Next Article in Journal
New Mixed Y0.5R0.5VO4 and RVO4:Bi Materials: Synthesis, Crystal Structure and Some Luminescence Properties
Next Article in Special Issue
High-Throughput Synthesis of Pillared-Layered Magnesium Tetraphosphonate Coordination Polymers: Framework Interconversions and Proton Conductivity Studies
Previous Article in Journal / Special Issue
Manganese Fluorene Phosphonates: Formation of Isolated Chains
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One-Dimensional Fluorene-Based Co(II) Phosphonate Co(H2O)2PO3C–C12H9·H2O: Structure and Magnetism

1
Normandie Univ, ENSICAEN, UNICAEN, CNRS, CRISMAT, 14000 Caen, France
2
IPCMS, UMR Unistra-CNRS 7504, 23 rue du Loess, BP 43, 67034 Strasbourg CEDEX 2, France
3
CEMCA UMR CNRS 6521, Université de Brest, IBSAM, 6 Avenue Victor Le Gorgeu, 29238 Brest, France
*
Authors to whom correspondence should be addressed.
Inorganics 2018, 6(3), 93; https://doi.org/10.3390/inorganics6030093
Submission received: 4 July 2018 / Revised: 31 August 2018 / Accepted: 3 September 2018 / Published: 5 September 2018
(This article belongs to the Special Issue Metal Phosphates and Phosphonates)

Abstract

:
A new Co(II) phosphonate, Co(H2O)2PO3C–C12H9·H2O, has been synthesized under hydrothermal conditions. The monoclinic P21/c structure of this organic–inorganic hybrid consists of isolated perovskite-type chains of corner-shared CoO4(H2O)2 octahedra interconnected via phosphonate groups. The unique one-dimensional structure of this phase is closely related to the single-chain magnet (SCM) phosphonate Co(H2L)(H2O), with L = 4-Me-C6H4-CH2N(CPO3H2)2, that contains isolated chains of CoO5N octahedra. Like the latter, this hybrid exhibits 1D antiferromagnetic interactions and the possibility of an effective pseudo spin contribution due to spin canting at low temperature, but, in contrast, is not an SCM. This different magnetic behavior is explained by the different geometry of the octahedral chains and by the possible existence of weak antiferromagnetic interactions between the chains. This opens the route to the investigation of a large series of compounds by tuning the chemical composition and structure of the phosphonic acid used as organic precursor of hybrid materials.

Graphical Abstract

1. Introduction

Among the hybrid materials, metal phosphonates represent a very large family that has been the object of numerous investigations [1] in view of various applications such as ion exchange properties [2], catalysis [3,4,5,6,7], luminescence [8,9,10], protonic conduction [11,12,13,14], or as a reservoir of bio-active metal ions [15,16]. Transition metal phosphonates have also been investigated for their attractive magnetic properties [17]. This is the case for Co(II) phosphonates, which can generate specific low-dimensional magnetic behavior, in connection with the high magnetic anisotropy of octahedral cobalt due to its spin–orbit coupling [18]. This is exemplified by several one-dimensional cobalt phosphonates with a ladder-like chain structure [19,20,21,22,23,24,25]. In those hybrids, the chains consist of isolated “Co2” dimers of edge-shared CoO6 octahedra connected through O–P–O bridges, causing them to be described by different authors as 1D antiferromagnets, with properties closely related to those of spin ladder systems. One-dimensional cobalt phosphonates with single chains built up of isolated CoO5N octahedra or CoO3N2 trigonal bipyramids bridged by PO3C tetrahedra were also obtained, showing weak magnetic interactions [26,27,28]. In contrast, the possibility to synthesize hybrid phosphonates involving isolated infinite chains of corner-shared cobalt CoX6 octahedra is quite rare. The compound Co(H2L)(H2O), where L = 4-Me-C6H4-CH2N(CPO3H2)2, which consists of CoO5N octahedra, is to our knowledge the only cobalt phosphonate to date that exhibits such a structural behavior [29,30]. Moreover, it has been shown to be a single-chain magnet (SCM), a magnetic behavior which has been the object of numerous investigations these last fifteen years [31,32,33,34,35,36] in view of its potential application in high-density data storage and quantum computing [37,38].
Based on these results, we believe that the investigation of one-dimensional Co(II) phosphonates is a challenging direction for the research of new magnetic properties. Bearing in mind that large molecular ligands in the phosphonate groups should allow the formation of isolated cobalt chains and control of the distances between them, we have thus embarked on the exploration of cobalt phosphonates prepared from 9-fluorenylphosphonic acid. We report herein a new phosphonate Co(H2O)2PO3C–C12H9·H2O, whose structure, built up of CoO4(H2O)2 octahedra, consists of isolated infinite perovskite-type Co(II) chains. We show that the latter exhibits 1D antiferromagnetic interactions and intra-chain spin canting, but, in spite of its structure closely related to that of Co(H2L)(H2O) [29,30], is not an SCM.

2. Results and Discussion

The phosphonate Co(H2O)2PO3C–C12H9·H2O was synthesized under hydrothermal conditions from cobalt(II) acetate Co(CH3CO2)2·4H2O, urea, and 9-fluorenylphosphonic acid (C13H11O3P) (Scheme 1a). The use of urea in the hydrothermal synthesis allows slow basification of the pH media during the hydrolysis process of urea leading to an increase of the pH and to the fully deprotonated phosphonate function. The thermogravimetric analysis (TGA) and CHNS elemental analysis allowed the chemical formula of this compound to be established. Oblong-shaped crystallites (Figure 1) were identified in the homogenous batches of this compound by scanning electron microscopy (SEM), with an average length from 28 µm to 54.8 µm, width from 6 µm to 11.2 µm, and thickness of 0.8 µm. Energy-dispersive X-ray spectrometry (EDS) analysis, coupled with SEM, confirmed the result obtained by TGA, i.e., the molar ratio Co/P = 1.

2.1. Structural Study

Data were corrected from absorption using the Twinabs program [39] developed for scaling and corrections for twin and area detector data. The structure was determined using SIR2011 [40] and refined using the program Jana 2006 [41]; the analysis of the electronic residue allows the location of all the missing atomic sites. Harmonic atomic displacement parameters were considered for Co, P, O, and C species. Hydrogen atoms were geometrically added; distance and angle restraints were introduced for H belonging to water molecules and these H were refined. Twin laws were introduced for the monoclinic phase and the ratio of the possible twin domains was refined. All the details of the refinement are summarized in Table 1. Atomic positions are listed in Table S1, thermal parameters in Table S2, bond angles in Table S3, and interatomic distances in Table S4. In order to check that the crystal is representative of the whole powder, a study by powder X-ray diffraction was performed. After analysis with FullProf software [1], a simulation of the pattern was realized. All the peaks were indexed and were shown to match with the P21/c space group and lattice parameters found by single crystal X-ray diffraction (Figure S1, Supplementary Materials).
The view of the structure along b (Figure 2) shows that it consists of chains of CoO4(H2O)2 octahedra running along that direction. Two PO3C–C12H9 groups are grafted on each cobalt chain in “trans” positions with respect to the chain, so that the cobalt chains and the fluorene groups form layers parallel to the (100) plane. Thus, in this structure, one single inorganic layer of cobalt chains alternates with a double organic layer of fluorene groups along a. Moreover, additional H2O molecules are located in the inorganic layers between the cobalt chains. The arrangement of the cobalt polyhedra and PO3C groups is better understood by considering the structure close to the a direction and by taking out the fluorene groups (Figure 3a). One observes that the cobalt [010] zig-zag chains consist of corner-sharing CoO4(H2O)2 octahedra with the perovskite-like geometry and can be formulated as [CoO3(H2O)2]. Each cobalt octahedron has its two free apical corners occupied by H2O and shares two of its four equatorial O apices with the neighboring octahedra of the chain. The two remaining O apices of each octahedron are shared with the PO3C tetrahedron linked to a fluorene group. In fact, each PO3C group shares two corners with two CoO4(H2O)2 octahedra and one edge with another octahedron. The arrangement of the cobalt chains in the form of (100) layers can also be seen from the projection of the structure along c (Figure 4). This view shows that the fluorene groups “C12H19” are directed along two transversal directions oriented at 40.13° and 42.93° with respect to the inorganic layers. In one column, the fluorene moieties, with distance of 3.58 Å, are stacked along c in a face-to-face packing (Figure 2b). The interatomic Co–O and P–O distances (Table S4) are in agreement with the iono-covalent radii given by Shannon and Prewitt [42] and confirm the strong bonding of the structure along b. In contrast, laterally, within the inorganic layers, the linking between the chains along c is only ensured by hydrogen bonds between H2O molecules and O atoms of the cobalt octahedral coordination environment and PO3C tetrahedra (Figure 2a), whereas the cohesion of the structure between the layers, along a, is ensured by van der Waals interactions between the fluorene groups (orange dashed lines, Figure 4). The study of the Hirshfeld surface [43,44] and its corresponding two-dimensional fingerprint plots [45] of Co(H2O)2PO3C–C12H9·H2O are in good agreement with the above statements (Figure S2, Supplementary Materials). Indeed, the most numerous intermolecular interactions are the C···H contacts (38.6%) and H···H contacts (51.1%), which evidence for the main part C–H···π interactions between the fluorene moieties (Figure S2). These π···π interactions are known to play in important role in the cohesion of many aromatic phosphonate-containing materials structures [46,47,48]. The other contact types’ contributions (P···C, C···O, C···C, and O···H) represent 10.3%.
The structure of this phosphonate-based hybrid is very closely related to that of Co(H2L)(H2O), where L = 4-Me-C6H4-CH2N(CPO3H2)2 (Scheme 1b), previously described by Palii et al. [1,30]. The latter consists indeed of rather similar isolated chains of corner-sharing Co(II) octahedra (Figure 3b) interconnected through chains of phosphonate groups. However, the two types of chains differ by their geometry. In the present hybrid, the perovskite-type chains form Co–O–Co bond angles close to 180° (Figure 3a), whereas in the zig-zag chains of the Co(H2L)(H2O) compound (Figure 3b), two successive Co(II) octahedra are strongly tilted with respect to each other so that the Co–O–Co bond angles deviate largely from 180° (i.e., 122.09°). It will be shown further that such a different geometry may influence the magnetic properties of this phase.

2.2. Magnetic Properties

The temperature dependence of the molar magnetic susceptibility χ is presented Figure 5. The inverse of the susceptibility curve (inset Figure 5), shows a linear behavior between 200 K and 300 K. The latter can be fit above 200 K using the Curie–Weiss law χ = C/(T − θ) which provides the Curie constant C = 3.14 emu·K·mol−1, in agreement with the expected value from the literature for six-coordinated “high-spin” Co(II) (C = 2.8–3.4 emu·K·mol−1) [49]. The experimental magnetic moment deduced from the C value, µeff = 5.01 µB (where µB is the Bohr magneton), is close to the calculated value of Co(II), µeff = 5.20 µB, which has [Ar]3d7 electronic configuration, implying a total spin S value of 3/2 and a total angular momentum L of 3. Note that the typical magnetic moments observed for octahedral complexes range from 4.7 to 5.2 µB [49]. The Weiss temperature, θ = −9.93 K, suggests the presence of weak antiferromagnetic exchange interactions and/or spin–orbit coupling. This feature is corroborated by the variation of the product χT as a function of temperature, which shows a continuous decrease from 3.00 emu·K·mol−1 at 300 K to 0.97 emu·K·mol−1 at 5 K (Figure 5). The existence of such antiferromagnetic interactions is in agreement with the Kanamori–Goodenough rules for a 180° super exchange mechanism between Co(II) [50,51].
When decreasing the temperature below 5 K, a small increase of χT can be observed, indicating some other magnetic contribution. This increase is observed in samples from different batches, indicating a likely intrinsic origin. In addition, the measurement of magnetization versus magnetic field at T = 2 K shows an S shape typical of spin canting (Figure 6).
This Ising-like behaviour of the chains with spin canting at low temperature is similar to that of Co(H2L)(H2O) [30]. Yet, contrarily to what is described for Co(H2L)(H2O), we do not observe here any hysteresis in the M = f(H) curve. Moreover, measurement of alternating current (ac) susceptibility indicates that no slow relaxation process occurs within the temperature range 2 K–6 K and frequency range 0.1 Hz–1 kHz under a zero direct current (dc) field (Figure S3, Supplementary Materials). Note that the absence of single-chain magnetism was also emphasized in the one-dimensional Co(II) hydroxide [Co4(phcina)6(OH)2)(H2O)4]·2H2O by Oka et al. [52] where single octahedral diamond chains are isolated from each other by 17 Å.

3. Experimental Section

Thermogravimetric analysis (TGA) was performed on polycrystalline samples using a SETARAM TAG 92 apparatus (SETERAM Instrumentation, Caluire, France) under an air atmosphere at the heating rate of 3 °C per minute from room temperature to 1000 °C. Scanning Electron Microscopy (SEM) characterization was performed with a Carl ZEISS SUPRA 55 (Carl Zeiss Microscopy GmbH, Jena, Germany) on a raw sample with gold metallization. Elemental analysis and atomic absorption testing were performed on a ThermoQuest NA2500 (THERMO FINNIGAN, Villebon sur Yvette, France) and on an Agilent Technologies 200 Series AA setup (Agilent Technologies, Les Ulis, France), respectively. Energy-dispersive X-ray spectrometry (EDS) analysis was performed with an EDAX analyzer (AMETEK Materials Analysis Division, Mahwah, NJ, USA). The magnetic susceptibility was recorded versus temperature with a SQUID Quantum Design magnetometer (Quantum Design, San Diego, CA, USA) operating in the temperature range from 2 K to 300 K according to Zero Field Cooling (ZFC) and Field Cooling (FC) procedures. The applied magnetic field was 0.1 T. For the alternated current (ac) susceptibility measurements, measurement was performed from 2 K to 6 K under the ac field of 0.05 T (linear regime) and the frequency range 0.1–1 KHz. A suitable single crystal of Co(H2O)2PO3C–C12H9·H2O was selected and an X-ray diffraction experiment was performed at room temperature using Mo Kα radiation produced with a microfocus Incoatec Iμ sealed X-ray tube on a Kappa CCD diffractometer (Bruker, Billerica, MA., USA) equipped with an Apex2 CCD detector. Details of the data collection are summarized in Table 1 and in Tables S1–S4. On experimental frames for the highest θ area, a splitting of the reflections into four close spots is observed; the coupling between Kα1 and Kα2 radiations and the twin leads to a quadruplet of reflections typical for a twinning by reticular pseudo-merohedry. Plots of reciprocal lattice planes were assembled from the different scans; they provide an overall view of the reciprocal space useful both for the symmetry and for twin law analysis. The twin law compatible with the experiment is a π rotation around the c axis. Moreover, the conditions limiting the possible reflections are in agreement with the P21/c space group (h0l:l = 2n, 0k0:k = 2n). CrystalExplorer software [53] was used to generate the Hirshfeld surface (HS) and the corresponding 2D fingerprint plots. The HSs were calculated on the fluorene moieties of the cobalt phosphonate in very high resolution with a standard void cluster mode (unit cell + 5 Å) and an isovalue of 0.002 e·au−3.
Crystallographic data have been deposited into the Cambridge Crystallographic Data Centre, with deposition numbers CCDC 1853487 for Co(H2O)2PO3C–C12H9·H2O. These data can be obtained free of charge via www.ccdc.cam.ac.uk/conts/retrieving.html (or from the Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: (+44) 1223-336-033; or [email protected]).

3.1. Synthesis

All reagents were purchased from Sigma Aldrich (Saint-Quentin Fallavier, France) and Epsilon Chimie (Guipavas, France) and were used without prior purification in a hydrothermal synthesis process according to the following procedure: The Co(H2O)2(PO3C–C12H9)·H2O phosphonate (CoPO6C13H15, M = 357.20 g/mol) was synthesized from an equimolar mixture of cobalt acetate tetrahydrated Co(C2H3O2)2·4H2O (0.0506 g, 0.2 mmol), 9-fluorenyl-phosphonic acid C13H11O3P (0.0502 g, 0.2 mmol) (Scheme 1a), and urea N2H4CO (0.0121 g, 0.2 mmol) dissolved in distillated water (15 mL) in a 20 mL polytetrafluoroethylene (PTFE) liner. Then, the liner was inserted in a Berghof DAB-2 digestive vessel and heated from room temperature to 140 °C over 20 h, maintained at 140 °C for 30 h, and finally allowed to cool to room temperature within 20 h. After filtration, the final product, obtained as pink oblong-shaped crystals, was washed with distilled water, rinsed with ethanol, and then dried in air. The medium pH values before and after the hydrothermal process were 5.04 and 6.18, respectively. The chemical formula was confirmed by elemental analysis (found (calc.) C 43.26 (43.72), H 4.61 (4.23)) and by atomic absorption spectroscopy (found (theo.): M(Co) (g/mol) = 58.94 (58.93)).

3.2. Thermogravimetric Analysis

The thermogravimetric analysis (TGA) curve recorded on a polycrystalline sample (Figure 7) shows two weight losses between room temperature and 300 °C which are in good agreement with the departure of the three water molecules per Co(H2O)2PO3C–C12H9·H2O formula (theoretical percentage of water: 15.13%). The dehydration process occurs in two steps: A first weight loss of 10.26% (theoretical value 10.1%) corresponding to the departure of two H2O molecules between room temperature and 110 °C and a second weight loss of 4.85% from 110 °C to 300 °C corresponding to the departure of the last H2O molecule (theoretical value 5%). Beyond 300 °C, the next weight losses are assigned to the ligand decomposition and to the transformation of the dehydrated compound into cobalt pyrophosphate Co2P2O7. The composition and structure of the latter phase was confirmed by powder X-ray diffraction on the TGA residue (Figure S4, Supplementary Materials) and is in agreement with the total weight loss recorded by TGA (experimental value: 58.55%, theoretical value: 59.15%).

4. Conclusions

A unique one-dimensional cobalt phosphonate that exhibits isolated perovskite-type chains of corner-sharing CoO4(H2O)2 octahedra was synthesized. The structure of this phosphonate is closely related to that of the compound Co(H2L)(H2O), with L = 4-Me-C6H4-CH2N(CPO3H2)2, which similarly consists of chains of corner-sharing Co5N octahedra. The magnetic behavior of this phase shows that, like the latter, it exhibits 1D antiferromagnetic interaction with possible spin canting at low temperature, but, in contrast, is not an SCM. This difference can be attributed to the different geometries of the octahedral chain in the two compounds and to the fact that weak antiferromagnetic interactions may appear between the chains in the present phosphonate. This study should foster further investigations of new 1D cobalt phosphonates, varying the nature of organic ligands in order to understand the magnetism of these compounds.

Supplementary Materials

The following are available online at https://www.mdpi.com/2304-6740/6/3/93/s1, Figure S1: X-ray diffraction data recorded on powder from Co(H2O)2PO3C–C12H9·H2O. Figure S2: (a) Hirshfeld Surface represented on the fluorene moiety of Co(H2O)2PO3C–C12H9·H2O with a shape index mapping; (b) Representation of the corresponding two-dimensional fingerprint plot showing the contribution of each contact-type in percentage. Figure S3: Representation of the AC susceptibility of Co(H2O)2PO3C–C12H9·H2O. Figure S4: X-ray diffraction data recorded on the TGA residue of Co(H2O)2PO3C–C12H9·H2O. Tables S1–S4: Atomic positions, thermal parameters, bond angles, and interatomic distances of Co(H2O)2PO3C–C12H9·H2O, Cif and Check cif files of Co(H2O)2PO3C–C12H9·H2O.

Author Contributions

Conceptualization, J.-M.R., C.B., and B.R.; Validation, all authors; Formal Analysis, A.P., O.P., C.B., V.C., J.-M.R.; Writing—Original Draft Preparation, B.R., C.B.; Writing—Review & Editing, all authors equally contributed; Visualization, C.B.; Project Administration, G.R., P.-A.J., J.-M.R.; Funding Acquisition, G.R., P.-A.J., J.-M.R.

Funding

This research was funded by the Agence Nationale de la Recherche, grant number [ANR-14-CE07-0004-01 (HYMN)].

Acknowledgments

We thank the Agence Nationale de la Recherche (Contract No ANR-14-CE07-0004-01 (HYMN)) for financial support. The authors also express their grateful acknowledgment of technical support from Sylvie Collin and Fabien Veillon from the CRISMAT laboratory.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Clearfield, A.; Demadis, K. Metal Phosphonate Chemistry: From Synthesis to Applications; Royal Society of Chemistry: Cambridge, UK, 2011; ISBN 978-1-84973-356-4. [Google Scholar]
  2. Silbernagel, R.; Martin, C.H.; Clearfield, A. Zirconium(IV) Phosphonate-Phosphates as Efficient Ion-Exchange Materials. Inorg. Chem. 2016, 55, 1651–1656. [Google Scholar] [CrossRef] [PubMed]
  3. Song, J.; Zhou, B.; Zhou, H.; Wu, L.; Meng, Q.; Liu, Z.; Han, B. Porous Zirconium–Phytic Acid Hybrid: A Highly Efficient Catalyst for Meerwein–Ponndorf–Verley Reductions. Angew. Chem. Int. Ed. 2015, 54, 9399–9403. [Google Scholar] [CrossRef] [PubMed]
  4. Hu, A.; Yee, G.T.; Lin, W. Magnetically Recoverable Chiral Catalysts Immobilized on Magnetite Nanoparticles for Asymmetric Hydrogenation of Aromatic Ketones. J. Am. Chem. Soc. 2005, 127, 12486–12487. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, Z.; Heising, J.M.; Clearfield, A. Sulfonated Microporous Organic−Inorganic Hybrids as Strong Bronsted Acids1. J. Am. Chem. Soc. 2003, 125, 10375–10383. [Google Scholar] [CrossRef] [PubMed]
  6. Hu, A.; Ngo, H.L.; Lin, W. Chiral Porous Hybrid Solids for Practical Heterogeneous Asymmetric Hydrogenation of Aromatic Ketones. J. Am. Chem. Soc. 2003, 125, 11490–11491. [Google Scholar] [CrossRef] [PubMed]
  7. Maillet, C.; Janvier, P.; Pipelier, M.; Praveen, T.; Andres, Y.; Bujoli, B. Hybrid Materials for Catalysis? Design of New Phosphonate-Based Supported Catalysts for the Hydrogenation of Ketones under Hydrogen Pressure. Chem. Mater. 2001, 13, 2879–2884. [Google Scholar] [CrossRef]
  8. Mao, J.-G. Structures and luminescent properties of lanthanide phosphonates. Coord. Chem. Rev. 2007, 251, 1493–1520. [Google Scholar] [CrossRef]
  9. Rueff, J.-M.; Barrier, N.; Boudin, S.; Dorcet, V.; Caignaert, V.; Boullay, P.; Hix, G.B.; Jaffrès, P.-A. Remarkable thermal stability of Eu(4-phosphonobenzoate): Structure investigations and luminescence properties. Dalton Trans. 2009, 10614–10620. [Google Scholar] [CrossRef] [PubMed]
  10. Mutelet, B.; Boudin, S.; Pérez, O.; Rueff, J.M.; Labbé, C.; Jaffrès, P.A. La1−xLnxH(O3PCH3)2 (Ln = Tb, Eu; 0 < x ≤ 1): An organic-inorganic hybrid with lanthanide chains and tunable luminescence properties. Dalton Trans. 2015, 44, 1186–1192. [Google Scholar] [CrossRef] [PubMed]
  11. Colodrero, R.M.P.; Olivera-Pastor, P.; Losilla, E.R.; Aranda, M.A.G.; Leon-Reina, L.; Papadaki, M.; McKinlay, A.C.; Morris, R.E.; Demadis, K.D.; Cabeza, A. Multifunctional lanthanum tetraphosphonates: Flexible, ultramicroporous and proton-conducting hybrid frameworks. Dalton Trans. 2012, 41, 4045–4051. [Google Scholar] [CrossRef] [PubMed]
  12. Horike, S.; Umeyama, D.; Kitagawa, S. Ion Conductivity and Transport by Porous Coordination Polymers and Metal–Organic Frameworks. Acc. Chem. Res. 2013, 46, 2376–2384. [Google Scholar] [CrossRef] [PubMed]
  13. Kim, S.; Dawson, K.W.; Gelfand, B.S.; Taylor, J.M.; Shimizu, G.K.H. Enhancing Proton Conduction in a Metal–Organic Framework by Isomorphous Ligand Replacement. J. Am. Chem. Soc. 2013, 135, 963–966. [Google Scholar] [CrossRef] [PubMed]
  14. Donnadio, A.; Nocchetti, M.; Costantino, F.; Taddei, M.; Casciola, M.; da Silva Lisboa, F.; Vivani, R. A layered mixed zirconium phosphate/phosphonate with exposed carboxylic and phosphonic groups: X-ray powder structure and proton conductivity properties. Inorg. Chem. 2014, 53, 13220–13226. [Google Scholar] [CrossRef] [PubMed]
  15. Berchel, M.; Gall, T.L.; Denis, C.; Hir, S.L.; Quentel, F.; Elléouet, C.; Montier, T.; Rueff, J.-M.; Salaün, J.-Y.; Haelters, J.-P.; et al. A silver-based metal–organic framework material as a ‘reservoir’ of bactericidal metal ions. New J. Chem. 2011, 35, 1000–1003. [Google Scholar] [CrossRef]
  16. Rueff, J.-M.; Perez, O.; Caignaert, V.; Hix, G.; Berchel, M.; Quentel, F.; Jaffrès, P.-A. Silver-Based Hybrid Materials from meta- or para-Phosphonobenzoic Acid: Influence of the Topology on Silver Release in Water. Inorg. Chem. 2015, 54, 2152–2159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Bao, S.-S.; Zheng, L.-M. Magnetic materials based on 3d metal phosphonates. Coord. Chem. Rev. 2016, 319, 63–85. [Google Scholar] [CrossRef]
  18. Zheng, L.-M.; Duan, Y. Structural and Magnetic Studies of Cobalt Phosphonates. In Metal Phosphonate Chemistry; Clearfield, A., Demadis, K., Eds.; Royal Society of Chemistry: Cambridge, UK, 2011; Chapter 8; pp. 235–278. ISBN 978-1-84973-356-4. [Google Scholar]
  19. Yin, P.; Gao, S.; Zheng, L.-M. Xin Magnetic Properties of Metal Diphosphonate Compounds with One-Dimensional Chain Structures. Chem. Mater. 2003, 15, 3233–3236. [Google Scholar] [CrossRef]
  20. Zheng, L.-M.; Gao, S.; Yin, P. Xin One-Dimensional Cobalt Diphosphonates Exhibiting Weak Ferromagnetism and Field-Induced Magnetic Transitions. Inorg. Chem. 2004, 43, 2151–2156. [Google Scholar] [CrossRef] [PubMed]
  21. Yin, P.; Gao, S.; Wang, Z.-M.; Yan, C.-H.; Zheng, L.-M. Xin Field-Induced Magnetic Transitions in Metal Phosphonates with Ladderlike Chain Structures: (NH3C6H4NH3)M2(hedpH)2·H2O [M = Fe, Co, Mn, Zn; hedp = C(CH3)(OH)(PO3)2]. Inorg. Chem. 2005, 44, 2761–2765. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, Z.-C.; Bao, S.-S.; Zheng, L.-M. Ladder-like metal diphosphonates exhibiting field-induced magnetic transitions. Inorg. Chem. Commun. 2007, 10, 1063–1066. [Google Scholar] [CrossRef]
  23. Zhang, Z.-C.; Gao, S.; Zheng, L.-M. Cobalt diphosphonate with a new double chain structure exhibiting field-induced magnetic transition. Dalton Trans. 2007, 4681–4684. [Google Scholar] [CrossRef] [PubMed]
  24. Yang, T.-H.; Knowles, E.S.; Pajerowski, D.M.; Xia, J.-S.; Yin, L.; Gao, S.; Meisel, M.W.; Zheng, L.-M. Metal Monophosphonates M{(2-C5H4NO)CH2PO3}(H2O)2 (M = Co, Ni, Mn, Cd): Synthesis, Structure, and Magnetism. Inorg. Chem. 2010, 49, 8474–8480. [Google Scholar] [CrossRef] [PubMed]
  25. Yang, Y.; Chen, M.; Tang, X.; Yuan, R.; Ma, Y. A one-dimensional cobalt phosphonate showing field-induced magnetic transition. Inorg. Chem. Commun. 2018, 89, 60–63. [Google Scholar] [CrossRef]
  26. Wang, P.; Duan, Y.; Zheng, L. One-dimensional metal phosphonates based on 6-phosphononicotinic acid: A structural and magnetic study. Sci. China Chem. 2010, 53, 2112–2117. [Google Scholar] [CrossRef]
  27. Cao, D.-K.; Xiao, J.; Li, Y.-Z.; Clemente-Juan, J.M.; Coronado, E.; Zheng, L.-M. Metal Phosphonates Based on {[(Benzimidazol-2-ylmethyl)imino]bis(methylene)}bis(phosphonic Acid): Syntheses, Structures and Magnetic Properties of the Chain Compounds [M{(C7H5N2)CH2N(CH2PO3H)2}](M = Mn, Fe, Co, Cu, Cd). Eur. J. Inorg. Chem. 2006, 2006, 1830–1837. [Google Scholar] [CrossRef]
  28. Han, G.-F.; Luo, H.-Z.; Ye, Q.; Xiong, R.-G. Two Novel Cobalt(II) Coordination Polymers Based on the Herbicide Glyphosate as a Building Block and their Magnetic Properties. Z. Anorg. Allg. Chem. 2008, 634, 1991–1995. [Google Scholar] [CrossRef]
  29. Sun, Z.-M.; Prosvirin, A.V.; Zhao, H.-H.; Mao, J.-G.; Dunbar, K.R. New type of single chain magnet based on spin canting in an antiferromagnetically coupled Co(II) chain. J. Appl. Phys. 2005, 97, 10B305. [Google Scholar] [CrossRef]
  30. Palii, A.V.; Ostrovsky, S.M.; Klokishner, S.I.; Reu, O.S.; Sun, Z.-M.; Prosvirin, A.V.; Zhao, H.-H.; Mao, J.-G.; Dunbar, K.R. Origin of the Single Chain Magnet Behavior of the Co(H2L)(H2O) Compound with a 1D Structure. J. Phys. Chem. A 2006, 110, 14003–14012. [Google Scholar] [CrossRef] [PubMed]
  31. Caneschi, A.; Gatteschi, D.; Lalioti, N.; Sangregorio, C.; Sessoli, R.; Venturi, G.; Vindigni, A.; Rettori, A.; Pini, M.G.; Novak, M.A. Cobalt(II)-Nitronyl Nitroxide Chains as Molecular Magnetic Nanowires. Angew. Chem. Int. Ed. 2001, 40, 1760–1763. [Google Scholar] [CrossRef]
  32. Caneschi, A.; Gatteschi, D.; Lalioti, N.; Sessoli, R.; Sorace, L.; Tangoulis, V.; Vindigni, A. Ising-Type Magnetic Anisotropy in a Cobalt(II) Nitronyl Nitroxide Compound: A Key to Understanding the Formation of Molecular Magnetic Nanowires. Chem. Eur. J. 2002, 8, 286–292. [Google Scholar] [CrossRef]
  33. Clérac, R.; Miyasaka, H.; Yamashita, M.; Coulon, C. Evidence for Single-Chain Magnet Behavior in a MnIII−NiII Chain Designed with High Spin Magnetic Units: A Route to High Temperature Metastable Magnets. J. Am. Chem. Soc. 2002, 124, 12837–12844. [Google Scholar] [CrossRef] [PubMed]
  34. Lescouëzec, R.; Vaissermann, J.; Ruiz-Pérez, C.; Lloret, F.; Carrasco, R.; Julve, M.; Verdaguer, M.; Dromzee, Y.; Gatteschi, D.; Wernsdorfer, W. Cyanide-Bridged Iron(III)–Cobalt(II) Double Zigzag Ferromagnetic Chains: Two New Molecular Magnetic Nanowires. Angew. Chem. 2003, 115, 1521–1524. [Google Scholar] [CrossRef]
  35. Miyasaka, H.; Clérac, R.; Mizushima, K.; Sugiura, K.; Yamashita, M.; Wernsdorfer, W.; Coulon, C. [3Mn2(saltmen)2Ni(pao)2(L)2](A)2 with L = Pyridine, 4-Picoline, 4-tert-Butylpyridine, N-Methylimidazole and A = ClO4−, BF4−, PF6−, ReO4−: A Family of Single-Chain Magnets. Inorg. Chem. 2003, 42, 8203–8213. [Google Scholar] [CrossRef] [PubMed]
  36. Coulon, C.; Clérac, R.; Lecren, L.; Wernsdorfer, W.; Miyasaka, H. Glauber dynamics in a single-chain magnet: From theory to real systems. Phys. Rev. B 2004, 69, 132408. [Google Scholar] [CrossRef]
  37. Gatteschi, D.; Sessoli, R. Quantum Tunneling of Magnetization and Related Phenomena in Molecular Materials. Angew. Chem. Int. Ed. 2003, 42, 268–297. [Google Scholar] [CrossRef] [PubMed]
  38. Gatteschi, D.; Sessoli, R.; Villain, J. Mesoscopic Physics and Nanotechnology. In Molecular Nanomagnets; Oxford University Press: Oxford, UK, 2006; ISBN 978-0-19-856753-0. [Google Scholar]
  39. Sheldrick, G.M. Twinabs; University of Göttingen: Göttingen, Germany, 1996. [Google Scholar]
  40. Burla, M.C.; Caliandro, R.; Camalli, M.; Carrozzini, B.; Cascarano, G.L.; Giacovazzo, C.; Mallamo, M.; Mazzone, A.; Polidori, G.; Spagna, R. SIR2011: A new package for crystal structure determination and refinement. J. Appl. Crystallogr. 2012, 45, 357–361. [Google Scholar] [CrossRef]
  41. Václav, P.; Michal, D.; Lukáš, P. Crystallographic Computing System JANA2006: General features. zkri 2014, 229, 345. [Google Scholar] [CrossRef]
  42. Shannon, R.D.; Prewitt, C.T. Effective ionic radii in oxides and fluorides. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1969, 25, 925–946. [Google Scholar] [CrossRef] [Green Version]
  43. Spackman, M.A.; Jayatilaka, D. Hirshfeld surface analysis. CrystEngComm 2009, 11, 19–32. [Google Scholar] [CrossRef]
  44. McKinnon, J.J.; Spackman, M.A.; Mitchell, A.S. Novel tools for visualizing and exploring intermolecular interactions in molecular crystals. Acta Crystallogr. Sect. B 2004, 60, 627–668. [Google Scholar] [CrossRef] [PubMed]
  45. Spackman, M.A.; McKinnon, J.J. Fingerprinting intermolecular interactions in molecular crystals. CrystEngComm 2002, 4, 378–392. [Google Scholar] [CrossRef]
  46. Bloyet, C.; Rueff, J.-M.; Caignaert, V.; Lohier, J.-F.; Cardin, J.; Jaffrès, P.-A.; Raveau, B. Fluorenyl Zinc Phosphonate Zn(H2O)PO3–C13H9·H2O: Hybrid Columnar Structure with Strong C–H···π Interactions. Z. Anorg. Allg. Chem. 2017, 643, 250–255. [Google Scholar] [CrossRef]
  47. Demadis, K.D.; Panera, A.; Anagnostou, Z.; Varouhas, D.; Kirillov, A.M.; Císařová, I. Disruption of “Coordination Polymer” Architecture in Cu2+ Bis-Phosphonates and Carboxyphosphonates by Use of 2,2′-Bipyridine as Auxiliary Ligand: Structural Variability and Topological Analysis. Cryst. Growth Des. 2013, 13, 4480–4489. [Google Scholar] [CrossRef]
  48. Bloyet, C.; Rueff, J.-M.; Cardin, J.; Caignaert, V.; Doualan, J.-L.; Lohier, J.-F.; Jaffrès, P.-A.; Raveau, B. Excimer and Red Luminescence Due to Aggregation-Induced Emission in Naphthalene Based Zinc Phosphonate. Eur. J. Inorg. Chem. 2018, 3095–3103. [Google Scholar] [CrossRef]
  49. Carlin, R.L. Magnetochemistry; Springer: Berlin/Heidelberg, Germany, 1986; ISBN 978-3-642-70735-3. [Google Scholar]
  50. Goodenough, J.B. An interpretation of the magnetic properties of the perovskite-type mixed crystals La1−xSrxCoO3−λ. J. Phys. Chem. Solids 1958, 6, 287–297. [Google Scholar] [CrossRef]
  51. Kanamori, J. Superexchange interaction and symmetry properties of electron orbitals. J. Phys. Chem. Solids 1959, 10, 87–98. [Google Scholar] [CrossRef]
  52. Oka, Y.; Inoue, K.; Kumagai, H.; Kurmoo, M. Long-Range Magnetic Ordering at 5.5 K for Cobalt(II)–Hydroxide Diamond Chains Isolated by 17 Å with α-Phenylcinnamate. Inorg. Chem. 2013, 52, 2142–2149. [Google Scholar] [CrossRef] [PubMed]
  53. Wolff, S.K.; Grimwood, D.J.; McKinnon, J.J.; Turner, M.J.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer; University of Western Australia: Crawley, WA, Australia, 2012. [Google Scholar]
Scheme 1. Chemical structure of organic precursors of hybrid materials: (a) 9-fluorenylphosphonic acid which is used in this study and (b) (4-methylphenyl)methylamine(bis-methylenephosphonic acid which is used in the synthesis of cobalt-based single-chain magnets (SCMs) [30].
Scheme 1. Chemical structure of organic precursors of hybrid materials: (a) 9-fluorenylphosphonic acid which is used in this study and (b) (4-methylphenyl)methylamine(bis-methylenephosphonic acid which is used in the synthesis of cobalt-based single-chain magnets (SCMs) [30].
Inorganics 06 00093 sch001
Figure 1. Oblong crystallites of Co(H2O)2PO3C–C12H9·H2O observed by scanning electron microscopy.
Figure 1. Oblong crystallites of Co(H2O)2PO3C–C12H9·H2O observed by scanning electron microscopy.
Inorganics 06 00093 g001
Figure 2. (a) Projection of the structure of Co(H2O)2PO3C–C12H9·H2O along the b axis. In blue dashed lines are represented the O–H···O hydrogen bonds between the H2O molecules, oxygen atoms of the cobalt octahedra, and PO3C tetrahedra, with H···O distances between 2.93 Å and 3.20 Å. The black square represents the unit cell. (b) Representation of one column along the c axis. The water molecules were removed in order to see clearly the π···π stacking between the fluorene moieties. The green dashed lines correspond to C–H···C interactions with H···C distances between 3.10 Å and 3.13 Å.
Figure 2. (a) Projection of the structure of Co(H2O)2PO3C–C12H9·H2O along the b axis. In blue dashed lines are represented the O–H···O hydrogen bonds between the H2O molecules, oxygen atoms of the cobalt octahedra, and PO3C tetrahedra, with H···O distances between 2.93 Å and 3.20 Å. The black square represents the unit cell. (b) Representation of one column along the c axis. The water molecules were removed in order to see clearly the π···π stacking between the fluorene moieties. The green dashed lines correspond to C–H···C interactions with H···C distances between 3.10 Å and 3.13 Å.
Inorganics 06 00093 g002
Figure 3. (a) Projection of the structure of Co(H2O)2PO3C–C12H9·H2O along the a axis without fluorene groups; (b) Projection of Co(H2L)(H2O) along the a axis without the organic part, adapted from the reference [30].
Figure 3. (a) Projection of the structure of Co(H2O)2PO3C–C12H9·H2O along the a axis without fluorene groups; (b) Projection of Co(H2L)(H2O) along the a axis without the organic part, adapted from the reference [30].
Inorganics 06 00093 g003
Figure 4. Projection of the structure of Co(H2O)2PO3C–C12H9·H2O along the c axis. The black square represents the unit cell. The orange dashed lines correspond to C–H···C interactions between two organic layers with H···C distances between 2.90 Å and 3.20 Å.
Figure 4. Projection of the structure of Co(H2O)2PO3C–C12H9·H2O along the c axis. The black square represents the unit cell. The orange dashed lines correspond to C–H···C interactions between two organic layers with H···C distances between 2.90 Å and 3.20 Å.
Inorganics 06 00093 g004
Figure 5. Variation of χ (black squares) and χT (open black circles) as function of the temperature for Co(H2O)2PO3C–C12H9·H2O under 0.1 T.
Figure 5. Variation of χ (black squares) and χT (open black circles) as function of the temperature for Co(H2O)2PO3C–C12H9·H2O under 0.1 T.
Inorganics 06 00093 g005
Figure 6. Variation of magnetization versus magnetic field at 2 K.
Figure 6. Variation of magnetization versus magnetic field at 2 K.
Inorganics 06 00093 g006
Figure 7. TGA curve of Co(H2O)2PO3C–C12H9·H2O recorded in air from room temperature to 1000 °C.
Figure 7. TGA curve of Co(H2O)2PO3C–C12H9·H2O recorded in air from room temperature to 1000 °C.
Inorganics 06 00093 g007
Table 1. Crystallographic data of Co(H2O)2PO3C–C12H9·H2O.
Table 1. Crystallographic data of Co(H2O)2PO3C–C12H9·H2O.
FormulaCo(H2O)2PO3C–C12H9·H2O
FW357.20
Space groupP21/c
a (Å) 15.216(3)
b (Å)4.827(8)
c (Å)19.590(3)
α, β, γ (°)90, 107.160(4), 90
Z4
V (Å3)1375.0(4)
dcalc (g/cm3)1.716
µ (mm−1)1.389
Radiation source λ (Å)Mo Kα 0.71069
Pattern range 2Ѳ (°)2.8–66.94
No. of reflexions21,476
No. of soft constraints0
Weighted R factor0.0877
R[F2 > 2σ(F2)]0.0462
Rint (internal R-value)0.0371
S (Goodness of the fit)1.720

Share and Cite

MDPI and ACS Style

Bloyet, C.; Rueff, J.-M.; Perez, O.; Pautrat, A.; Caignaert, V.; Raveau, B.; Rogez, G.; Jaffrès, P.-A. One-Dimensional Fluorene-Based Co(II) Phosphonate Co(H2O)2PO3C–C12H9·H2O: Structure and Magnetism. Inorganics 2018, 6, 93. https://doi.org/10.3390/inorganics6030093

AMA Style

Bloyet C, Rueff J-M, Perez O, Pautrat A, Caignaert V, Raveau B, Rogez G, Jaffrès P-A. One-Dimensional Fluorene-Based Co(II) Phosphonate Co(H2O)2PO3C–C12H9·H2O: Structure and Magnetism. Inorganics. 2018; 6(3):93. https://doi.org/10.3390/inorganics6030093

Chicago/Turabian Style

Bloyet, Clarisse, Jean-Michel Rueff, Olivier Perez, Alain Pautrat, Vincent Caignaert, Bernard Raveau, Guillaume Rogez, and Paul-Alain Jaffrès. 2018. "One-Dimensional Fluorene-Based Co(II) Phosphonate Co(H2O)2PO3C–C12H9·H2O: Structure and Magnetism" Inorganics 6, no. 3: 93. https://doi.org/10.3390/inorganics6030093

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