: 166766) and β-HBO2 [13] (collection code: 34640) from the ICSD [14], [15], [16]. The reflection shapes were modelled using modified Thompson-Cox-Hastings pseudo-Voigt profiles [17], [18]. Instrumental contributions on reflection profiles were corrected from the refinement of a standard (LaB6) [19]. The background was fitted with Chebychev polynomials up to the 6th order [20].

Under a polarisation microscope, a suitable single crystal of K2[Pd(NO3)4]·2HNO3 with a diameter of 30 μm was fixed on the tip of a MicroMount™ (MiTeGen, LLC, Ithaca, NY, USA) and immediately placed into a stream of cold N2 inside the diffractometer. The intensity data was collected with a Bruker D8 Quest diffractometer (Bruker, Karlsruhe, Germany) equipped with a Photon 100 detector system and an Incoatec microfocus source generator (multi-layered optic, monochromatized MoKα radiation, λ=0.71073 Å). The collection strategy, concerning the ω and φ scans, was optimized using the Apex 3 [21] program package. Thus, a complete data set up to high angles with high redundancies was received. For data processing and data reduction, the program Saint [22] was employed. Thereafter, multi-scan absorption corrections were applied with the program Sadabs [23].

In the course of the structure solution and parameter refinement with anisotropic displacement parameters for all non-hydrogen atoms with the Shelxs/l-17 [24], [25], [26], [27] software suite, the monoclinic space group P21/c was found to be correct. The hydrogen atom H7 was refined isotropically, while employing a bond restraint of 0.84(2) Å using the DFIX command.

Relevant details of the data collection and evaluation are listed in Table 1, the atomic coordinates and the isotropic displacement parameters in Table 2, and the anisotropic displacement parameters in Table 3. Interatomic distances and bond angles are shown in Table 4 and the hydrogen bond data in Table 5.

Table 1:

Crystal data and structure refinement of K2[Pd(NO3)4]·2HNO3 (standard deviations in parentheses).

Empirical formulaK2[Pd(NO3)4]·2HNO3
Molar mass, gmol−1558.68
Crystal systemMonoclinic
Space groupP21/c (no. 14)
Single-crystal diffractometerBruker D8 Quest Photon 100
Radiation; wavelength, ÅMo; λ=0.71073
a, Å10.1715(4)
b, Å8.9294(3)
c, Å8.8055(3)
β, deg98.13(1)
V, Å3791.73(5)
Formula units per cell, Z2
Calculated density, g cm−32.34
Crystal size, mm30.05×0.05×0.05
Temperature, K213(2)
Absorption coefficient, mm−11.8
F(000), e544
2θ range, deg6.1–75.0
Range in hkl±17, ±15, ±15
Total no. of reflections53342
Independent reflections/Rint4162/0.0474
Data/restraints/parameters4162/1/138
Absorption correctionMulti-scan (Bruker Sadabs 2016/2)
Final R1/wR2 [I>2 σ(Io)]0.0334/0.0766
Final R1/wR2 (all data)0.0531/0.0820
Goodness-of-fit on Fi21.061
Largest diff. peak/hole, e Å−30.98/–1.28
Table 2:

Fractional atomic coordinates and equivalent isotropic displacement parameters Ueq2) for K2[Pd(NO3)4]·2HNO3 (space group P21/c).

AtomxyzUeq
Pd1½½½0.01676(5)
K10.7079(2)0.06281(8)0.6449(2)0.0405(4)
K20.7638(9)0.0555(7)0.5909(9)0.121(2)
O10.3180(2)0.4251(2)0.5267(2)0.0277(3)
O20.1773(2)0.2423(2)0.4866(3)0.0483(5)
O30.3592(2)0.2337(2)0.3863(2)0.0387(4)
O40.5818(2)0.3378(2)0.6396(2)0.0259(3)
O50.5786(2)0.4923(2)0.8317(2)0.0358(3)
O60.6698(2)0.2737(2)0.8664(2)0.0340(3)
O70.0719(2)0.4393(3)0.6630(3)0.0486(5)
O80.9092(2)0.2801(3)0.6397(4)0.0738(8)
O90.9018(3)0.4768(3)0.7782(5)0.080(1)
N10.2861(2)0.2960(2)0.4637(2)0.0300(3)
N20.6107(2)0.3706(2)0.7844(2)0.0238(3)
N30.9519(2)0.3945(3)0.6972(3)0.0461(5)
H70.101(4)0.370(4)0.609(4)0.07(2)
  1. Ueq is defined as one third of the trace of the orthogonalized Uij tensor (standard deviations in parentheses). All atoms are on the Wyckoff position 4e, except palladium, which is on the Wyckoff position 2b. All sites are fully occupied with the exception of K1 exhibiting a s.o.f. of 0.72 and K2 with a s.o.f. of 0.28.

Table 3:

Anisotropic displacement parameters (Å2) of K2[Pd(NO3)4]·2HNO3 (space group P21/c) with standard deviations in parentheses.

AtomU11U22U33U23U13U12
Pd10.02185(8)0.01332(7)0.01527(7)0.00013(6)0.00313(5)–0.00148(7)
K10.0531(6)0.0218(3)0.0427(5)–0.0001(3)–0.0067(3)–0.0011(3)
K20.144(5)0.101(4)0.128(5)0.014(4)0.052(5)0.016(4)
O10.0274(6)0.0202(6)0.0371(7)–0.0050(5)0.0100(6)–0.0059(5)
O20.0396(9)0.0407(10)0.0673(13)–0.0089(9)0.0165(9)–0.0226(8)
O30.0432(9)0.0251(7)0.0501(10)–0.0123(7)0.0147(8)–0.0072(7)
O40.0386(8)0.0205(6)0.0178(5)0.0019(5)0.0019(5)0.0050(5)
O50.0523(10)0.0312(7)0.0237(6)–0.0046(6)0.0043(6)0.0046(7)
O60.0399(8)0.0361(8)0.0248(7)0.0115(6)0.0011(6)0.0050(7)
O70.0344(9)0.0584(13)0.0545(12)–0.0122(11)0.0110(8)–0.0187(9)
O80.0472(13)0.0578(14)0.119(2)–0.0187(15)0.0217(14)–0.0255(11)
O90.0636(17)0.0754(19)0.110(3)–0.0270(17)0.0451(18)–0.0115(14)
N10.0324(9)0.0213(7)0.0366(9)–0.0013(6)0.0060(7)–0.0074(6)
N20.0267(7)0.0261(7)0.0191(6)0.0035(5)0.0053(5)0.0002(6)
N30.0294(10)0.0487(13)0.0607(15)0.0020(11)0.0078(10)–0.0076(9)
Table 4:

Selected interatomic distances (Å) and selected bond angles (deg) in K2[Pd(NO3)4]·2HNO3 (standard deviations in parentheses).

AtomsDistanceAtomsBond angle
Pd1–O42.003(2)O4–Pd–O189.83(6)
Pd1–O4a2.003(2)O4–Pd–O190.17(6)
Pd1–O12.013(2)Ø=90
Pd1–O1a2.013(2)
Ø=2.01
N2–O41.301(2)O3–N1–O2122.8(2)
N2–O51.225(2)O3–N1–O1120.7(2)
N2–O61.228(2)O2–N1–O1116.5(2)
Ø=1.25Ø=120
N1–O11.301(2)O5–N2–O6123.9(2)
N1–O21.247(2)O5–N2–O4119.9(2)
N1–O31.213(3)O6–N2–O4116.2(2)
Ø=1.25Ø=120
N3–O71.358(3)O9–N3–O8128.5(3)
N3–O81.194(3)O9–N3–O7115.3(3)
N3–O91.188(4)O8–N3–O7116.2(3)
Ø=1.25Ø=120
  1. aSymmetry operation: 1–x, 1–y, 1–z.

Table 5:

The hydrogen bond paramters (Å, deg) in K2[Pd(NO3)4]·2HNO3.

H BondD–HHLADLAD–HLA
O7–H7LO20.86(4)1.82(4)2.670(4)1.72(4)

CCDC 1897204 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre viawww.ccdc.cam.ac.uk/data_request/cif.

3 Results and discussion

3.1 Crystal structure

According to the single-crystal diffraction data, K2[Pd(NO3)4]·2HNO3 crystallizes in the monoclinic space group P21/c (no. 14) with the cell parameters a=10.1715(4), b=8.9294(3), c=8.8055(3) Å, and β=98.13(1)°. The [Pd(NO3)4]2− anion consists of a central palladium atom, which is coordinated monodentately in a square planar fashion by four NO3 groups, as is depicted in Fig. 2 (left) in comparison to the similar [Pd(NO3)4]2− anion found in β-K2[Pd(NO3)4] (right). The mean Pd–O bond length in K2[Pd(NO3)4]·2HNO3 is 2.01 Å and, as the palladium atom is situated on a centre of inversion, only two crystallographically independent nitrate groups are distinguishable. The deviation of the O–Pd–O angle from the theoretical value for a perfect square planar coordination is 0.17°. The nitrate groups are nearly planar, deviating only 0.2(2)° for the nitrate group of N1 and 0.3(2)° for the nitrate group of N2 from planarity. The terminal N–O distances with 1.213(3) to 1.247(2) Å are noticeably shorter than the palladium coordinating O–N distances possessing values of 1.301(2) Å each. The O–N–O bond angles vary from 116.2(2) to 123.9(2)° and are in good accordance with the theoretical value of 120° for a trigonal planar structure.

Fig. 2: The [Pd(NO3)4]2− anion in K2[Pd(NO3)4]·2HNO3 (left) is built up from a central palladium atom, which is coordinated in a monodentate fashion by four NO3− anions. It is very similar to the anion in β-K2[Pd(NO3)4] (right).
Fig. 2:

The [Pd(NO3)4]2− anion in K2[Pd(NO3)4]·2HNO3 (left) is built up from a central palladium atom, which is coordinated in a monodentate fashion by four NO3 anions. It is very similar to the anion in β-K2[Pd(NO3)4] (right).

Each [Pd(NO3)4]2− anion is surrounded by eight potassium cations with Pd···K distances of 4.053(2) to 5.800(1) Å, which is in good agreement with Pd···K distances of 4.3905(4) to 5.5658(4) Å found for β-K2[Pd(NO3)4] [11]. The two crystallographically different potassium sites are not fully occupied. The K1 position reveals a site occupancy factor (s.o.f.) of 0.72 and the K2 site a s.o.f. of 0.28. The initial results were obtained by refining the potassium positions with a free variable. Determining the same positions freely results in similar values of K1 s.o.f. 0.70 and K2 s.o.f. 0.15. The potassium cation K1 is surrounded by ten oxygen atoms, with K–O distances ranging from 2.738(2) to 3.242(2) Å. The potassium cation K2 is surrounded by nine oxygen atoms, with K–O distances ranging from 2.493(8) to 3.354(9) Å. Both potassium sites are depicted in Fig. 3. The potassium cations are surrounded by NO3 anions and HNO3 molecules. The distance between the two potassium sites is 0.794(9) Å.

Fig. 3: Each potassium site is surrounded by NO3− anions. Left: The K1 site is surrounded by six NO3− anions and one HNO3 molecule. Right: The K2 site is surrounded by four NO3− anions and two HNO3 molecules.
Fig. 3:

Each potassium site is surrounded by NO3 anions. Left: The K1 site is surrounded by six NO3 anions and one HNO3 molecule. Right: The K2 site is surrounded by four NO3 anions and two HNO3 molecules.

The [Pd(NO3)4]2− anion is surrounded by four HNO3 molecules. The array allows for hydrogen bonds between hydrogen atoms H7 and the oxygen atoms O2 with an H7···O2 distance of 1.82(4) Å and an O7–H7···O2 bond angle of 172(4)° (see Fig. 4 and for crystallographic details Table 5). The hydrogen atom was fixed at a bond length of 0.84(2) Å to the oxygen atom O7 using the DFIX command. The HNO3 molecule is nearly planar with a O9–O8–O7–H7 torsion angle of 4(3)° and exhibits two shorter and one longer N–O distances with 1.188(4), 1.194(3), and 1.358(3) Å, respectively. The hydrogen atom is bonded to the oxygen atom belonging to the elongated N–O bond. The O–N–O bond angles comprise two smaller and one larger angle with 115.3(3), 116.2(3), and 128.5(3)°.

Fig. 4: The moderately strong hydrogen bond O7–H7···O2 is shown as a fragmented line [28].
Fig. 4:

The moderately strong hydrogen bond O7–H7···O2 is shown as a fragmented line [28].

The general packing motif for the complex anion is determined by two alternating orientations for the square-planar complex anion with a change of 74.73(4)°, as is shown in Fig. 5. K2[Pd(NO3)4]·2HNO3 can be regarded as a HNO3 intercalation variant of β-K2[Pd(NO3)4]. The [Pd(NO3)4]2− anion in β-K2[Pd(NO3)4] is surrounded by potassium cations and further complex anions. In K2[Pd(NO3)4]·2HNO3 on the other hand, the complex anion is surrounded by disordered potassium cations and HNO3 molecules along the a axis. Both compounds are very similar with the main difference being the intercalation of HNO3 groups.

Fig. 5: Left: K2[Pd(NO3)4]·2HNO3 featuring layers along the bc plane and stacks along the a axis. It can be regarded as a HNO3 intercalation variant of β-K2[Pd(NO3)4] (right). The different potassium sites are shown in purple. The less occupied K2 site is emphasized in form of transparent spheres.
Fig. 5:

Left: K2[Pd(NO3)4]·2HNO3 featuring layers along the bc plane and stacks along the a axis. It can be regarded as a HNO3 intercalation variant of β-K2[Pd(NO3)4] (right). The different potassium sites are shown in purple. The less occupied K2 site is emphasized in form of transparent spheres.

3.2 Powder diffraction data

In a glovebox, the bulk material was ground to a powder in an agate mortar and transferred to a glass capillary (Ø=0.3 mm). Figure 6 shows the experimentally obtained powder diffraction data (black; top) in comparison to the data simulated from the single-crystal diffraction data (red; bottom). It is obvious that the diffraction patterns do not match. Upon further investigation it became clear that the sample already decomposed, probably caused by vacuum treatment, which was necessary for the transfer into the glovebox. The powder pattern can be identified as a mixture of β-K2[Pd(NO3)4] and β-HBO2 [13], which formed as a side product of the reaction. The Rietveld refinement of the decomposition products is shown in Fig. 7. The experimental data are shown in black, the best fit profile in red, and the difference curve in blue.

Fig. 6: Observed powder pattern (MoKα1 radiation, λ=0.7093 Å) (top), compared to the theoretical powder pattern of K2[Pd(NO3)4]·2HNO3 calculated from the single-crystal data (bottom).
Fig. 6:

Observed powder pattern (Mo1 radiation, λ=0.7093 Å) (top), compared to the theoretical powder pattern of K2[Pd(NO3)4]·2HNO3 calculated from the single-crystal data (bottom).

Fig. 7: Rietveld refinement of the powder X-ray diffraction data. The observed powder pattern can be attributed to β-K2[Pd(NO3)4] and β-HBO2.
Fig. 7:

Rietveld refinement of the powder X-ray diffraction data. The observed powder pattern can be attributed to β-K2[Pd(NO3)4] and β-HBO2.

4 Conclusion

A new potassium tetranitratopalladate with the composition K2[Pd(NO3)4]·2HNO3 was synthesized via solvothermal synthesis in a torch-sealed glass ampoule. The structure is composed of [Pd(NO3)4]2− anions and potassium cations, similar to β-K2[Pd(NO3)4]. However, HNO3 molecules reside between the layer-like arrangement of [Pd(NO3)4]2− anions and potassium cations along the bc plane.

Acknowledgement

Jörn Bruns is thankful for a research stipend (BR5269/1) donated by the DFG.

References

[1] D. Vitzthum, K. Wurst, H. Huppertz, Z. Naturforsch.2018, 73b, 919.10.1515/znb-2018-0141Search in Google Scholar

[2] S. Schönegger, S. G. Jantz, A. Saxer, L. Bayarjargal, B. Winkler, F. Pielnhofer, H. A. Höppe, H. Huppertz, Chem. Eur. J.2018, 24, 16036.10.1002/chem.201803478Search in Google Scholar

[3] H. Li, H. Wu, X. Su, H. Yu, S. Pan, Z. Yang, Y. Lu, J. Han, K. Poeppelmeier, J. Mater. Chem.2014, C2, 1704.10.1039/c3tc32002bSearch in Google Scholar

[4] T. S. Ortner, K. Wurst, L. Perfler, M. Tribus, H. Huppertz, J. Solid State Chem.2015, 221, 66.10.1016/j.jssc.2014.09.016Search in Google Scholar

[5] T. S. Ortner, K. Wurst, C. Hejny, H. Huppertz, J. Solid State Chem.2016, 233, 329.10.1016/j.jssc.2015.11.005Search in Google Scholar

[6] T. S. Ortner, H. Huppertz, Z. Naturforsch.2017, 72b, 677.10.1515/znb-2017-0121Search in Google Scholar

[7] L. I. Elding, B. Norén, Å. Oskarsson, Inorg. Chim. Acta1986, 114, 71.10.1016/S0020-1693(00)84591-0Search in Google Scholar

[8] L. I. Elding, Å. Oskarsson, Inorg. Chim. Acta1985, 103, 127.10.1016/S0020-1693(00)87478-2Search in Google Scholar

[9] S. Khranenko, I. Baidina, S. Gromilov, A. Belyaev, J. Struct. Chem.2000, 41, 709.10.1007/BF02683937Search in Google Scholar

[10] S. Khranenko, I. Baidina, S. Gromilov, A. Belyaev, J. Struct. Chem.2005, 46, 1060.10.1007/s10947-006-0242-7Search in Google Scholar

[11] S. Khranenko, I. Baidina, S. Gromilov, J. Struct. Chem.2009, 50, 361.10.1007/s10947-009-0051-xSearch in Google Scholar

[12] A. A. Coelho, Topas-academic, Coelho Software, Brisbane (Australia) 2007.Search in Google Scholar

[13] W. Zachariasen, Acta Crystallogr.1963, 16, 385.10.1107/S0365110X6300102XSearch in Google Scholar

[14] G. Bergerhoff, I. Brown, F. Allen, Crystallographic Databases, Vol. 360, International Union of Crystallography, Chester (U.K.) 1987.Search in Google Scholar

[15] A. Belsky, M. Hellenbrandt, V. L. Karen, P. Luksch, Acta Crystallogr.2002, B58, 364.10.1107/S0108768102006948Search in Google Scholar PubMed

[16] R. Allmann, R. Hinek, Acta Crystallogr.2007, A63, 412.10.1107/S0108767307038081Search in Google Scholar PubMed PubMed Central

[17] P. Thompson, D. Cox, J. Hastings, J. Appl. Crystallogr.1987, 20, 79.10.1107/S0021889887087090Search in Google Scholar

[18] R. Young, P. Desai, Arch. Nauki Mater.1989, 10, 71.Search in Google Scholar

[19] M. Morris, H. McMurdie, E. Evans, B. Paretzkin, H. Parker, N. Pyrros, Natl. Bur. Stand. (US) Monogr.1984, 25, 62.Search in Google Scholar

[20] L. McCusker, R. Von Dreele, D. Cox, D. Louër, P. Scardi, J. Appl. Crystallogr.1999, 32, 36.10.1107/S0021889898009856Search in Google Scholar

[21] Apex3 (version 2017 3-0), Bruker AXS GmbH, Karlsruhe (Germany) 2017.Search in Google Scholar

[22] Saint (version 8.38A), Bruker AXS GmbH, Karlsruhe (Germany) 2016.Search in Google Scholar

[23] Sadabs (version 2016/2), Bruker AXS GmbH, Karlsruhe (Germany) 2016.Search in Google Scholar

[24] G. M. Sheldrick, Shelxs/l-17, Program Suite for the Solution and Refinement of Crystal Structures, University of Göttingen, Göttingen (Germany) 2017.Search in Google Scholar

[25] G. M. Sheldrick, Acta Crystallogr2015, A71, 3.Search in Google Scholar

[26] G. M. Sheldrick, Shelxl (version 2017-/1), Crystal Structure Refinement-Multi-CPU Version, University of Göttingen, Göttingen (Germany) 2017.Search in Google Scholar

[27] G. M. Sheldrick, Acta Crystallogr.2015, C71, 3.Search in Google Scholar

[28] G. A. Jeffrey, An Introduction to Hydrogen Bonding, Oxford University Press, New York, 1997.Search in Google Scholar

Received: 2019-02-14
Accepted: 2019-02-28
Published Online: 2019-03-21
Published in Print: 2019-04-24

© 2019 Hubert Huppertz et al., published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 Public License.

Downloaded on 24.4.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2019-0018/html
Scroll to top button