Next Article in Journal
Influence of Nanoscale Surface Arrangements on the Oxygen Transfer Ability of Ceria–Zirconia Mixed Oxide
Next Article in Special Issue
Reactivity of Coordinated 2-Pyridyl Oximes: Synthesis, Structure, Spectroscopic Characterization and Theoretical Studies of Dichlorodi{(2-Pyridyl)Furoxan}Zinc(II) Obtained from the Reaction between Zinc(II) Nitrate and Pyridine-2-Chloroxime
Previous Article in Journal
CO2 Derivatives of Molecular Tin Compounds. Part 1: Hemicarbonato and Carbonato Complexes
Previous Article in Special Issue
Copper(II) Halide Salts with 1-(4′-Pyridyl)-Pyridinediium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rare Nuclearities in Ni(II) Cluster Chemistry: An Unprecedented {Ni12} Nanosized Cage from the Use of N-Naphthalidene-2-Amino-5-Chlorobenzoic Acid

by
Panagiota S. Perlepe
1,
Konstantinos N. Pantelis
2,
Luís Cunha-Silva
3,
Vlasoula Bekiari
4,
Albert Escuer
5 and
Theocharis C. Stamatatos
1,2,*
1
Department of Chemistry, Brock University, 1812 Sir Isaac Brock Way, St. Catharines, ON L2S 3A1, Canada
2
Chemistry Department, University of Patras, 265 04 Patras, Greece
3
LAQV/REQUIMTE & Department of Chemistry and Biochemistry, Faculty of Sciences, University of Porto, 4169-007 Porto, Portugal
4
School of Agricultural Sciences, University of Patras, 30200 Messolonghi, Greece
5
Departament de Quimica Inorgànica i Orgànica, Secció Inorgànica and Institut de Nanociència i Nanotecnologia (IN2UB), Universitat de Barcelona, Martí Franqués 1-11, 08028 Barcelona, Spain
*
Author to whom correspondence should be addressed.
Submission received: 10 March 2020 / Revised: 24 April 2020 / Accepted: 6 May 2020 / Published: 9 May 2020

Abstract

:
The self-assembly reaction between NiI2, benzoic acid (PhCO2H) and the Schiff base chelate, N-naphthalidene-2-amino-5-chlorobenzoic acid (nacbH2), in the presence of the organic base triethylamine (NEt3), has resulted in the isolation and the structural, spectroscopic, and physicochemical characterization of the dodecanuclear [Ni12I2(OH)6(O2CPh)5(nacb)5(H2O)4(MeCN)4]I (1) cluster compound in ~30% yield. Complex 1 has a cage-like conformation, comprising twelve distorted, octahedral NiII ions that are bridged by five μ3-OH, one μ-OH, an I in 55% occupancy, five PhCO2 groups (under the η11:μ, η123 and η224 modes), and the naphthoxido and carboxylato O-atoms of five doubly deprotonated nacb2− groups. The overall {Ni12} cluster exhibits a nanosized structure with a diameter of ~2.5 nm and its metallic core can be conveniently described as a series of nine edge- or vertex-sharing {Ni3} triangular subunits. Complex 1 is the highest nuclearity coordination compound bearing the nacbH2 chelate, and a rare example of polynuclear NiII complex containing coordinating I ions. Direct current (DC) magnetic susceptibility studies revealed the presence of predominant antiferromagnetic exchange interactions between the NiII ions, while photophysical studies of 1 in the solid-state showed a cyan-to-green centered emission at 520 nm, upon maximum excitation at 380 nm. The reported results demonstrate the rich coordination chemistry of the deprotonated nacb2− chelate in the presence of NiII metal ions, and the ability of this ligand to adopt a variety of different bridging modes, thus fostering the formation of high-nuclearity molecules with rare, nanosized dimensions and interesting physical (i.e., magnetic and optical) properties.

1. Introduction

Polynuclear 3d-metal complexes (or 3d-metal clusters) remain one of the most attractive research fields in the cross-disciplinary areas of chemistry, physics, and materials science [1]. This is mainly due to the ability of these nanosized molecular species to exhibit very interesting magnetic, optical, biological, and catalytic properties, to name just a few [2]. From a structural perspective, the motifs of these cluster compounds often resemble the aesthetically beautiful structures of highly-symmetric inorganic solids, such as cubic and hexagonal structures, perovskites, brucites, and supertetrahedra, due to the presence of multiple bridging oxido and hydroxido groups [3]. In the molecular magnetism arena, ferromagnetically-coupled 3d-systems with a large ground state spin value, S, and an appreciable magnetic anisotropy of the Ising (or easy-axis) type can behave as single-molecule magnets (SMMs) [4]. SMMs exhibit slow magnetization relaxation over and/or through an anisotropy barrier and they represent a molecular or “bottom-up” approach to nanoscale magnetism with potential applications in the fields of information storage, molecular electronics and spintronics [5].
The first and most well-studied family of SMMs is the mixed-valence [MnIII8MnIV4O12(O2CR)16(X)4], where RCO2 and X are various carboxylate bridging groups and terminal solvate molecules, respectively [6]. The SMM behavior of these compounds originates from the combined S = 10 spin ground state and the enhanced magnetic anisotropy resulting from the parallel alignment of the MnIII Jahn-Teller axes. In polynuclear NiII cluster chemistry, the number of SMMs is substantially smaller [7] and only a few of them show slow relaxation of magnetization in the absence of an external DC field, as well as magnetization hysteresis, the diagnostic property of a magnet. This is predominately due to the small zero-field splitting parameter, D, that a polynuclear NiII complex often exhibits when the NiII atoms adopt the favorable octahedral coordination geometry. Exceptional examples of NiII cluster-based SMMs are the ferromagnetic [Ni12(chp)12(O2CMe)12(THF)6(H2O)6] with a ring-like topology [8], and the family of [Ni4(hmp)4(ROH)4Cl4] complexes with a distorted cubane [Ni4(OR)4]4+ core [9], where chpH and hmpH are the organic chelates chloro-2-hydroxypyridine and 2-hydroxymethylpyridine, respectively, and ROH are various terminally-bound alcohol solvates.
It becomes apparent that the choice of the organic chelating/bridging ligand is of fundamental importance in the self-assembly synthesis of high-nuclearity NiII complexes with high-spin values and interesting magnetic dynamics. To this end, we have recently started a research program aiming at the exploration of Schiff base chelates, which are based on the tridentate N-salicylidene-o-aminophenol (saphH2, Scheme 1) scaffold, in 3d-metal cluster chemistry as a means of obtaining nanosized molecular materials, primarily those with interesting magnetic properties [10]. To increase the coordination and bridging potential of the organic chelate, we initially turned our attention to the tetradentate ligand N-salicylidene-2-amino-5-chlorobenzoic acid (sacbH2, Scheme 1); this has led to the structurally impressive {Ni18} and {Ni26} clusters [11], and a {Dy2} SMM with a large energy barrier for magnetization reversal [12]. A reasonable leap forward would be the replacement of the phenyl ring of the N-salicylidene moiety with a naphthalene one. The resulting ligand N-naphthalidene-2-amino-5-chlorobenzoic acid (nacbH2, Scheme 1) shows the following salient features: (a) it is still a tetradentate like sacbH2, but is undoubtedly more rigid and sterically demanding than sacbH2, and (b) it includes the naphthalene substituent, a well-known fluorescent group [13], which could open new prospects in the emission properties of NiII coordination compounds with O- and N-donor atoms. Both features presage the synthesis of new cluster compounds with potentially interesting magnetic and emission properties. Indeed, the initial employment of nacbH2 in NiII chemistry has afforded a series of {Ni5} and {Ni6} clusters with diverse magnetic and optical properties, but with limited bridging affinity for nacb2− [14]. In this work, we have unveiled the bridging capacity of nacb2− in conjunction with ancillary bridging benzoate groups. We herein report an unprecedented {Ni12} cluster compound with the highest nuclearity in metal cluster chemistry of nacbH2, and one of the rarest nuclearities in NiII cluster chemistry.

2. Results and Discussion

2.1. Synthetic Comments

The general reaction system NiX2/nacbH2, where X are various anions with either a strong (i.e., NO3, β-diketonates and pseudohalides) or weak coordinating (Cl, Br and ClO4) ability, has been studied by us [15], and—in almost all the cases—small in nuclearity clusters (i.e., {Ni5} and {Ni6}) were isolated and structurally characterized [14]. These results have demonstrated the unpredictability of the nacbH2 chelate toward the coordination with 3d-metal ions in solution and subsequently the stabilization of different cluster compounds in the solid-state. We have thus decided to introduce to the NiII/nacbH2 system anions with very limited coordinating capacity, such as iodides (I), in conjunction with organic anionic groups with a superior bridging ability, such as benzoates (PhCO2), in an attempt to harness a flexible ”synthetic blend” which would potentially lead to high-nuclearity NiII clusters. Indeed, the reaction between NiI2, nacbH2, PhCO2H, and NEt3 in a 2:1:1:3 molar ratio, in solvent acetonitrile (MeCN), afforded dark-green crystals of the dodecanuclear cluster compound [Ni12I2(OH)6(O2CPh)5(nacb)5(H2O)4(MeCN)4]I (1) in 30% yield. The general formation of 1 is summarized by the following stoichiometric Equation (1).
12 NiI2 + 5 nacbH2 + 5 PhCO2H + 21 NEt3 + 10 H2O + 4 MeCN → [Ni12I2(OH)6(O2CPh)5(nacb)5(H2O)4(MeCN)4]I + 21 NHEt3I
Under the context of chemical reactivity, several synthetic parameters were explored to either increase the yield of the isolated product 1 or alter the nuclearity of the {Ni12} compound and subsequently isolate a new product. In particular, the employment of NiCl2 or NiBr2 in place of NiI2 afforded the already reported (NHEt3)2[Ni6(OH)2(nacb)6(H2O)4] [14], whereas the replacement of PhCO2H by other carboxylic acids, such as MeCO2H or EtCO2H, led to green-colored microcrystalline products, of which we were unable to determine the crystal structures due to the very small size of the obtained crystallites. The presence of NEt3 as an external base was also essential for the clean preparation of 1, providing a proton acceptor to facilitate the complete deprotonation of nacbH2 and PhCO2H, and fostering the metal-assisted deprotonation of H2O molecules in solution to the coordinating OH groups (vide infra). Various similar reactions in other organic solvents (i.e., alcohols (ROH), CH2Cl2 and mixtures of MeCN/ROH) and/or external bases (i.e., trimethylamine (NMe3), tripropylamine (NPr3), diethylamine (Et2NH), dimethylamine (Me2NH) and tetramethylammonium hydroxide (Me4NOH) yielded amorphous solids that we were unable to recrystallize and further characterize. Finally, it is worth mentioning that the presence of iodide ions, either as bound groups or counterions, or both, in NiII coordination chemistry is limited to a handful of previously reported dinuclear or trinuclear compounds [16,17,18,19].

2.2. Description of Structure

A partially labeled representation of the cation of complex 1 is shown in Figure 1. The positively charged cluster cation [Ni12I2(OH)6(O2CPh)5(nacb)5(H2O)4(MeCN)4]+ is counterbalanced by an I anion in the crystal lattice of 1. The I counterion (I7) is closely held with the {Ni12} cluster through H-bonding interactions with three of the bridging OH groups; these separations are: O1∙∙∙I7 = 3.226(3) Å, O2∙∙∙I7 = 3.425(1) Å and O3∙∙∙I7 = 3.222(3) Å. Selected interatomic distances and angles of 1 are listed in Table 1. Bond valence sum (BVS) calculations for the inorganic bridging O-atoms with 100% occupancies gave values of: 1.23 (for O1), 1.16 (for O2 and O8), 1.10 (for O4), and 1.18 (for O9), in excellent agreement with their assignment as OH groups. Oxygen BVS values in the ~1.7–2.0, ~1.0–1.2, and ~0.2–0.4 ranges are indicative of non-, single- and double-protonation, respectively [20].
Complex 1 is a closed cage-like cluster consisting of 12 NiII ions that are bridged by five μ3-OH, one μ-OH, an I in 55% occupancy, and the naphthoxido and carboxylato O-atoms of five double-deprotonated nacb2− groups. The latter are arranged into three classes (Figure 2), with all of them acting as tridentate chelates to a NiII ion, and additionally bridging two (η11213 mode), three (η21214 mode) and four (η21225 mode) metal ions. The variety of binding modes of nacb2− in complex 1 clearly emphasizes the coordination affinity of this Schiff base ligand with NiII ions, and its ability to stabilize high-nuclearity 3d-metal clusters with unprecedented structural motifs and nanosized dimensions. To this end, the space-filling plot (Figure 3) shows that 1 has a nearly ”bowl”-shaped conformation with the longest intramolecular C···C distance being ~25 Å, excluding the H atoms. The shortest Ni···Ni distance between neighboring {Ni12} clusters in the crystal is 12.246(2) Å, thus confirming the good separation of the cluster compounds due to the bulky naphthalene substituents of the nacb2− ligands.
Additional bridging about the twelve NiII ions is provided by five PhCO2 groups, which are arranged into three classes; three of them are bridging under the η11:μ mode, one is acting as an η123 ligand, and the last one adopts the rare η224 mode. Thus, the resulting core is [Ni123-OH)53-Ι/H2O)(μ-OR)15]3+ (Figure 4), and peripheral ligation about this core is further provided by four terminally bound MeCN molecules (on Ni5, Ni9, Ni11, and Ni12) and a total of five I/H2O group combinations (on Ni1, Ni4, Ni8, Ni9, and Ni12). All NiII atoms are six-coordinate with distorted octahedral geometries. The metallic core of 1 can be conveniently described as a series of nine edge- or vertex-sharing {Ni3} triangular subunits (Figure 5), which are held together by μ3-OH and μ-OR groups. An alternative description of the {Ni12} metal arrangement is that of a central {Ni7} subunit possessing a distorted disk-like topology [Ni(1,2,3,5,6,7,10)], which is attached to a {Ni3} triangular [Ni(3,4,12)] and a {Ni4} rhombus-shaped [Ni(7,8,9,11)] subunits by sharing the common Ni3 and Ni7 vertices, respectively. Finally, the nuclearity of complex 1 is the largest reported to date of a metal cluster bearing nacb2− chelate, and it joins a relatively rare family of {Ni12} Werner-type compounds with a cage-like conformation [21,22,23,24].

2.3. Solid-State Magnetic Susceptibility Studies

Variable-temperature (2.0–300 K range), direct-current (DC) magnetic susceptibility measurements were performed on a freshly-prepared microcrystalline solid of 1 under a weak DC field of 0.03 T to avoid saturation effects. The data are shown as χMT versus T plot in Figure 6. The value of the χΜT product at 300 K is 10.40 cm3 Kmol−1, slightly lower than the value of 12 cm3 Kmol−1 (calculated with g = 2.0) expected for twelve non-interacting, high-spin NiII (S = 1) atoms. Upon cooling, the χΜT product continuously decreases down to a value of 3.06 cm3 Kmol−1 at 2 K. A slightly different curvature of the plot is observed below ~5 K, and this likely due to the onset of zero-field splitting, intermolecular antiferromagnetic interactions between the {Ni12} clusters, and/or Zeeman effects [11,15]. The overall shape of the χMT versus T plot is suggestive of the presence of predominant antiferromagnetic exchange interactions between the metal centers, as frequently observed in many high-nuclearity and low-symmetry NiII cage-like clusters where many different magnetic exchange pathways are in effect [11]. To this end, a fit of the experimental data to a theoretical model (H = −2JijŜi·Ŝj convention) was not feasible. Undoubtedly, 1 possesses a small ground-state spin value, with the χΜT value at 2 K being consistent with an S ~ 2 ground state (for g = 2). The antiferromagnetic response of the {Ni12} compound can be tentatively assigned to the majority of obtuse Ni–O–Ni bond angles (close to or larger than 100°) and the presence of {Ni3} triangular subunits, which are prone to spin frustration effects [25]. Magnetization (M) versus field (H) measurements (Figure 6, inset) at 2 K show a continuous increase of M as the field increases, reaching a non-saturated value of 6.0 NμB at 5 T; this is likely due to the presence of low-lying excited states, as reported previously for other high-nuclearity NiII complexes [26]. As a result, attempts to fit the reduced magnetization data assuming that only the ground state is populated were very poor.

2.4. Solid-State Emission Studies

The photophysical properties of complex 1 were carried out in the solid-state and at room temperature due to its structural instability in solution. This was confirmed by performing electrospray ionization mass spectrometry (ESI-MS) studies in various solvent media (Figure S1). The optical response of the free-ligand nacbH2 has been reported by us in a previous work [15]. Briefly, it was shown that nacbH2 is a promising “antenna” group for the promotion of energy transfer effects. Upon maximum excitation at ~350 nm, nacbH2 exhibits a strong emission in the visible range with two clear maxima at ~390 and 410 nm, and a weak shoulder at ~480 nm. Complex 1 shows an interesting photophysical response, given its large nuclearity, the presence of many different binding groups, and the possible quenching effects from the coordinating O- and N-atoms. The dodecanuclear compound 1 exhibits a cyan-to-green centered emission at 520 nm, upon maximum excitation at 380 nm (Figure 7). The red-shifted emission of 1 with respect to the free nacbH2 can be tentatively assigned to the coordination of the deprotonated nacb2− ligands with the metal ions and/or the presence of additional binding groups with emission efficiency, such as benzoates, which could affect the charge transfer process and resulting emission [14,15].
In general, the loss of energy due to vibrations is reduced as a result of the coordination of a ligand to a metal center; this binding enhances the organic ligand’s rigidity [27]. In addition, the usually observed optical quenching effects from the paramagnetic metal ions can be prevented using organic fluorescent groups, such as the naphthalene, anthracene, and phenanthrene substituents [28]. Red-shifted emissions are commonly observed in most fluorescent compounds in the solid-state, likely due to the π–π stacking interactions of the aromatic rings [29]. Due to the structural complexity of 1, as a result of the many different ligands present and the number of metal ions, among other electronic and steric perturbations, an in-depth analysis of the photophysical properties of 1 would be unrealistic.

3. Materials and Methods

3.1. Materials, Physical and Spectroscopic Measurements

All manipulations were performed under aerobic conditions using materials (reagent grade) and solvents as received unless otherwise noted. The Schiff base ligand nacbH2 was prepared, purified, and characterized as described elsewhere [14,15]. Elemental analyses (C, H, and N) were performed on a Perkin-Elmer 2400 Series II Analyzer (Foster City, CA, USA). Infrared spectra were recorded in the solid state on a Bruker’s FT-IR spectrometer (ALPHA’s Platinum ATR single reflection) (Billerica, MA, USA) in the 4000–400 cm−1 range. Excitation and emission spectra were recorded in the solid state at room temperature conditions using a Cary Eclipse spectrofluorometer. The repeatability and reproducibility of the emission was verified by recording the emission spectra of the material three times in two different days using the same scan rate and the same excitation and emission monochromator slits. Variable-temperature magnetic susceptibility studies were performed on a MPMS5 Quantum Design magnetometer equipped with a 5 T magnet and operating in the 2–300 K range. The microcrystalline sample was embedded in solid eicosane to prevent torquing. Diamagnetic corrections were applied to the observed paramagnetic susceptibility using Pascal’s constants [30].

3.2. Synthesis of [Ni12I2(OH)6(O2CPh)5(nacb)5(H2O)4(MeCN)4]I (1)

To a stirred, orange suspension of nacbH2 (0.07 g, 0.20 mmol) in MeCN (20 mL) was added NEt3 (84 μL, 0.60 mmol). Solids NiI2 (0.13 g, 0.40 mmol) and PhCO2H (0.03 g, 0.20 mmol) were added to the resulting yellow solution, and a noticeable color change to a dark-green solution was observed over the period of 1 h, under a continuous magnetic stirring. The final solution was filtered, and the filtrate was carefully layered with Et2O (40 mL). After 20 days, X-ray quality dark-green plate-like crystals of 1 were formed, and these were collected by filtration, washed with cold MeCN (2 × 3 mL) and Et2O (2 × 3 mL), and dried in air. The yield was 30% (based on the ligand available). The air-dried solid was found to be slightly hygroscopic and it was satisfactorily analyzed as 1∙3H2O. Anal. calc. for C133H107N9O38Ni12Cl5I3 (found values in parentheses): C 43.16% (43.31%), H 2.91% (3.06%), N 3.41% (3.28%). Selected IR data (ATR): ν = 3300 (mb), 1598 (s), 1575 (s), 1535 (s), 1503 (w), 1472 (m), 1450 (m), 1427 (m), 1406 (s), 1382 (w), 1337 (s), 1298 (m), 1249 (m), 1217 (m), 1180 (m), 1156 (m), 1113 (m), 1088 (m), 985 (m), 961 (w), 885 (w), 852 (m), 828 (m), 743 (s), 718 (s), 670 (w), 630 (w), 555 (w), 451 (m).

3.3. Single-Crystal X-ray Crystallography

A suitable single-crystal of complex 1 was selected and mounted on the respective cryoloop using adequate inert oil [31]. Diffraction data were collected on a Bruker X8 Kappa APEX II Charge-Coupled Device (CCD) area-detector diffractometer (Billerica, MA, USA) controlled by the APEX2 software [32] package (Mo Kα graphite-monochromated radiation, λ = 0.71073 Å), and equipped with an Oxford Cryosystems Series 700 cryostream, monitored remotely with the software interface Cryopad [33]. Images were processed with the software SAINT+ [34], and the absorption effects were corrected by the multi-scan method implemented in SADABS [35]. The structure was solved using the algorithm implemented in SHELXT-2014 [36,37], and refined by successive full-matrix least-squares cycles on F2 using the latest SHELXL-v.2014 [36,38]. The non-hydrogen atoms of the crystal structure were successfully refined using anisotropic displacement parameters, and H-atoms bonded to carbon of the ligands were placed at their idealized positions using appropriate HFIX instructions in SHELXL. All these atoms were included in subsequent refinement cycles in riding-motion approximation with isotropic thermal displacement parameters (Uiso) fixed at 1.2 or 1.5 × Ueq of the relative atom. The refinement model revealed the presence of two coordinated iodide (I) ions and four coordinated H2O molecules partially disordered over six coordinative positions with distinct complementary occupancies (I1/O1W = 0.55/0.45; I2/O2W = 0.35/0.65; I3/O3W = 0.30/0.70; I4/O4W = 0.25/0.75; I5/O5W = 0.25/0.75, and I6/O6W = 0.30/0.70). In addition, the non-coordinated I ion is disordered over two positions with occupancies of 0.65 and 0.35. As a result of the severely disordered structure of 1, the H-atoms of the coordinated water molecules and hydroxido groups were not included in the refined model, but they were considered in the final molecular formula of the compound.
Substantial electron density was found on the data of complex 1, most probably due to additional disordered solvate molecules occupying the spaces originated by the close packing of the cluster compound. Various efforts to properly locate, model, and refine these residues were unsuccessful, and the examination for the total potential solvent area using the software package PLATON [39] clearly confirmed the existence of cavities with potential solvent accessible void volume. Thus, the original data set was treated with the program SQUEEZE [40], which calculates the contribution of the smeared electron density in the lattice voids and adds this to the calculated structure factors from the structural model when refining against the hkl file. The programs used for molecular graphics were MERCURY [41] and DIAMOND [42]. Unit cell parameters, structure solution and refinement details for 1 are summarized in Table 2. Further crystallographic details can be found in the corresponding CIF file provided in the ESI. Crystallographic data (excluding structure factors) for the structure reported in this work have been deposited to the Cambridge Crystallographic Data Centre (CCDC) as supplementary publication number: CCDC-1988904. Copies of the data can be obtained online using https://summary.ccdc.cam.ac.uk/structure-summary-form.

4. Conclusions and Perspectives

In conclusion, we have herein reported the synthesis, and structural and physicochemical characterization, of a new dodecanuclear NiII cluster compound with an unprecedented structural motif and nanoscale dimensions, resulted from the successful employment of the nacbH2/I/PhCO2 ”ligand blend” in NiII chemistry. The N-naphthalidene-2-amino-5-chlorobenzoic acid (nacbH2) ligand also contributed to the observation of unquenched optical emission from the {Ni12} complex 1 in the solid-state, a very unusual phenomenon in high-nuclearity 3d-metal cluster chemistry with N-/O-donor atoms. Complex 1 is by far the highest nuclearity NiII cluster compound with coordinating I ions, also supported by ancillary chelating and bridging organic groups, thus presaging a new synthetic approach to nanoscale molecular materials with interesting structural motifs and physical properties. We are currently investigating the NiII/RCO2/nacbH2 tertiary system as a means of obtaining higher-nuclearity, nanosized NiII clusters with interesting magneto-optical properties.

Supplementary Materials

The CIF and the checkCIF output files of complex 1 are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2304-6740/8/5/32/s1. Figure S1. Positive ion ES mass spectrum of 1 in a solvent mixture of MeCN/CH2Cl2.

Author Contributions

P.S.P. and K.N.P. conducted the syntheses, crystallization, purification, optimization, conventional characterization and interpretation of the structural and magnetic data of complex 1; L.C.-S. collected single-crystal X-ray diffraction data, solved the structure and performed the complete refinement; V.B. collected, plotted, rationalized and discussed the optical properties of complex 1; A.E. collected, plotted and discussed magnetic and part of the spectroscopic data of the compound; T.C.S. coordinated the research, contributed to the interpretation of the results and wrote the paper based on the reports of his collaborators; All the authors exchanged ideas and comments regarding the explanation of the results and discussed upon the manuscript at all stages. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by NSERC-DG and ERA to T.C.S. P.S.P thanks the Alexander S. Onassis Public Benefit Foundation for a graduate scholarship. L.C.-S. thanks the Fundação para a Ciência e a Tecnologia (FCT/Ministério da Ciência, Tecnologia e Ensino Superior, Portugal) for the financial support to the LAQV/REQUIMTE (UID/QUI/50006/2019) through national funds. A.E. acknowledges financial support from Ministerio de Economia y Competitividad, Project CTQ2018-094031-B-100.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Papatriantafyllopoulou, C.; Moushi, E.E.; Christou, G.; Tasiopoulos, A.J. Filling the gap between the quantum and classical worlds of nanoscale magnetism: Giant molecular aggregates based on paramagnetic 3d metal ions. Chem. Soc. Rev. 2016, 45, 1597–1628. [Google Scholar] [CrossRef] [PubMed]
  2. Maayan, G.; Gluz, N.; Christou, G. A bioinspired soluble manganese cluster as a water oxidation electrocatalyst with low overpotential. Nat. Catal. 2018, 1, 48–54. [Google Scholar] [CrossRef]
  3. Deng, Y.-K.; Su, H.-F.; Xu, J.-H.; Wang, W.-G.; Kurmoo, M.; Lin, S.-C.; Tan, Y.-Z.; Jia, J.; Sun, D.; Zheng, L.-S. Hierarchical assembly of a {MnII15MnIII4} brucite disc: Step-by-step formation and ferrimagnetism. J. Am. Chem. Soc. 2016, 138, 1328–1334. [Google Scholar] [CrossRef] [PubMed]
  4. Holynska, M. (Ed.) Single-Molecule Magnets: Molecular Architectures and Building Blocks for Spintronics; Wiley-VCH: Weinheim, Germany, 2018. [Google Scholar]
  5. Bogani, L.; Wernsdorfer, W. Molecular spintronics using single-molecule magnets. Nat. Mater. 2008, 7, 179–186. [Google Scholar] [CrossRef] [PubMed]
  6. Bagai, R.; Christou, G. The drosophila of single-molecule magnetism: [Mn12O12(O2CR)16(H2O)4]. Chem. Soc. Rev. 2009, 38, 1011–1026. [Google Scholar] [CrossRef] [PubMed]
  7. Milios, C.J.; Winpenny, R.E.P. Cluster-based single-molecule magnets. Struct. Bond. 2015, 164, 1–110. [Google Scholar]
  8. Blake, A.J.; Grant, C.M.; Parsons, S.; Rawson, J.M.; Winpenny, R.E.P. The synthesis, structure and magnetic properties of a cyclic dodecanuclear nickel complex. J. Chem. Soc. Chem. Commun. 1994, 2363–2364. [Google Scholar] [CrossRef]
  9. Yang, E.-C.; Wernsdorfer, W.; Hill, S.; Edwards, R.S.; Nakano, M.; Maccagnano, S.; Zakharov, L.N.; Rheingold, A.L.; Christou, G.; Hendrickson, D.N. Exchange bias in Ni4 single-molecule magnets. Polyhedron 2003, 22, 1727–1733. [Google Scholar] [CrossRef]
  10. Alexandropoulos, D.I.; Nguyen, T.N.; Cunha-Silva, L.; Zafiropoulos, T.F.; Escuer, A.; Christou, G.; Stamatatos, T.C. Slow magnetization relaxation in unprecedented MnIII4DyIII3 and MnIII4DyIII5 clusters from the use of N-salicylidene-o-aminophenol. Inorg. Chem. 2013, 52, 1179–1181. [Google Scholar] [CrossRef]
  11. Athanasopoulou, A.A.; Pilkington, M.; Raptopoulou, C.P.; Escuer, A.; Stamatatos, T.C. Structural aesthetics in molecular nanoscience: A unique Ni26 cluster with a ‘rabbit-face’ topology and a discrete Ni18 ‘molecular chain’. Chem. Commun. 2014, 50, 14942–14945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Mazarakioti, E.C.; Regier, J.; Cunha-Silva, L.; Wernsdorfer, W.; Pilkington, M.; Tang, J.; Stamatatos, T.C. Large energy barrier and magnetization hysteresis at 5 K for a symmetric {Dy2} complex with spherical tricapped trigonal prismatic DyIII ions. Inorg. Chem. 2017, 56, 3568–3578. [Google Scholar] [CrossRef] [PubMed]
  13. Alexandropoulos, D.I.; Mowson, A.M.; Pilkington, M.; Bekiari, V.; Christou, G.; Stamatatos, T.C. Emissive molecular nanomagnets: Introducing optical properties in triangular oximato {MnIII3} SMMs from the deliberate replacement of simple carboxylate ligands with their fluorescent analogues. Dalton Trans. 2014, 43, 1965–1969. [Google Scholar] [CrossRef] [PubMed]
  14. Perlepe, P.S.; Cunha-Silva, L.; Bekiari, V.; Gagnon, K.J.; Teat, S.J.; Escuer, A.; Stamatatos, T.C. Structural diversity in NiII cluster chemistry: Ni5, Ni6, and {NiNa2}n complexes bearing the Schiff-base ligand N-naphthalidene-2-amino-5-chlorobenzoic acid. Dalton Trans. 2016, 45, 10256–10270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Perlepe, P.S.; Cunha-Silva, L.; Gagnon, K.J.; Teat, S.J.; Lampropoulos, C.; Escuer, A.; Stamatatos, T.C. “Ligands-with-benefits”: Naphthalene-substituted Schiff bases yielding new NiII metal clusters with ferromagnetic and emissive properties and undergoing exciting transformations. Inorg. Chem. 2016, 55, 1270–1277. [Google Scholar] [CrossRef] [Green Version]
  16. Prema, D.; Oshin, K.; Desper, J.; Levy, C.J. Mono- and dinuclear nickel(II) complexes of resolved Schiff-base ligands with extended quinoline substituents. Dalton Trans. 2012, 41, 4998–5009. [Google Scholar] [CrossRef] [PubMed]
  17. Szacilowski, K.T.; Puhui, X.; Malkhasian, A.Y.S.; Heeg, M.J.; Udugala-Ganehenege, M.Y.; Wenger, L.E.; Endicott, J.F. Solid-state structures and magnetic properties of halide-bridged, face-to-face bis-nickel(II)-macrocyclic ligand complexes: Ligand-mediated interchanges of electronic configuration. Inorg. Chem. 2005, 44, 6019–6033. [Google Scholar] [CrossRef] [PubMed]
  18. Fossey, J.S.; Matsubara, R.; Kiyohara, H.; Kobayashi, S. Heterochiral triangulo nickel complex as evidence of a large positive nonlinear effect in catalysis. Inorg. Chem. 2008, 47, 781–783. [Google Scholar] [CrossRef] [PubMed]
  19. Morgenstern, D.A.; Ferrence, G.M.; Washington, J.; Henderson, J.I.; Rosenhein, L.; Heise, J.D.; Fanwick, P.E.; Kubiak, C.P. A class of halide-supported trinuclear nickel clusters [Ni33-L)(μ3-X)(μ2-dppm)3]n+ (L = I, Br, CO, CNR.; X = I, Br-; n = 0, 1; dppm = Ph2PCH2PPh2): Novel physical properties and the fermi resonance of symmetric μ31 bound isocyanide ligands. J. Am. Chem. Soc. 1996, 118, 2198–2207. [Google Scholar] [CrossRef]
  20. Liu, W.; Thorp, H.H. Bond valence sum analysis of metal-ligand bond lengths in metalloenzymes and model complexes. Refined distances and other enzymes. Inorg. Chem. 1993, 32, 4102–4105. [Google Scholar] [CrossRef]
  21. Efthymiou, C.G.; Cunha-Silva, L.; Perlepes, S.P.; Brechin, E.K.; Inglis, R.; Evangelisti, M.; Papatriantafyllopoulou, C. In search of molecules displaying ferromagnetic exchange: Multiple-decker Ni12 and Ni16 complexes from the use of pyridine-2-amidoxime. Dalton Trans. 2016, 45, 17409–17419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Breeze, B.A.; Shanmugam, M.; Tuna, F.; Winpenny, R.E.P. A series of nickel phosphonate-carboxylate cages. Chem. Commun. 2007, 5185–5187. [Google Scholar] [CrossRef] [PubMed]
  23. Feltham, H.L.C.; Clérac, R.; Brooker, S. Hexa-, hepta- and dodeca-nuclear nickel(II) complexes of three Schiff-base ligands derived from 1,4-diformyl-2,3-dihydroxybenzene. Dalton Trans. 2009, 2965–2973. [Google Scholar] [CrossRef] [PubMed]
  24. Pons-Balagué, A.; Piligkos, S.; Teat, S.J.; Sánchez Costa, J.; Shiddiq, M.; Hill, S.; Castro, G.R.; Ferrer-Escorihuela, P.; Sañudo, E.C. New nanostructured materials: Synthesis of dodecanuclear NiII complexes and surface deposition studies. Chem. Eur. J. 2013, 19, 9064–9071. [Google Scholar] [CrossRef] [PubMed]
  25. Athanasopoulou, A.A.; Raptopoulou, C.P.; Escuer, A.; Stamatatos, T.C. Rare nuclearities in Ni(II) cluster chemistry: A Ni11 cage from the first use of N-salicylidene-2-amino-5-chlorobenzoic acid in metal cluster chemistry. RSC Adv. 2014, 4, 12680–12684. [Google Scholar] [CrossRef]
  26. Papatriantafyllopoulou, C.; Diamantopoulou, E.; Terzis, A.; Tangoulis, V.; Lalioti, N.; Perlepes, S.P. High-nuclearity nickel(II) clusters: Ni13 complexes from the use of 1-hydroxybenzotriazole. Polyhedron 2009, 28, 1903–1911. [Google Scholar] [CrossRef]
  27. Bünzli, J.-C.G.; Piguet, C. Taking advantage of luminescent lanthanide ions. Chem. Soc. Rev. 2005, 34, 1048–1077. [Google Scholar] [CrossRef]
  28. Alaimo, A.; Takahashi, D.; Cunha-Silva, L.; Christou, G.; Stamatatos, T.C. Emissive {MnIII4Ca} clusters with square pyramidal topologies: Syntheses and structural, spectroscopic, and physicochemical characterization. Inorg. Chem. 2015, 54, 2137–2151. [Google Scholar] [CrossRef]
  29. Binnemans, K. Lanthanide-based luminescent hybrid materials. Chem. Rev. 2009, 109, 4283–4374. [Google Scholar] [CrossRef] [Green Version]
  30. Bain, G.A.; Berry, J.F. Diamagnetic corrections and Pascal’s constants. J. Chem. Educ. 2008, 85, 532–536. [Google Scholar] [CrossRef]
  31. Kottke, T.; Stalke, D. Crystal handling at low temperatures. J. App. Cryst. 1993, 26, 615–619. [Google Scholar] [CrossRef] [Green Version]
  32. APEX2. Data Collection Software Version 2012.4; Bruker AXS: Delft, The Netherlands, 2012. [Google Scholar]
  33. Cryopad. Remote Monitoring and Control, Version 1.451; Oxford Cryosystems: Oxford, UK, 2006. [Google Scholar]
  34. SAINT+. Data Integration Engine v. 7.23a ©; Bruker AXS: Madison, WI, USA, 1997–2005. [Google Scholar]
  35. Sheldrick, G.M. SADABS v.2.01. Bruker/Siemens Area Detector Absorption Correction Program; Bruker AXS: Madison, WI, USA, 1998. [Google Scholar]
  36. Sheldrick, G.M. A short history of SHELX. Acta Cryst. A 2008, 64, 112–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Sheldrick, G.M. SHELXT, Version 2014/3; Program for Crystal Structure Solution; University of Göttingen: Gottingen, Germany, 2014. [Google Scholar]
  38. Sheldrick, G.M. SHELXL, Version 2014; Program for Crystal Structure Refinement; University of Göttingen: Gottingen, Germany, 2014. [Google Scholar]
  39. Spek, A.L. PLATON, An integrated tool for the analysis of the results of a single crystal structure determination. Acta Cryst. A 1990, 46, C34. [Google Scholar]
  40. Spek, A.L. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar]
  41. Bruno, I.J.; Cole, J.C.; Edgington, P.R.; Kessler, M.K.; Macrae, C.F.; McCabe, P.; Pearson, J.; Taylor, R. New software for searching the Cambridge structural database and visualizing crystal structures. Acta Cryst. 2002, B58, 389–397. [Google Scholar] [CrossRef] [PubMed]
  42. Bradenburg, K. DIAMOND, Release 3.1f; Crystal Impact GbR: Bonn, Germany, 2008. [Google Scholar]
Scheme 1. Structural formulae and abbreviations of the Schiff base ligands discussed in the text.
Scheme 1. Structural formulae and abbreviations of the Schiff base ligands discussed in the text.
Inorganics 08 00032 sch001
Figure 1. Partially labeled representation of the cation of complex 1. Color scheme: NiII: green; Cl: cyan; I: purple; O: red; N: blue; C: gray. H atoms are omitted for clarity.
Figure 1. Partially labeled representation of the cation of complex 1. Color scheme: NiII: green; Cl: cyan; I: purple; O: red; N: blue; C: gray. H atoms are omitted for clarity.
Inorganics 08 00032 g001
Figure 2. Crystallographically established coordination modes of nacb2− ligands present in complex 1. Color scheme as in Figure 1.
Figure 2. Crystallographically established coordination modes of nacb2− ligands present in complex 1. Color scheme as in Figure 1.
Inorganics 08 00032 g002
Figure 3. Space-filling representation of 1. Color scheme as in Figure 1.
Figure 3. Space-filling representation of 1. Color scheme as in Figure 1.
Inorganics 08 00032 g003
Figure 4. Complete, labelled core of the {Ni12} cluster. Color scheme as in Figure 1.
Figure 4. Complete, labelled core of the {Ni12} cluster. Color scheme as in Figure 1.
Inorganics 08 00032 g004
Figure 5. Metal topology of the {Ni12} cluster; the black dashed lines are virtual bonds to emphasize the edge- and vertex-sharing {Ni3} triangles.
Figure 5. Metal topology of the {Ni12} cluster; the black dashed lines are virtual bonds to emphasize the edge- and vertex-sharing {Ni3} triangles.
Inorganics 08 00032 g005
Figure 6. Temperature dependence of the χΜT product for complex 1 at 0.03 T. (inset) Plot of magnetization (M) versus field (H) for 1 at 2 K.
Figure 6. Temperature dependence of the χΜT product for complex 1 at 0.03 T. (inset) Plot of magnetization (M) versus field (H) for 1 at 2 K.
Inorganics 08 00032 g006
Figure 7. Excitation (1) and emission (2) spectra of complex 1 in the solid-state and at room temperature.
Figure 7. Excitation (1) and emission (2) spectra of complex 1 in the solid-state and at room temperature.
Inorganics 08 00032 g007
Table 1. Selected interatomic distances (Å) and angles (°) for complex 1.
Table 1. Selected interatomic distances (Å) and angles (°) for complex 1.
BondDistancesBondDistancesBondAnglesBondAngles
Ni1–O11.994(5)Ni7–O82.017(5)Ni3–I1–Ni278.6(4)Ni2–O8–Ni6116.0(2)
Ni1–O262.020(6)Ni7–O92.017(5)Ni3–I1–Ni181.6(4)Ni7–O8–Ni699.2(2)
Ni1–O152.087(5)Ni7–O232.114(5)Ni2–I1–Ni178.4(4)Ni8–O9–Ni7126.9(3)
Ni1–O32.151(6)Ni7–O172.128(5)Ni1–O1–Ni1096.6(2)Ni8–O9–Ni11112.0(3)
Ni1–I22.484(2)Ni7–O122.137(5)Ni1–O1–Ni2112.8(2)Ni7–O9–Ni11102.8(2)
Ni1–I12.643(1)Ni7–O142.176(5)Ni10–O1–Ni2104.7(2)Ni9–O10–Ni898.8(2)
Ni2–O12.005(5)Ni8–O212.005(6)Ni3–O2–Ni5113.3(2)Ni9–O11–Ni899.2(2)
Ni2–O82.007(5)Ni8–O92.006(5)Ni3–O2–Ni2107.2(2)Ni9–O12–Ni7113.7(2)
Ni2–O272.014(5)Ni8–O192.031(6)Ni5–O2–Ni2120.9(2)Ni10–O14–Ni7118.6(2)
Ni2–O22.057(5)Ni8–O112.113(6)Ni3–O3–Ni1103.8(2)Ni10–O15–Ni192.9(2)
Ni2–O162.134(5)Ni8–O102.138(6)Ni3–O4–Ni12101.5(2)Ni2–O16–Ni1095.3(2)
Ni2–I12.628(1)Ni8–I42.479(5)Ni3–O4–Ni497.5(2)Ni7–O17–Ni1191.2(2)
Ni3–O22.001(5)Ni9–N11.988(7)Ni12–O4–Ni499.4(2)Ni6–O23–Ni797.4(2)
Ni3–O42.040(6)Ni9–O102.025(6)Ni3–O5–Ni494.9(2)Ni6–O24–Ni5100.7(2)
Ni3–O52.061(5)Ni9–O112.039(6)Ni6–O7–Ni594.9(2)Ni3–O29–Ni1299.6(2)
Ni3–O292.063(5)Ni9–N22.092(8)Ni2–O8–Ni7119.0(2)Ni12–O30–Ni499.2(2)
Ni3–O32.150(5)Ni9–O122.117(5)
Ni3–I12.532(1)Ni9–I52.367(9)
Ni4–O311.996(6)Ni10–N31.994(6)
Ni4–N92.016(7)Ni10–O12.000(5)
Ni4–O42.049(5)Ni10–O152.026(5)
Ni4–O302.069(6)Ni10–O252.039(6)
Ni4–O52.113(5)Ni10–O142.126(5)
Ni4–I32.498(5)Ni10–O162.158(5)
Ni5–O22.011(5)Ni11–O202.015(6)
Ni5–O242.063(5)Ni11–O182.016(6)
Ni5–O282.064(6)Ni11–O92.019(6)
Ni5–O62.066(5)Ni11–O222.050(6)
Ni5–N82.093(7)Ni11–N52.115(9)
Ni5–O72.126(5)Ni11–O172.283(5)
Ni6–O241.997(5)Ni12–N62.024(7)
Ni6–O232.002(5)Ni12–O302.030(6)
Ni6–N42.013(6)Ni12–O42.045(6)
Ni6–O82.044(5)Ni12–N72.076(8)
Ni6–O132.052(5)Ni12–O292.081(5)
Ni6–O72.120(5)Ni12–I62.482(4)
Table 2. Crystallographic data for complex 1.
Table 2. Crystallographic data for complex 1.
Parameter1
Empirical formula C133H101N9O35Ni12Cl5I3
FW/g mol−13647.69
Temperature/K 150(2)
Crystal typeGreen plate
Crystal size/mm30.22 × 0.10 × 0.04
Crystal system Triclinic
Space group P-1
a/Å19.422(2)
b/Å22.654(3)
c25.321(3)
α115.783(4)
β92.992(5)
γ109.118(5)
Volume/Å39443(2)
Z2
ρcalc/g cm−31.283
μ/mm−11.786
F(000)3644
θ range/°3.65 to 25.03
RadiationMo Kα (λ = 0.71073)
Index ranges−23 ≤ h ≤ 22
−22 ≤ k ≤ 26
−30 ≤ l ≤ 30
Reflections collected134,933
Independent reflections32,481 (Rint = 0.0476)
Goodness-of-fit on F21.040
Final R indexes [I ≥ 2σ(I)] a,bR1 = 0.0872
wR2 = 0.2099
Final R indexes [all data]R1 = 0.1221
wR2 = 0.2361
ρ)max,min/e Å−31.613 and −1.594
a R1 = ∑(||Fo| − |Fc||)/∑|Fo|. bwR2 = [∑[w(Fo2Fc2)2]/∑[w(Fo2)2]1/2, w = 1/[σ2(Fo2) + (ap)2 + bp], where p = [max(Fo2, 0) + 2Fc2]/3.

Share and Cite

MDPI and ACS Style

Perlepe, P.S.; Pantelis, K.N.; Cunha-Silva, L.; Bekiari, V.; Escuer, A.; Stamatatos, T.C. Rare Nuclearities in Ni(II) Cluster Chemistry: An Unprecedented {Ni12} Nanosized Cage from the Use of N-Naphthalidene-2-Amino-5-Chlorobenzoic Acid. Inorganics 2020, 8, 32. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8050032

AMA Style

Perlepe PS, Pantelis KN, Cunha-Silva L, Bekiari V, Escuer A, Stamatatos TC. Rare Nuclearities in Ni(II) Cluster Chemistry: An Unprecedented {Ni12} Nanosized Cage from the Use of N-Naphthalidene-2-Amino-5-Chlorobenzoic Acid. Inorganics. 2020; 8(5):32. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8050032

Chicago/Turabian Style

Perlepe, Panagiota S., Konstantinos N. Pantelis, Luís Cunha-Silva, Vlasoula Bekiari, Albert Escuer, and Theocharis C. Stamatatos. 2020. "Rare Nuclearities in Ni(II) Cluster Chemistry: An Unprecedented {Ni12} Nanosized Cage from the Use of N-Naphthalidene-2-Amino-5-Chlorobenzoic Acid" Inorganics 8, no. 5: 32. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics8050032

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