Next Article in Journal
Unraveling Nanoscale Magnetic Ordering in Fe3O4 Nanoparticle Assemblies via X-rays
Next Article in Special Issue
Closing the Circle of the Lanthanide-Murexide Series: Single-Molecule Magnet Behavior and Near-Infrared Emission of the NdIII Derivative
Previous Article in Journal
Effects of Vertical Magnetohydrodynamic Flows on Chiral Surface Formation in Magnetoelectrolysis
Previous Article in Special Issue
Slow Relaxation of the Magnetization in Bis-Decorated Chiral Helicene-Based Coordination Complexes of Lanthanides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

How to Make a Better Magnet? Insertion of Additional Bridging Ligands into a Magnetic Coordination Polymer

Faculty of Chemistry, Jagiellonian University, Gronostajowa 2, 30-387 Kraków, Poland
*
Author to whom correspondence should be addressed.
Magnetochemistry 2018, 4(3), 41; https://doi.org/10.3390/magnetochemistry4030041
Submission received: 31 July 2018 / Revised: 6 September 2018 / Accepted: 10 September 2018 / Published: 15 September 2018
(This article belongs to the Special Issue Multifunctional Molecule-based Magnetic Materials)

Abstract

:
A three-dimensional cyanide-bridged coordination polymer based on FeII (S = 2) and NbIV (S = 1/2) {[FeII(H2O)2]2[NbIV(CN)8]·4H2O}n (Fe2Nb) was modified at the self-assembly stage by inserting an additional formate HCOO bridge into its cyanide framework. The resulting mixed-bridged {(NH4)[(H2O)FeII-(μ-HCOO)-FeII(H2O)][NbIV(CN)8]·3H2O}n (Fe2NbHCOO) exhibited additional FeII-HCOO-FeII structural motifs connecting each of the two FeII centers. The insertion of HCOO was possible due to the substitution of some of the aqua ligands and crystallization water molecules in the parent framework by formate anions and ammonium cations. The formate molecular bridge not only shortened the distance between FeII ions in Fe2NbHCOO from 6.609 Å to 6.141 Å, but also created additional magnetic interaction pathways between the magnetic centers, resulting in an increase in the long range magnetic ordering temperature from 43 K for Fe2Nb to 58 K. The mixed-bridged Fe2NbHCOO also showed a much broader magnetic hysteresis loop of 0.102 T, compared to 0.013 T for Fe2Nb.

1. Introduction

The rational design of crystalline molecular magnets enables the combination of magnetism (ferromagnetism) with other properties within a single material. The additional functionalities co-exist or are strongly coupled with the magnetic ordering, providing a convenient route to multifunctionality [1]. In terms of future technological applications, it is crucial to develop new strategies towards higher magnetic ordering temperatures in molecular magnets [2] to eliminate the need for expensive cooling resources like liquid helium. Efforts thus far have focused mainly on Prussian blue analogs [3] and cyanido-bridged materials based on hepta- or octacyanidometallates of 4d and 5d metal ions [4,5]. This is due to the ability of CN to efficiently mediate moderate-to-strong magnetic interactions between the bridged paramagnetic metal ions. Most of the room-temperature molecular magnets are based on VII/VIII and CrIII metal ions, where the magnetic coupling was found to be the strongest [6,7,8,9,10,11,12]. Unfortunately, these compounds are often obtained as amorphous powders and are air-sensitive, therefore, there is still a high demand for air-stable, crystalline, high-Tc, molecule-based magnets. One way to achieve this is changing the metal centers (spins and coupling constants). Here, a slightly different approach is proposed based on increasing the number of the closest magnetic neighbors. This approach is dictated by the molecular field theory Equation (1) [13], where the Tc depends on three factors: the coupling constant JAB between the two metal centers A and B, the spin values SA and SB and the number of the closest magnetic neighbors nA and nB (kB is the Boltzmann constant).
T c = 2 n A n B | J A B | S A ( S A + 1 ) S B ( S B + 1 ) 3 k B ,
The aim of our research was to modify the well-known compound {[FeII(H2O)2]2[NbIV(CN)8]·4H2O}n (Fe2Nb) [14] by inserting additional bridging ligands connecting the FeII ions in order to increase the number of the nearest magnetic neighbors, and re-inforce the existing long-range magnetic ordering. The implemented changes were similar to the modification of the MnII-NbIV analogue {[MnII(H2O)2]2[NbIV(CN)8]·4H2O}n, described previously [15]. The insertion of formate into the newly obtained mixed-bridged compound {(NH4)[(H2O)FeII-(μ-HCOO)-FeII(H2O)][NbIV(CN)8]·3H2O}n (Fe2NbHCOO) led to a very significant increase of the critical temperature Tc from 43 K (Fe2Nb parent framework) to 58 K, and a larger coercive field.

2. Results and Discussion

2.1. Synthesis and X-ray Crystal Structure Description

The insertion of additional formate anions into the three-dimensional (3D), CN-bridged parent framework of Fe2Nb at the self-assembly stage resulted in the formation of Fe2NbHCOO with mixed 3d-4d and 3d-3d metal bridging. The key to its formation was a large excess of ammonium formate in the solution of the building blocks: Mohr’s salt and potassium octacyanidoniobate(IV) dihydrate. The formation of the desired product was fully confirmed by single crystal X-ray diffraction structural analysis (sc-XRD; Figure 1 and Table 1), infrared spectroscopy (IR), elemental analysis (EA), and powder X-ray diffraction (PXRD).
Fe2NbHCOO is isostructural with its MnII analogue {(NH4)[(H2O)MnII-(μ-HCOO)-MnII(H2O)][NbIV(CN)8]·3H2O}n (Mn2NbHCOO) [15]. The coordination framework of Fe2NbHCOO becomes anionic compared to Fe2Nb, due to the insertion of negatively charged formate. As such, its electroneutrality is maintained by a simultaneous incorporation of an ammonium cation, which replaces one of the crystallization water molecules present in the parent compound.
Insertion of formate anions causes mild changes in the geometry of the CN-framework of Fe2NbHCOO, compared to the parent Fe2Nb. However, a noticeable change occurred between the adjacent FeII centers, where the formate ion was bound (Figure 1b), the FeII···FeII distance was shortened by 0.468 Å (from 6.609 Å to 6.141 Å). The corresponding change in the Mn2NbHCOO [15] compared to its MnII-NbIV parent [16] was smaller, shortening by only 0.337 Å (from 6.590 Å to 6.253 Å). A comparison of the most important distances and angles in Fe2NbHCOO and Fe2Nb is presented in Table 2. The cyanide bridges in Fe2NbHCOO were slightly less bent compared to Fe2Nb, while other structural parameters were quite similar (except for the aforementioned Fe···Fe distances).
In both Fe2NbHCOO and Fe2Nb, each [NbIV(CN)8]4− was connected to eight nearly octahedral [FeII(CN)4(H2O)L] moieties (L = H2O for Fe2Nb and L = formate for Fe2NbHCOO). The coordination geometry of NbIV in Fe2NbHCOO was slightly closer to an ideal square antiprism than in Fe2Nb, according to the continuous shape measure analysis executed using SHAPE software [17]. Conversely, the geometry of the octahedral FeII centers was much more distorted in Fe2NbHCOO due to the strain imposed by the bridging formate anion (Figure 1a,b and Table 3), which might lead to a completely different magnetic anisotropy.
The insertion of the additional formate bridge into the parent Fe2Nb framework changed the symmetry from a centrosymmetric I4/m to a non-centrosymmetric I4cm. The structure of Fe2NbHCOO was refined as an inversion twin with the following scales: 0.49(10) and 0.51(10). The asymmetric unit (Figure 2) was comprised of NbIV and FeII ions linked with a CN ligand (C2N2). The second CN ligand (C1N1) in the asymmetric unit connected the FeII ion with the next NbIV ion in the structure. The asymmetric unit also included an aqua ligand coordinated to FeII (O1) and half of the bridging formate anion (C3O2). There were also two oxygen atoms of the crystallization water molecules (O3 with occupancy 0.5 and O4 with occupancy 1.0), and a nitrogen atom of the ammonium cation (N3 with 0.5 occupancy in the same position as O3). H-atoms could not be located from the Fourier difference map. The presence of the ammonium cation was confirmed by IR spectroscopy (see below) and is consistent with the elemental analysis (EA) results.

2.2. IR Spectroscopy

The infrared (IR) spectrum of Fe2NbHCOO (Figure 3) exhibited bands typical for the parent framework: two bands characteristic for stretching (3576 cm−1) and bending (1635 cm−1) vibrations of O-H, and a strong band characteristic for stretching vibrations of CN- ligands (2141 cm−1). Additional specific bands as compared to the parent framework indicate the presence of the inserted formate anions (1584 cm−1, 1383 cm−1, 1365 cm−1 and 812 cm−1) and the ammonium cations (3261 cm−1 and 1403 cm−1).

2.3. Identity and Purity Confirmation by Powder X-ray Diffraction

The identity and purity of the bulk sample of Fe2NbHCOO was confirmed by powder X-ray diffraction measurements (Figure 4). The experimental PXRD diffraction patterns were in very good agreement with the simulated results from the sc-XRD structural model obtained at room temperature.

2.4. Magnetic Properties

The shape of the χT(T) dependence for Fe2NbHCOO recorded in the 2–300 K range at 0.1 T magnetic field (Figure 5) suggests the presence of a long-range magnetic ordering in this system below 70 K (discussed below). The experimental χT value at 300 K of 8.31 cm3·K·mol‒1 was slightly lower than the expected 9.02 cm3·K·mol‒1 for two FeII (S = 2), assuming gFe = 2.4, and one NbIV (S = ½), assuming gNb = 2.0. This was most probably caused by the presence of antiferromagnetic interactions within the NbIV-CN-FeII coordination framework. χT increased significantly as the temperature was lowered to reach a maximum value of 889 cm3·K·mol‒1 at 45 K, confirming the long-range magnetic ordering.
The magnetization vs. temperature M(T) curves recorded in the zfc-fc (zero field-cooled, field-cooled) modes at 0.3 mT for Fe2NbHCOO are presented in Figure 6. The most striking feature is the significant increase of the magnetization at 58 K, which constitutes the long-range magnetic ordering temperature Tc for this system. The critical temperature Tc was defined as the zfc-fc bifurcation point and was confirmed by the position of the maximum of the AC (alternating current) magnetic susceptibility signal χ’ at 57.5 K (Figure 7). There was essentially no frequency dependence of the in-phase and out-of-phase AC susceptibilities at the magnetic ordering temperature, which supports the claim that the reported compound is a magnetically ordered system. However, a slight frequency drift below 50 K indicates some dynamic processes taking place in the compound. Understanding this behavior will require further measurements using more advanced techniques (i.e., muon spin spectroscopy).
The magnetic field dependence of the molar magnetization for Fe2NbHCOO (recorded up to 7 T at T = 2.0 K) is presented in Figure 8. The magnetization curve showed a fast increase to circa 4.5 Nβ at 0.2 T, followed by a further slower increase and a saturation value of ca. 8.8 Nβ at 7.0 T. This was similar to the value obtained for the parent Fe2Nb [14], and close to the expected 8.6 Nβ for two high-spin FeII (S = 2, gFe = 2.4) antiferromagnetically coupled with NbIV ion (S = ½, gNb = 2.0). This suggests that Fe2NbHCOO exhibits long-range ferrimagnetic ordering. Fe2NbHCOO exhibited a magnetic hysteresis loop (Figure 8, inset) with the coercive field Hc = 102 mT and the remanence MR = 4.8 Nβ. Both values were much larger than the parent framework Fe2Nb (Hc = 13 mT and MR = 1.24 Nβ). The significant increase of the magnetic ordering temperature and a larger coercive field/remanence of Fe2NbHCOO compared to the parent Fe2Nb were most likely caused by the inserted bridging formate connecting the neighboring FeII ions. This formate bridge must promote the FeII ions magnetic anisotropy change and most probably additional local ferromagnetic interactions between them, which work in concert with the postulated ferrimagnetic structure of the CN-bridged framework (Figure 9). Nevertheless, an unlikely scenario where strong ferromagnetic ordering within the NbIV-CN-FeII framework [14] is accompanied by very weak antiferromagnetic interactions through FeII-HCOO-FeII pairs is also possible. Please note that the MnII-based analog {(NH4)[(H2O)MnII-(μ-OOCH)-MnII(H2O)][NbIV(CN)8]·3H2O}n (Mn2NbHCOO) that we recently reported [15] showed a significant lowering of the Tc from 49 K to 45 K compared to its {[MnII(H2O)2]2[NbIV(CN)8]·4H2O}n (Mn2Nb) parent framework, due to the antiferromagnetic interactions transmitted through the MnII-HCOO-MnII motifs.

3. Materials and Methods

3.1. Materials

Chemicals of analytical grade were purchased from commercial sources (Sigma-Aldrich Co., St. Louis, MO, USA) and used as received. K4[Nb(CN)8]·2H2O was synthesized according to the last reported procedure [18]. All operations were carried out in an ambient atmosphere.

3.2. Synthesis of {(NH4)[(H2O)FeII-(µ-HCOO)-FeII(H2O)][NbIV(CN)8]·3H2O}n (Fe2NbHCOO)

A water (160 mL) solution of HCOONH4 (1.26 g, 20.0 mmol) with a small amount of ascorbic acid (ca. 10–15 mg) was prepared. Half of the volume of this solution was used to dissolve K4[Nb(CN)8]∙2H2O (149.1 mg, 0.30 mmol), and the other half was used to dissolve (NH4)2Fe(SO4)2∙6H2O (235.3 mg, 0.60 mmol). After adding the greenish solution of Mohr’s salt to the yellow solution of potassium octacyanidoniobiate(IV), the resulting orange mixture was left for one day for crystallization. Dark purple crystals (60.2 mg, 35%) of Fe2NbOOCH were isolated by repeated decantation with distilled water followed by ethanol (92%), and were dried for a short while in air (the compound is slightly sensitive to oxygen and needs to be stored at low temperature). We found C, 18.81; H, 2.59; N, 21.40. C9H15Fe2N9NbO7 (565.87 g/mol) requires C, 19.10; H, 2.67; N, 22.28%. Slightly lower N content is caused by the air-sensitivity of the compound. IR (νmax/cm−1): 3576 s ν(OH), 3261 s ν(NH), 2141 s ν(CN), 1635 m δ(OH), 1584 vs ν(COO), 1403 ν(NH4+), 1383 m and 1365 ν(COO), and 812 ν(COO).

3.3. Single-Crystal X-ray Diffraction

The single crystal diffraction data for Fe2NbHCOO was collected on a Bruker D8 Quest Eco Photon50 CMOS machine equipped with a Mo Kα radiation source and a graphite monochromator (λ = 0.71073 Å). Measurements were performed at ambient temperature (details in Table 1). Data reduction and unit cell determination were carried out using SAINT and SADABS (Apex3 package). Absorption correction using the multi-scan method was applied for all reflection intensities. The structure was solved using direct methods (Apex3 package). Non-hydrogen atoms were refined anisotropically using weighted full-matrix least-squares on F2 [19,20]. The crystallographic data is summarized in Table 1. CCDC 1859396 (Fe2NbHCOO) contains the supplementary crystallographic data for this paper. This data can be obtained free of charge from the Cambridge Crystallographic Data Centre via ww.ccdc.cam.ac.uk/data_request/cif.

3.4. Magnetic Measurements

Magnetic measurements were performed using a Quantum Design MPMS-3 Evercool magnetometer equipped with a 7 T superconducting magnet. The sample was loaded into a double polypropylene bag and sealed. The magnetic susceptibility was corrected for the diamagnetic contribution of the sample holder and the diamagnetism of the samples themselves using Pascal constants.

3.5. Other Physical Measurements and Calculations

Infrared spectra were collected on Nicolet iS 5 FT-IR spectrometer (Thermo Fisher Scientific, Wltham, MA, USA) in the range 4000–650 cm−1. Powder X-ray diffraction experiments (PXRD) were carried out using PANalytical X’Pert Pro MPD diffractometer (Cu Kα radiation, Malvern PANalytical, Royston, UK) at ambient temperature for dry well-ground samples loaded into a narrow diameter borosilicate-glass capillary (0.7 mm). Elemental analysis was performed using an ELEMENTAR Vario Micro Cube CHNS analyzer (Elementar, Langenselbold, Germany). Continuous shape measure analysis for coordination spheres of NbIV and FeII was performed using the SHAPE software [21]. The results are summarized in Table 3.

4. Conclusions

A modification of the structure and magnetic properties of a parent coordination polymer Fe2Nb was performed at the self-assembly stage by “forcing” an additional bridging formate anion into its structure. The formate anion formed direct coordination connections with two adjacent FeII centers within the CN-bridged FeII-NbIV framework. The presence of this additional molecular bridge promoting ferromagnetic interactions between iron(II) centers reinforced the ferrimagnetic structure of the FeII-NbIV framework, and made Fe2NbHCOO a much better magnet than the parent Fe2Nb. Our approach demonstrates a new efficient route towards molecular magnets with higher critical temperatures.

Author Contributions

D.P. conceived and designed the experiments and performed the magnetic measurements; G.H. synthesized the compounds and performed the single crystal XRD structural analysis; and D.P. and G.H. analyzed the data and wrote the paper together.

Funding

This work was funded by the Polish National Science Centre within the Sonata Bis project NO. 2016/22/E/ST5/00055.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pinkowicz, D.; Czarnecki, B.; Reczyński, M.; Arczyński, M. Multifunctionality in Molecular Magnetism. Sci. Prog. 2015, 98, 346–378. [Google Scholar] [CrossRef] [PubMed]
  2. Miller, J.S.; Ohkoshi, S. High-Tc Ordered Molecular Magnets. In Molecular Magentic Materials. Concepts and Applications, 1st ed.; Sieklucka, B., Pinkowicz, D., Eds.; Wiley: Weinheim, Germany, 2017; pp. 161–186. ISBN 978-3-527-33953-2. [Google Scholar]
  3. Tokoro, H.; Ohkoshi, S. Novel magnetic functionalities of Prussian blue analogs. Dalton Trans. 2011, 40, 6825–6833. [Google Scholar] [CrossRef] [PubMed]
  4. Nowicka, B.; Korzeniak, T.; Stefańczyk, O.; Pinkowicz, D.; Chorąży, S.; Podgajny, R.; Sieklucka, B. The impact of ligands upon topology and functionality of octacyanidometallate-based assemblies. Coord. Chem. Rev. 2012, 256, 1946–1971. [Google Scholar] [CrossRef]
  5. Wang, X.-Y.; Avendaño, C.; Dunbar, K.R. Malecular magnetic materials based on 4d and 5d transition metals. Chem. Soc. Rev. 2011, 40, 3213–3238. [Google Scholar] [CrossRef] [PubMed]
  6. Entley, W.R.; Girolami, G.S. High-temperature molecular magnets based on cyanovanadate building blocks: Spontaneous magnetization at 230 K. Science 1995, 268, 397–402. [Google Scholar] [CrossRef] [PubMed]
  7. Imoto, K.; Takemura, M.; Tokoro, H.; Ohkoshi, S. A Cyano-Bridged Vanadium-Niobium Bimetal Assembly Exhibiting a High Curie Temperature of 210 K. Eur. J. Inorg. Chem. 2012, 2649–2652. [Google Scholar] [CrossRef]
  8. Mizuno, M.; Ohkoshi, S.; Hashimoto, K. Electrochemical Synthesis of High-Tc, Colored, Magnetic Thin Films Composed of Vanadium(II/III)-Chromium(II) Hexacyanochromate(III). Adv. Mater. 2000, 12, 1955–1958. [Google Scholar] [CrossRef]
  9. Ohkoshi, S.; Mizuno, M.; Hung, G.J.; Hashimoto, K. Magnetooptical Effects of Room Temperature Molecular-Based Magnetic Films Composed of Vanadium Hexacyanochromates. J. Phys. Chem. B 2000, 104, 9365–9367. [Google Scholar] [CrossRef]
  10. Ferlay, S.; Mallah, T.; Ouahés, R.; Veillet, P.; Verdaguer, M. A room-temperature organometallic magnet based on Prossian blue. Nature 1995, 378, 701–703. [Google Scholar] [CrossRef]
  11. Holmes, S.M.; Girolami, G.S. Sol-Gel Synthesis of KVII[CrIII(CN)6]·2H2O: A Crystalline Molecule-Based Magnet with a Magnetic Ordering Temperature above 100 °C. J. Am. Chem. Soc. 1999, 121, 5593–5594. [Google Scholar] [CrossRef]
  12. Hatlevik, Ø.; Buschmann, W.E.; Zhang, J.; Manson, J.L.; Miller, J.S. Enhancement of the Magnetic Ordering Temperature and Air Stability of a Mixed Valent Vanadium Hexacyanochromate(III) Magnet to 99 °C (372 K). Adv. Mater. 1999, 11, 914–918. [Google Scholar] [CrossRef]
  13. Ohkoshi, S.-I.; Iyoda, T.; Fujishima, A.; Hashimoto, K. Magnetic properties of mixed ferro-ferrimagnets composed of Prussian blue analogs. Phys. Rev. B 1997, 56, 11642–11652. [Google Scholar] [CrossRef]
  14. Pinkowicz, D.; Podgajny, R.; Pełka, R.; Nitek, W.; Bałanda, M.; Makarewicz, M.; Czapla, M.; Żukowski, J.; Kapusta, C.; Zając, D.; et al. Iron(II)-octacyanoniobate(IV) ferromagnet with TC 43 K. Dalton Trans. 2009, 7771–7777. [Google Scholar] [CrossRef] [PubMed]
  15. Handzlik, G.; Sieklucka, B.; Pinkowicz, D. Cross-linking of cyanide magnetic coordination polymers by rational insertion of formate, cyanide or azide. Dalton Trans. 2018. [Google Scholar] [CrossRef] [PubMed]
  16. Herrera, J.M.; Franz, P.; Podgajny, R.; Pilkington, M.; Biner, M.; Decurtins, S.; Stoeckli-Evans, H.; Neels, A.; Grade, R.; Dromzee, Y.; et al. Three-dimensional bimetallic octacyanidometalates [MIV{(μ-CN)4MnII(H2O)2}2·4H2O]n (M=Nb, Mo, W): Synthesis, single-crystal X-ray diffraction and magnetism. C. R. Chim. 2008, 11, 1192. [Google Scholar] [CrossRef]
  17. Alvarez, S.; Alemany, P.; Casanova, D.; Cirera, J.; Llunell, M.; Avnir, D. Shape maps and polyhedral interconversion paths in transition metal chemistry. Coord. Chem. Rev. 2005, 249, 1693–1708. [Google Scholar] [CrossRef]
  18. Handzlik, G.; Magott, M.; Sieklucka, B.; Pinkowicz, D. Alternative Synthetic Route to Potassium Octacyanidoniobate(IV) and Its Molybdenum Congener. Eur. J. Inorg. Chem. 2016, 30, 4872–4877. [Google Scholar] [CrossRef]
  19. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A: Found. Crystallogr. 2008, 64, 112–122. [Google Scholar] [CrossRef] [PubMed]
  20. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C: Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  21. Llunell, M.; Casanova, D.; Cirera, J.; Alemany, P.; Alvarez, S. SHAPE v. 2.1; University of Barcelona: Barcelona, Spain, 2013. [Google Scholar]
Figure 1. Structural diagrams showing the potential “cavity” between adjacent FeII ions in the parent framework Fe2Nb (a), the additional formate bridge occupying the “cavity” in Fe2NbHCOO (b), a slice of Fe2NbHCOO crystal packing parallel to the ab crystallographic plane (c), and a packing diagram presenting the three-dimensional (3D) CN-bridged coordination framework (red) cross-linked by local formate bridges (green) (d). Note: Nb—cyan, Fe—yellow, C—gray, N—blue, and Oformate—red. Hydrogen and oxygen atoms of water molecules have been omitted for the sake of clarity.
Figure 1. Structural diagrams showing the potential “cavity” between adjacent FeII ions in the parent framework Fe2Nb (a), the additional formate bridge occupying the “cavity” in Fe2NbHCOO (b), a slice of Fe2NbHCOO crystal packing parallel to the ab crystallographic plane (c), and a packing diagram presenting the three-dimensional (3D) CN-bridged coordination framework (red) cross-linked by local formate bridges (green) (d). Note: Nb—cyan, Fe—yellow, C—gray, N—blue, and Oformate—red. Hydrogen and oxygen atoms of water molecules have been omitted for the sake of clarity.
Magnetochemistry 04 00041 g001
Figure 2. Asymmetric unit of Fe2NbHCOO.
Figure 2. Asymmetric unit of Fe2NbHCOO.
Magnetochemistry 04 00041 g002
Figure 3. IR spectrum of Fe2NbHCOO.
Figure 3. IR spectrum of Fe2NbHCOO.
Magnetochemistry 04 00041 g003
Figure 4. Experimental (red) and simulated from the single-crystal structural model (blue) powder X-ray diffraction patterns for Fe2NbHCOO (the broad peak at 8° is from the glass capillary—the sample holder).
Figure 4. Experimental (red) and simulated from the single-crystal structural model (blue) powder X-ray diffraction patterns for Fe2NbHCOO (the broad peak at 8° is from the glass capillary—the sample holder).
Magnetochemistry 04 00041 g004
Figure 5. Temperature dependence of the molar magnetic susceptibility and temperature product for Fe2NbHCOO at H = 0.1 T.
Figure 5. Temperature dependence of the molar magnetic susceptibility and temperature product for Fe2NbHCOO at H = 0.1 T.
Magnetochemistry 04 00041 g005
Figure 6. Temperature dependence of the molar magnetization measured in the zfc-fc (red and black points, respectively) modes for Fe2NbHCOO.
Figure 6. Temperature dependence of the molar magnetization measured in the zfc-fc (red and black points, respectively) modes for Fe2NbHCOO.
Magnetochemistry 04 00041 g006
Figure 7. Temperature dependence of the in-phase and out-of-phase AC magnetic susceptibility of Fe2NbHCOO at HAC = 0.1 mT and at three different frequencies (7, 70 and 700 Hz).
Figure 7. Temperature dependence of the in-phase and out-of-phase AC magnetic susceptibility of Fe2NbHCOO at HAC = 0.1 mT and at three different frequencies (7, 70 and 700 Hz).
Magnetochemistry 04 00041 g007
Figure 8. The magnetic field dependence of the molar magnetization of Fe2NbHCOO recorded at 2.0 K in the 0–7 T range and the magnetic hysteresis loop at the same temperature (inset).
Figure 8. The magnetic field dependence of the molar magnetization of Fe2NbHCOO recorded at 2.0 K in the 0–7 T range and the magnetic hysteresis loop at the same temperature (inset).
Magnetochemistry 04 00041 g008
Figure 9. Schematic representation of the postulated ferrimagnetic structure of Fe2NbHCOO with an antiparallel alignment of the NbIV and FeII magnetic moments. The ferromagnetic interactions transmitted through the newly introduced FeII-HCOO-FeII structural motifs work in concert with the antiferromagnetic coupling within the -NbIV-CN-FeII- 3D skeleton.
Figure 9. Schematic representation of the postulated ferrimagnetic structure of Fe2NbHCOO with an antiparallel alignment of the NbIV and FeII magnetic moments. The ferromagnetic interactions transmitted through the newly introduced FeII-HCOO-FeII structural motifs work in concert with the antiferromagnetic coupling within the -NbIV-CN-FeII- 3D skeleton.
Magnetochemistry 04 00041 g009
Table 1. Sc-XRD structural, solution and refinement parameters for Fe2NbHCOO.
Table 1. Sc-XRD structural, solution and refinement parameters for Fe2NbHCOO.
FormulaC9HFe2N9NbO7
Temperature, K296(2)
λ, Å0.71073 Å
Molecular weight, g/mol551.80
Crystallographic systemtetragonal
Space groupI4cm
Unit cella = 11.8782(12) Å
b = 13.2770(14) Å
Volume V, Å31873.3(4)
Z4
Density ρcalc, g/cm31.957
F (000)1068
θ, deg3.4–28.0
Abs. coeff. µ, mm−12.18
Data/parameters /restrains1211/81/8
R [F2 > 2σ(F2)]0.032
wR (F2)0.064
GOF on F21.103
Flack parameter0.472(25)
max/min resid. density, e· Å−31.13/−0.59
Reflections collected13274
Unique reflections1211
Rint0.048
Completeness, %99.6
Table 2. Comparison of selected distances (Å) and angles (°) in the parent framework and modified Fe2NbHCOO.
Table 2. Comparison of selected distances (Å) and angles (°) in the parent framework and modified Fe2NbHCOO.
Atom NamesFe2NbFe2NbHCOO
Nb···Fe5.409
5.508
5.461
5.480
Nb···Feav5.4595.471
Fe···Fe6.609(1)6.141(1)
Fe-NCN2.144(2)
2.175(2)
2.119(10)
2.171(10)
Fe-N-C154.5(2)
167.5(2)
157.8(11)
170.0(9)
Nb-C-N174.6(2)
176.7(2)
173.0(10)
178.4(11)
Fe-Oaq2.099(3)
2.130(3)
2.191(9)
Fe-OHCOO-1.977(9)
Table 3. Analysis of the coordination spheres of FeII and NbIV ions in Fe2Nb and Fe2NbHCOO.
Table 3. Analysis of the coordination spheres of FeII and NbIV ions in Fe2Nb and Fe2NbHCOO.
NbIVSquare antiprism (SAPR-8)0.1920.113
FeIIOctahedron (OC-6)0.1610.562

Share and Cite

MDPI and ACS Style

Handzlik, G.; Pinkowicz, D. How to Make a Better Magnet? Insertion of Additional Bridging Ligands into a Magnetic Coordination Polymer. Magnetochemistry 2018, 4, 41. https://doi.org/10.3390/magnetochemistry4030041

AMA Style

Handzlik G, Pinkowicz D. How to Make a Better Magnet? Insertion of Additional Bridging Ligands into a Magnetic Coordination Polymer. Magnetochemistry. 2018; 4(3):41. https://doi.org/10.3390/magnetochemistry4030041

Chicago/Turabian Style

Handzlik, Gabriela, and Dawid Pinkowicz. 2018. "How to Make a Better Magnet? Insertion of Additional Bridging Ligands into a Magnetic Coordination Polymer" Magnetochemistry 4, no. 3: 41. https://doi.org/10.3390/magnetochemistry4030041

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