Next Article in Journal
Preparing a Ca-Bi-O System by the Precipitation Method and Studying Its Intermediate Structural Properties for Applications in Water Treatment
Next Article in Special Issue
Neptunyl(VI) Nitrate Coordination Polymer with Bis(2-pyrrolidone) Linkers Highlighting Crystallographic Analogy and Solubility Difference in Actinyl(VI) Nitrates
Previous Article in Journal
Supramolecular Host–Guest Assemblies of [M6Cl14]2–, M = Mo, W, Clusters with γ-Cyclodextrin for the Development of CLUSPOMs
Previous Article in Special Issue
Phases and Interfaces in the Cr–Fe–Si Ternary System: X-ray Diffraction and Electron Microscopy Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of the Solvent on the Crystallographic and Magnetic Properties of Rhenium(IV) Complexes Based on 2,2′-Bipyrimidine Ligand

by
Adrián Sanchis-Perucho
,
Marta Orts-Arroyo
,
Nicolas Moliner
and
José Martínez-Lillo
*
Departament de Química Inorgànica/Instituto de Ciencia Molecular (ICMol), Universitat de València, c/Catedrático José Beltrán 2, Paterna, 46980 València, Spain
*
Author to whom correspondence should be addressed.
Inorganics 2023, 11(2), 78; https://doi.org/10.3390/inorganics11020078
Submission received: 29 December 2022 / Revised: 1 February 2023 / Accepted: 6 February 2023 / Published: 9 February 2023
(This article belongs to the Special Issue Inorganics: 10th Anniversary)

Abstract

:
Two solvated rhenium(IV) complexes with formula [ReCl4(bpym)]·MeCN (1) and [ReCl4(bpym)]·CH3COOH·H2O (2) (bpym = 2,2′-bipyrimidine) have been prepared and characterized by means of Fourier transform infrared spectroscopy (FT–IR), scanning electron microscopy and energy dispersive X-ray analysis (SEM–EDX), single-crystal X-ray diffraction (XRD) and SQUID magnetometer. 1 and 2 crystallize in the monoclinic system with space groups P21/n and P21/c, respectively. In both compounds, the Re(IV) ion is six-coordinate and bound to four chloride ions and two nitrogen atoms of a 2,2′-bipyrimidine molecule forming a distorted octahedral geometry around the metal ion. In the crystal packing of 1 and 2, intermolecular halogen⋯halogen and π⋯halogen-type interactions are present. Hydrogen bonds take place only in the crystal structure of 2. Both compounds exhibit a similar crystal framework based on halogen bonds. Variable temperature dc magnetic susceptibility measurements performed on microcrystalline samples of 1 and 2 show a similar magnetic behavior for both compounds, with antiferromagnetic exchange between the Re(IV) ions connected mainly through intermolecular Re-Cl⋯Cl-Re interactions.

Graphical Abstract

1. Introduction

The study of solvate formation in crystalline compounds is becoming an increasingly significant topic given the academic and industrial interest in elucidating the properties and variations in the morphology of different crystal forms [1,2]. In many cases, the crystallization solvent can explain the observed changes in physical properties of a crystalline material, and this investigation can assist in the ongoing efforts to improve the design of crystallization processes from solution [1,2,3,4,5].
In the coordination chemistry of Re(IV) (a 5d3 ion), the polymorphism and the effect of the solvent on the magnetic properties of Re(IV) complexes have been relatively little studied [6,7]. The mononuclear complexes based on halides of Re(IV) ion have been investigated during the past few decades and are characterized by large values of magnetic anisotropy and significant intermolecular magnetic exchange [8,9,10,11,12,13,14,15,16,17,18,19,20,21,22]. Indeed, these compounds can display relatively strong dipolar exchange through Re–X⋯X–Re type contacts (X = halogen), which can result in a magnetic order as, for instance, metamagnetism or spin-canting phenomenon [23,24,25]. Such studies have also revealed the occurrence of significant antiferromagnetic interactions between the paramagnetic Re(IV) metal ions [Re(IV) has three unpaired electrons and a 4A2g term as ground electronic state]. These interactions are not transmitted across a chemical bond but through the aforementioned intermolecular contacts. These exchange pathways are supported by DFT-type calculations on the mononuclear Re(IV) complexes, which showed that spin density from the metal ion is somewhat delocalized onto the peripheral atoms of the ligands [7]. Given that the intermolecular halogen⋯halogen distance depends on the countercation size, bulky cations such as tetraphenylarsonium (AsPh4)+ and tetra-n-butylammonium (NBu4)+ preclude the magnetic coupling between adjacent halides of Re(IV) ions due to the large X⋯X separation they generate [7]. In singular cases, an unusual Re–Cl⋯(H2O)⋯Cl–Re pathway leads these Re(IV) compounds to ferromagnetic interactions. Hence, this halogen⋯halogen separation can also be affected by solvent molecules. As such, the use of adequate solvents could provide deeper insights into the magnetic properties of these types of Re(IV) systems.
The neutral mononuclear [ReCl4(bpym)] complex has been previously investigated, showing remarkable properties [26]. It displays potent in vitro anti-proliferative activity against a series of cancer cells [27]. Regarding its magnetic properties, it exhibits magnetic ordering below 7.0 K through the spin-canting phenomenon, with coercive field (Hc) and remanent magnetization (Mr) values of 1750 G and 0.05 μB, respectively [28]. It has also been employed as a starting material for the preparation of rhenium-based compounds such as, for instance, the first heterodinuclear 2,2′-bipyrimidine-bridged complex magneto-structurally studied, in the formula [ReCl4(μ-bpym)NiBr2(H2O)2] [29], and in the chiral, photoluminescent and spin-canted compound of formula {CuReCl4(μ-Cl)(μ-pyz)[ReCl4(μ-bpym)]}·MeNO2 [30]. More recently, two one-dimensional coordination polymers of general formula {[ReCl4(μ-bpym)CuX2]·solvent}n have been characterized structurally and magnetically [31].
To develop our investigation on the effect of the solvent and the intermolecular interactions on the magnetic properties of Re(IV) compounds, we herein report the synthesis, crystal structure and magnetic properties of two solvated Re(IV) complexes based on the 2,2′-bipyrimidine ligand, with formula [ReCl4(bpym)]·MeCN (1) and [ReCl4(bpym)]·CH3COOH·H2O (2), whose properties are compared with those of the non-solvated complex.

2. Results and Discussion

2.1. Preparation of the Complexes

Ammonium hexachlororhenate and 2,2’-bipyrimidine (bpym) were stirred in N,N-dimethylformamide (DMF) for 4 h at 110 °C. After cooling, the dark brown solution was filtered and evaporated at 60 °C until the elimination of the solvent was complete. The brown residue was washed with acidified H2O and left to air dry. This solid was then shaken with dichloromethane for 30 min and the resulting solution was filtered. The insoluble residue was discarded. The filtrate was washed several times with portions of H2O in a separation funnel until the aqueous layer became almost colorless. The dark orange dichloromethane layer was then dried and the solvent evaporated under reduced pressure to leave an orange–yellow microcrystalline solid which was finally dried at 110 °C. The same compound can be obtained from the tetrabutylammonium salt instead of the ammonium salt in acetic acid. Compound 1 was obtained via the recrystallization in a N,N-dimethylformamide-acetonitrile mixture, followed by slow diffusion in isopropanol at room temperature. Compound 2 was prepared by means of ligand substitution reaction, by replacing the oxalate (ox) ligand by 2,2′-bipyrimidine in the [ReCl4(ox)]2- complex in a mixture based on acetic acid, which was heated to 100 °C and stirred for 4 h.

2.2. Description of the Crystal Structures

The crystal structures of 1 and 2 were studied through the single-crystal X-ray diffraction technique. Both compounds crystallized in the monoclinic crystal system with space groups P21/n and P21/c, respectively (Table 1). The non-solvated [ReCl4(bpym)] complex crystallizes in the orthorhombic space group P212121 [26]. The crystal structure of 1 is made up of neutral [ReCl4(bpym)] complexes and MeCN molecules, whereas the crystal structure of 2 is based on [ReCl4(bpym)], CH3COOH and H2O molecules, as shown in their respective asymmetric units (Figure 1 and Figure 2).
In the mononuclear [ReCl4(bpym)] complexes of 1 and 2, each Re(IV) ion is bonded to four chloride ions and two nitrogen atoms of a 2,2′-bipyrimidine molecule, forming a distorted octahedral geometry around the metal ion. In both compounds, the average values of the Re–N bond lengths [2.129(1) Å for 1 and 2.137(1) Å for 2] are shorter than those of the Re–Cl bond lengths [2.317(1) Å for 1 and 2.308(1) Å for 2], displaying values which are in agreement with those previously published for similar Re(IV) complexes [26,32]. It is worth noting that the Re-Cl bond lengths in axial positions are longer than those located in equatorial positions in both 1 and 2 (Figure 1 and Figure 2). The chelating 2,2′-bipyrimidine molecule is effectively planar in both compounds. However, in the non-solvated complex, the two pyrimidine rings show a tilted position with a dihedral angle between them of ca. 16.0(3)° [26]. In 1 and 2, the 2,2′-bipyrimidine ligand exhibits average C−C, N−N, and C−N bond length values in agreement with those found in similar bpym-based complexes containing 4d and 5d metal ions [33,34].
In the crystal packing of 1, intermolecular short Cl⋯Cl contacts of 3.559(1) Å [Cl(3)⋯Cl(7a) distance with (a) = x − 1, y, z] and 3.560(1) Å [Cl(2)⋯Cl(6b) distance with (b) = x − 1, y − 1, z] direct alternate chains formed by complexes of Re(1) and Re(2) ions, which grow along the crystallographic c axis (Figure 3). These chains are interlinked through C-H⋯N interactions involving neighboring bpym ligands [C(15c)⋯N(4) distance being ca. 3.50 Å; (c) = x, y − 1, z], and also with somewhat longer Cl⋯Cl contacts of ca. 3.667(1) Å, resulting in a corrugated layered structure that grows in the ac plane (Figure 3). Further C-H⋯N interactions [C(6d)⋯N(9) distance being ca. 3.62 Å; (d) = −x + 1, −y + 2, −z + 1] occur between [ReCl4(bpym)] complexes and MeCN solvent molecules. π⋯Cl contacts of ca. 3.283(1) Å can be considered halogen bonds, which, along with several weak C-H⋯Cl interactions, draw together a 3D crystal structure in compound 1.
In the crystal packing of 2, several O-H⋯O and O-H⋯N hydrogen bonds connect the solvent molecules CH3COOH and H2O to the [ReCl4(bpym)] complexes [O1w⋯N(7a) distance of ca. 2.866 Å; (a) = -x + 1, y-1/2, -z + 3/2] (Table 2). As in 1, chains formed through alternate complexes of Re(1) and Re(2) ions are generated by very short Cl⋯Cl contacts (Figure 3). In the case of 2, these types of halogen⋯halogen bonds are shorter than those found in 1 [the shortest one being Cl(3)⋯Cl(7b) with a distance of ca. 3.309 Å and (b) = x + 1, -y + 3/2, z + 1/2] (Figure 3). Further Cl⋯Cl contacts of approximately 3.708 Å, together with C-H⋯N interactions between bipyrimidine rings of adjacent [ReCl4(bpym)] complexes, link the chains forming 2D sheets [C(6c)⋯N(8) distance of ca. 3.507 Å; (c) = x, -y + 3/2, z + 1/2] (Figure 4).
In 2, there are π⋯Cl contacts involving bipyrimidine rings and chloride anions, which are of approximately 3.283(1) Å. There are also weak C-H⋯Cl interactions, which implicate the methyl group of the CH3COOH molecules and chloride anions of the [ReCl4(bpym)] complexes [C(18)⋯Cl(5d) distance of ca. 3.617 Å; (d) = x + 1, -y + 3/2, z-1/2]. These last intermolecular interactions contribute to stabilizing the crystal structure in compound 2 (Figure 4).

2.3. Hirshfeld Surface Analysis

Hirshfeld surfaces of the neutral [ReCl4(bpym)] complex in compounds 1 and 2 were calculated. The close intermolecular interactions were analyzed through the CrystalExplorer program and compared with those of the non-solvated [ReCl4(bpym)] complex [35,36,37]. CrystalExplorer calculates a series of surfaces which allow users to obtain both a qualitative and quantitative visualization of the main intermolecular interactions. In this case, taking place in 1 and 2, the shorter contacts are shown using a red color [38,39]. This is performed by mapping the distances from the surface to the nearest atom outside (de) and inside (di) of each surface and, at the same time, assuming a normalized contact distance (dnorm) that considers some limitations generated by the atomic radii [35,36,37]. The Hirshfeld surfaces for these bpym-based compounds are given in Figure 5, Figure 6 and Figure 7 and Figures S1–S3 (Supplementary Materials). According to their fingerprint plots, both 1 and 2 show a very similar percentage regarding the intermolecular Cl⋯Cl contacts between adjacent [ReCl4(bpym)] complexes (ca. 6.4% for 1 and ca. 6.3% for 2), which is consistent with the similar framework based on halogen bonds that both compounds exhibit (Figure 8). The Cl⋯H contacts, primarily connecting chloride anions and C-H groups of neighboring [ReCl4(bpym)] complexes in 1 and 2, and the methyl groups of CH3COOH molecules in the case of 2, are the main interactions observed on the Hirshfeld surfaces for both compounds, which cover ca. 34.9% for 1 and ca. 39.2% for 2 on their respective fingerprint plots (Figure 5 and Figure 6). Finally, further N⋯H contacts involving solvent molecules, mainly MeCN (in 1) and H2O (in 2), and N atoms and C-H groups of adjacent bipyrimidine rings, are approximately 11.6% and 9.7% of the complete fingerprint plot of 1 and 2, respectively (Figure 5 and Figure 6). In Figure 6, a small asymmetric region close to the di values (1.2–1.4) is observed, which could be the result of some structural disorder, as previously reported [40,41,42,43]. The fingerprint plots for the non-solvated [ReCl4(bpym)] complex show a similar percentage for the intermolecular Cl⋯Cl contacts (ca. 6.0%) and a higher contribution of the Cl⋯H interactions (ca. 43.5%), involving chloride anions and C-H groups of bipyrimidine rings of adjacent [ReCl4(bpym)] complexes (Figure 7 and Figure S3).

2.4. Magnetic Properties

Dc magnetic susceptibility measurements were performed on microcrystalline samples of 1 and 2 in the 2–300 K temperature range and under an external magnetic field of 0.5 T. The resulting χMT versus T plots (χM being the molar magnetic susceptibility per mononuclear Re(IV) complex) for compounds 1 and 2 are shown in Figure 9. At 300 K, the χMT values are ca. 1.57 (1) and ca. 1.54 cm3mol−1K (2). These are very close to those previously published for magnetically isolated complexes based on Re(IV) ion (5d3 ion with S = 3/2) [7]. Upon cooling, the curve that the χMT values draw for 1 follows the Curie law, with decreasing temperature to approximately 100 K before they decrease reaching minimum values of approximately 0.17 cm3mol−1K at 2.0 K (Figure 9). In the case of 2, the χMT values decrease gradually with the decreasing temperature, and more abruptly at approximately 50 K, reaching a minimum value of 0.49 cm3mol−1K at 2.0 K (Figure 9). The decrease in the χMT value observed for both 1 and 2 is assignable to antiferromagnetic interactions as well as zero-field splitting (ZFS) effects [44,45,46]. No maximum of the magnetic susceptibility is observed for either compound, as shown in their respective χM vs. T plots (insets in Figure 9). Room temperature magnetization values for 1 and 2 are ca. 1.55 and ca. 1.58 µB, respectively. They are very close to those values of earlier published mononuclear Re(IV) complexes [7].
As indicated in the description of the crystal structure of 1 and 2, both compounds exhibit short Cl⋯Cl contacts between the paramagnetic [ReCl4(bpym)] complexes in their crystal lattices (Figure 3). Hence, these relatively important through-space interactions between Re(IV) ions precludes the occurrence of SIM behavior [44,45]. Additionally, the presence of solvent molecules drastically changes the magnetic properties, given that the unsolvated [ReCl4(bpym)] complex exhibits magnetic ordering through spin-canting. As previously reported [7,28], this is a magnetic behavior observed in neither 1 nor in 2.
H ^ = D [ ( S ^ Z ) 2 S ( S + 1 ) / 3 ] + g β H S ^
χ M =   χ / / + 2   χ 3 χ / / =   N β 2 g / / 2 4 k   ( T θ )   1 + 9 exp   ( 2 D / k T ) 1 + exp   ( 2 D / k T ) χ =   N β 2 g 2 k   ( T θ )   1 + ( 3 k T / 4 D )   [ 1 exp   ( 2 D / k T ) ] 1 + exp   ( 2 D / k T )
To analyze the magnetic behavior of 1 and 2, we employed the Hamiltonian of Equation (1) and its derived theoretical expression for the magnetic susceptibility of Equation (2), by including a θ term to account for the observed intermolecular interactions, where Ŝz is the easy-axis spin operator, H is the applied field, β is the Bohr magneton, the g parameter is the Landé factor and D is the zero-field splitting (ZFS) for the Re(IV) ion [7]. As previously reported, the zero-field splitting values for Re(IV) (a 5d3 ion) are very large in a distorted octahedral environment because of the high value of the spin-orbit coupling constant (λ ca. 1000 cm−1 in the free ion). In fact, for a six-coordinate Re(IV) ion, as in compounds 1 and 2, the first excited term which arises from the 4F free-ion ground term is 4T2g. Under a tetragonal distortion, this excited state is split into orbital singlet 4B2 and orbital doublet 4E at energies Δ and Δ, respectively [7]. Under this pattern, the interaction of these two excited terms, with the quartet ground spin state, leads to two Kramers doublets, |±3/2〉 and |±1/2〉, which are separated by an energy gap of |2D|; that is, the zero-field splitting resulting from the combined action of second-order spin–orbit interaction and the tetragonal crystal field.
Thus, the first term in Equation (1) corresponds to the ZFS and the second term to the Zeeman effect. Additionally, we have assumed that g = g = g for complexes 1 and 2. The best least-squares fit gave the parameters D = 31.9 cm−1, g = 1.86, θ = −5.3 K and R = 2.94 × 10−5 for 1 and D = 35.8 cm−1, g = 1.83, θ = −8.8 K and R = 4.52 × 10−5 for 2 {R being the agreement factor defined as Σi[(χMT)iobs − (χMT)icalcd]2/[(χMT)iobs]2}. As shown in Figure 9, the calculated curves (solid red lines) reproduce the experimental magnetic data in the whole temperature range quite well. The D and g values calculated for 1 and 2 are in agreement with those earlier reported for mononuclear Re(IV) complexes [7,8]. The sign and magnitude of the θ values corroborate the presence of relatively strong antiferromagnetic exchanges between the Re(IV) ions connected through intermolecular Cl⋯Cl and π⋯Cl pathways (Figure 3), with the shorter interactions being in 2 [ca. 3.6 Å (in 1) versus ca. 3.3 Å (in 2)]. According to the literature, these are considered halogen bonds and result in stronger antiferromagnetic exchange [47,48].

3. Experimental Section

3.1. Materials

All manipulations were performed under aerobic conditions, using all solvents and chemicals as received. The Re(IV) precursors, [ReCl4(bpym)] and (NBu4)2[ReCl4(ox)], were prepared following their respective literature procedures [26,49].

3.2. Synthesis of the Complexes

3.2.1. Synthesis and Crystallization of [ReCl4(bpym)]·MeCN (1)

The preparation of 1 consisted of dissolving [ReCl4(bpym)] (0.05 mmol, 24.3 mg) in 2 mL of a N,N-dimethylformamide-acetonitrile (1:1, v/v) mixture, followed by a slow diffusion in isopropanol at room temperature. Brown crystals of 1 were thus obtained in a few days and were suitable for single-crystal X-ray diffraction studies. Yield: ca. 75%. Anal. Calcd. for C18H15N9Cl8Re2 (1): C, 21.3; H, 1.5; N, 12.4. Found: C, 20.9; H, 1.1; N, 12.3. SEM-EDX: a molar ratio of 1:4 for Re/Cl was found for 1 (Figure S4). IR (KBr pellet) peaks are observed at 3100 (m), 3067 (s), 2245 (m), 1578 (vs), 1541 (m), 1453 (w), 1407 (vs), 1269 (w), 1213 (m), 1113 (m), 1021 (s), 987 (w), 818 (m), 747 (s), 693 (m), 665 (s), 490 (w), 440 (w) cm−1 (Figure S5).

3.2.2. Synthesis and Crystallization of [ReCl4(bpym)]·CH3COOH·H2O (2)

The preparation of 2 consisted of the ligand substitution of the previously reported (NBu4)2[ReCl4(ox)] complex. (NBu4)2[ReCl4(ox)] (0.12 mmol, 110 mg) and 2,2’-bipyrimidine (0.25 mmol, 39.5 mg) were mixed in glacial acetic acid (4.0 mL). The mixture was heated to 100 °C and stirred for 4 h. The resulting solution was filtered while hot and left to evaporate at room temperature. Dark orange crystals of 2 were grown in less than 1 week and were suitable for X-ray diffraction data collection. Yield: ca. 60%. Anal. Calcd for C18H18N8O3Cl8Re2 (2): C, 20.6; H, 1.7; N, 10.7. Found: C, 21.0; H, 2.0; N, 10.6. SEM-EDX: a molar ratio of 1:4 for Re/Cl was found for 2 (Figure S6). IR (KBr pellet) peaks are observed at 3424 (br), 3091 (m), 3066 (m), 1711 (m), 1579 (vs), 1547 (m), 1452 (w), 1436 (w), 1406 (vs), 1303 (w), 1258 (m), 1211 (w), 1109 (m), 1024 (s), 989 (w), 877 (w), 819 (m), 748 (s), 695 (m), 669 (m), 617 (w), 515 (w) cm−1 (Figure S7).

3.3. X-ray Data Collection and Structure Refinement

X-ray diffraction data collection on single crystals of dimensions 0.28 × 0.21 × 0.21 (1) and 0.17 × 0.16 × 0.06 mm3 (2) was carried out on a Bruker D8 Venture diffractometer with PHOTON II detector, using monochromatized Mo-Kα radiation (λ = 0.71073 Å). Crystal parameters and refinement results for 1 and 2 are summarized in Table 1. The structures were solved by standard direct methods and subsequently completed by Fourier recycling using the SHELXTL [50] software packages and refined by the full-matrix least-squares refinements based on F2 with all observed reflections. The final graphical manipulations were performed with the DIAMOND [51] and CRYSTALMAKER [52] programs. Crystallographic data were deposited in the Cambridge Structural Data Centre (CCDC) with numbers 2233243 and 2233244 for 1 and 2, respectively.

3.4. Physical Measurements

Elemental analyses (C, H, N) were performed in an Elemental Analyzer CE Instruments CHNS1100 and the molar ratio between heavier elements was found by means of a Philips XL-30 scanning electron microscope (SEM-EDX), equipped with system of X-ray microanalysis, in the Central Service for the Support to Experimental Research (SCSIE) at the University of Valencia. Infrared spectra (IR) of 1 and 2 were recorded with a PerkinElmer Spectrum 65 FT-IR spectrometer in the 4000–400 cm−1 region. Variable-temperature, solid-state dc magnetic susceptibility data were collected on a Quantum Design MPMS-XL SQUID magnetometer equipped with a 5 T dc magnet. To keep the samples of compounds 1 and 2 both immobilized and well isolated from the moisture of the air at all times, the organic compound eicosene was employed. Experimental magnetic data were corrected for the diamagnetic contributions of both the sample holder and the eicosene. The diamagnetic contribution of the involved atoms was corrected using Pascal’s constants [53].

4. Conclusions

In summary, the synthesis, crystal structure and magnetic properties of two novel solvated Re(IV) complexes based on the 2,2’-bipyrimidine ligand, with formula [ReCl4(bpym)]·MeCN (1) and [ReCl4(bpym)]·CH3COOH·H2O (2), have been reported. In the crystal structures of 1 and 2, there are intermolecular halogen⋯halogen and π⋯halogen-type interactions present and, in compound 2 only, there exist O-H⋯O and O-H⋯N hydrogen bonds. Having taken into account the presence of Cl⋯Cl interactions and halogen bonds, both compounds exhibit corrugated, reticular crystal frameworks, which host the solvent molecules. The investigation of the magnetic properties of 1 and 2 through dc magnetic susceptibility measurements reveals a similar magnetic behavior, since both compounds display antiferromagnetic exchange couplings between neighboring Re(IV) ions. The magnetic properties can be drastically modified by the presence of solvent molecules, given that the non-solvated [ReCl4(bpym)] complex exhibits a very different magnetic behavior (that is, magnetic ordering) through spin-canting, as previously reported. As such, by changing suitable solvents, it is possible to tune the magnetic properties in this type of molecular-based Re(IV) compounds. Further work based on other halides of the Re(IV) ion and other solvents is in progress.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/inorganics11020078/s1, Figure S1. Fingerprint plots for the [ReCl4(bpym)] complex in 1highlighting the region assigned to the N⋯H interactions (left) and to the Cl⋯H interactions (right). Figure S2. Fingerprint plots for the [ReCl4(bpym)] complex in 2highlighting the region assigned to the O⋯H interactions (left) and to the Cl⋯H interactions (right). Figure S3. Fingerprint plots for the non-solvated [ReCl4(bpym)] complex highlighting the region assigned to the N⋯H interactions (left) and to the Cl⋯H interactions (right). Figure S4. SEM-EDX spectrum for compound 1. Figure S5. Infrared spectrum (FT-IR) for compound 1. Figure S6. SEM-EDX spectrum for compound 2. Figure S7. Infrared spectrum (FT-IR) for compound 2.

Author Contributions

Conceptualization, J.M.-L.; funding acquisition, J.M.-L.; methodology, A.S.-P., M.O.-A., N.M. and J.M.-L.; investigation, A.S.-P., M.O.-A., N.M. and J.M.-L.; formal analysis, A.S.-P., M.O.-A., N.M. and J.M.-L.; writing–original draft preparation, J.M.-L.; writing–review and editing, J.M.-L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Spanish Ministry of Science and Innovation (Grant numbers PID2019-109735GB-I00 and CEX2019-000919-M (Excellence Unit “María de Maeztu”)).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

A.S.-P. and M.O.-A. thank the Spanish “FPU fellowships” and “FPI fellowships” programs, respectively.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Price, C.P.; Glick, G.D.; Matzger, A.J. Dissecting the Behavior of a Promiscuous Solvate Former. Angew. Chem. Int. Ed. 2006, 45, 2062–2066. [Google Scholar] [CrossRef] [PubMed]
  2. Shan, N.; Zaworotko, M.J. The role of cocrystals in pharmaceutical science. Drug Discov. 2008, 13, 440–446. [Google Scholar] [CrossRef] [PubMed]
  3. Cruz-Cabeza, A.J.; Reutzel-Edens, S.M.; Bernstein, J. Facts and fictions about polymorphism. Chem. Soc. Rev. 2015, 44, 8619–8635. [Google Scholar] [CrossRef] [PubMed]
  4. Takieddin, K.; Khimyak, Y.Z.; Fábián, L. Prediction of Hydrate and Solvate Formation Using Statistical Models. Cryst. Growth Des. 2016, 16, 70–81. [Google Scholar] [CrossRef]
  5. Braga, D.; Casali, L.; Grepioni, F. The Relevance of Crystal Forms in the Pharmaceutical Field: Sword of Damocles or Innovation Tools? Int. J. Mol. Sci. 2022, 23, 9013. [Google Scholar] [CrossRef]
  6. Martínez-Lillo, J.; Armentano, D.; Mastropietro, T.F.; Julve, M.; Faus, J.; De Munno, G. Self-Assembled One- and Two-Dimensional Networks Based on NH2Me2[ReX5(DMF)] (X = Cl and Br) Species: Polymorphism and Supramolecular Isomerism in Re(IV) Compounds. Cryst. Growth Des. 2011, 11, 1733–1741. [Google Scholar] [CrossRef]
  7. Martínez-Lillo, J.; Faus, J.; Lloret, F.; Julve, J. Towards multifunctional magnetic systems through molecular-programmed self assembly of Re(IV) metalloligands. Coord. Chem. Rev. 2015, 289–290, 215–237. [Google Scholar] [CrossRef]
  8. Woodall, C.H.; Craig, G.A.; Prescimone, A.; Misek, M.; Cano, J.; Faus, J.; Probert, M.R.; Parsons, S.; Moggach, S.; Martínez-Lillo, J.; et al. Pressure induced enhancement of the magnetic ordering temperature in rhenium(IV) monomers. Nat. Commun. 2016, 7, 13870. [Google Scholar] [CrossRef]
  9. González, R.; Chiozzone, R.; Kremer, C.; Guerra, F.; De Munno, G.; Lloret, F.; Julve, M.; Faus, J. Magnetic Studies on Hexahalorhenate(IV) Salts of Ferrocenium Cations [Fe(C5R5)2]2[ReX6] (R = H, CH3; X = Cl, Br, I). Inorg. Chem. 2004, 43, 3013–3019. [Google Scholar] [CrossRef]
  10. Nelson, C.M.; Boyd, G.E.; Smith, W.T. Magnetochemistry of Technetium and Rhenium. J. Am. Chem. Soc. 1954, 76, 348–352. [Google Scholar] [CrossRef]
  11. Figgis, B.N.; Lewis, J.; Mabbs, F.E. The Magnetic Properties of Some d3-Complexes. J. Chem. Soc. 1961, 3138–3145. [Google Scholar] [CrossRef]
  12. Busey, R.; Sonder, E. Magnetic Susceptibility of Potassium Hexachlororhenate (IV) and Potassium Hexabromorhenate (IV) from 5° to 300 °K. J. Chem. Phys. 1962, 36, 93–97. [Google Scholar] [CrossRef]
  13. Rouschias, G.; Wilkinson, G. The chemistry of rhenium–nitrile complexes. J. Chem. Soc. A 1968, 489–496. [Google Scholar] [CrossRef]
  14. Mroziński, J. Magnetic properties of methylammonium hexachlororhenate (4) salts in the range 1.5–300 K. Bull. Pol. Acad. Sci. Chem. 1978, 26, 789–798. [Google Scholar]
  15. Mroziński, J. Low temperature magnetic properties of some methylammonium hexabromorhenate (4) salts. Bull. Pol. Acad. Sci. Chem. 1980, 28, 559–567. [Google Scholar]
  16. Reynolds, P.A.; Moubaraki, B.; Murray, K.S.; Cable, J.W.; Engelhardt, L.M.; Figgis, B.N. Metamagnetism in tetrachlorobis(N-phenylacetamidine)rhenium(IV). J. Chem. Soc. Dalton Trans. 1997, 263–268. [Google Scholar] [CrossRef]
  17. Małecka, J.; Jäger, L.; Wagner, C.; Mroziński, J. Preparation, crystal structure and properties of (Ph4P)2ReCl6·2CH3CN. Pol. J. Chem. 1998, 72, 1879–1885. [Google Scholar]
  18. Reynolds, P.A.; Figgis, B.N.; Martín y Marero, D.J. Magnetic structure and covalence in tetrachlorobis(N-phenylacetamidinato)rhenium(IV) by neutron diffraction. J. Chem. Soc. Dalton Trans. 1999, 945–950. [Google Scholar] [CrossRef]
  19. Thornton, P. 13 Manganese, technetium and rhenium. Annu. Rep. Prog. Chem. Sect. A Inorg. Chem. 2000, 96, 215–228. [Google Scholar] [CrossRef]
  20. Coronado, E.; Day, P. Magnetic Molecular Conductors. Chem. Rev. 2004, 104, 5419–5448. [Google Scholar] [CrossRef]
  21. Feng, X.; Liu, J.-L.; Pedersen, K.S.; Nehrkorn, J.; Schnegg, A.; Holldack, K.; Bendix, J.; Sigrist, M.; Mutka, H.; Samohvalov, D.; et al. Multifaceted magnetization dynamics in the mononuclear complex [ReIVCl4(CN)2]2−. Chem. Commun. 2016, 52, 12905–12908. [Google Scholar] [CrossRef] [PubMed]
  22. Pedersen, K.S.; Sigrist, M.; Sørensen, M.A.; Barra, A.-L.; Weyhermüller, T.; Piligkos, S.; Thuesen, C.A.; Vinum, M.G.; Mutka, H.; Weihe, H.; et al. [ReF6]2−: A Robust Module for the Design of Molecule-Based Magnetic Materials. Angew. Chem. Int. Ed. 2014, 53, 1351–1354. [Google Scholar] [CrossRef] [PubMed]
  23. González, R.; Chiozzone, R.; Kremer, C.; De Munno, G.; Nicolò, F.; Lloret, F.; Julve, M.; Faus, J. Magnetic Studies on Hexaiodorhenate(IV) Salts of Univalent Cations. Spin Canting and Magnetic Ordering in K2[ReI6] with Tc = 24 K. Inorg. Chem. 2003, 42, 2512–2518. [Google Scholar] [CrossRef] [PubMed]
  24. Martínez-Lillo, J.; Kong, J.; Barros, W.P.; Faus, J.; Julve, M.; Brechin, E.K. Metamagnetic behaviour in a new Cu(II)Re(IV) chain based on the hexachlororhenate(IV) anion. Chem. Commun. 2014, 50, 5840–5842. [Google Scholar] [CrossRef]
  25. Louis-Jean, J.; Balasekaran, S.M.; Lawler, K.V.; Sanchis-Perucho, A.; Martínez-Lillo, J.; Smith, D.; Forster, P.M.; Salamat, A.; Poineau, F. Coexistence of metamagnetism and slow relaxation of magnetization in ammonium hexafluoridorhenate. RSC Adv. 2021, 11, 6353–6360. [Google Scholar] [CrossRef]
  26. Chiozzone, R.; González, R.; Kremer, C.; Cerdá, M.F.; Armentano, D.; De Munno, G.; Martínez-Lillo, J.; Faus, J. A novel series of rhenium-bipyrimidine complexes: Synthesis, crystal structure and electrochemical properties. Dalton Trans. 2007, 6, 653–660. [Google Scholar] [CrossRef]
  27. Martínez-Lillo, J.; Mastropietro, T.F.; Lappano, R.; Madeo, A.; Alberto, M.E.; Russo, N.; Maggiolini, M.; De Munno, G. Rhenium(IV) compounds inducing apoptosis in cancer cells. Chem. Commun. 2011, 47, 5283–5285. [Google Scholar] [CrossRef]
  28. Martínez-Lillo, J.; Lloret, F.; Julve, M.; Faus, J. Spin canting in Re(IV) complexes: Magnetic properties of [ReX4(bpym)] (X = Cl and Br; bpym = 2,2′-bipyrimidine). J. Coord. Chem. 2008, 62, 92–99. [Google Scholar] [CrossRef]
  29. Martínez-Lillo, J.; Armentano, D.; De Munno, G.; Cano, J.; Lloret, F.; Julve, M.; Faus, J. First Magnetostructural Study on a Heterodinuclear 2,2′-Bipyrimidine-Bridged Complex. Inorg. Chem. 2011, 50, 12405–12407. [Google Scholar] [CrossRef]
  30. Martínez-Lillo, J.; Armentano, D.; Fortea-Pérez, F.R.; Stiriba, S.E.; De Munno, G.; Lloret, F.; Julve, M.; Faus, J. A Chiral, Photoluminescent, and Spin-Canted {CuIReIV2}n Branched Chain. Inorg. Chem. 2015, 54, 4594–4596. [Google Scholar] [CrossRef]
  31. Armentano, D.; Sanchis-Perucho, A.; Rojas-Dotti, C.; Martínez-Lillo, J. Halogen⋯halogen interactions in the self-assembly of one-dimensional 2,2′-bipyrimidine-based CuIIReIV systems. CrystEngComm 2018, 20, 4575–4581. [Google Scholar] [CrossRef]
  32. Martínez-Lillo, J.; Armentano, D.; De Munno, G.; Faus, J. Magneto-structural study on a series of rhenium(IV) complexes containing biimH2, pyim and bipy ligands. Polyhedron 2008, 27, 1447–1454. [Google Scholar] [CrossRef]
  33. Roy, N.; Sen, U.; Moharana, P.; Babu, L.T.; Kar, B.; Vardhan, S.; Sahoo, S.K.; Bose, B.; Paira, P. 2,2′-Bipyrimidine-based luminescent Ru(II)/Ir(III)–arene monometallic and homo- and hetero-bimetallic complexes for therapy against MDA-MB-468 and caco-2 cells. Dalton Trans. 2021, 50, 11725–11729. [Google Scholar] [CrossRef]
  34. Yao, S.-Y.; Ou, Y.-L.; Ye, B.-H. Asymmetric Synthesis of Enantiomerically Pure Mono- and Binuclear Bis(cyclometalated) Iridium(III) Complexes. Inorg. Chem. 2016, 55, 6018–6026. [Google Scholar] [CrossRef]
  35. Turner, M.J.; McKinnon, J.J.; Wolff, S.K.; Grimwood, D.J.; Spackman, P.R.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer 17; University of Western Australia: Perth, Australia, 2017. [Google Scholar]
  36. McKinnon, J.J.; Jayatilaka, D.; Spackman, M.A. Towards quantitative analysis of intermolecular interactions with Hirshfeld surfaces. Chem. Commun. 2007, 3814–3816. [Google Scholar] [CrossRef]
  37. Spackman, M.A.; Jayatilaka, D. Hirshfeld surface analysis. CrystEngComm 2009, 11, 19–32. [Google Scholar] [CrossRef]
  38. Spackman, P.R.; Turner, M.J.; McKinnon, J.J.; Wolff, S.K.; Grimwood, D.J.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer: A program for Hirshfeld surface analysis, visualization and quantitative analysis of molecular crystals. J. Appl. Cryst. 2021, 54, 1006–1011. [Google Scholar] [CrossRef]
  39. Li, S.; Bu, R.; Gou, R.-J.; Zhang, C. Hirshfeld Surface Method and Its Application in Energetic Crystals. Cryst. Growth Des. 2021, 21, 6619–6634. [Google Scholar] [CrossRef]
  40. Martin, A.D.; Britton, J.; Easun, T.L.; Blake, A.J.; Lewis, W.; Schröder, M. Hirshfeld Surface Investigation of Structure-Directing Interactions within Dipicolinic Acid Derivatives. Cryst. Growth Des. 2015, 15, 1697–1706. [Google Scholar] [CrossRef]
  41. Orts-Arroyo, M.; Castro, I.; Lloret, F.; Martínez-Lillo, J. Molecular Self-Assembly in a Family of Oxo-Bridged Dinuclear Ruthenium(IV) Systems. Cryst. Growth Des. 2020, 20, 2044–2056. [Google Scholar] [CrossRef]
  42. Sanchis-Perucho, A.; Orts-Arroyo, M.; Camús-Hernández, J.; Rojas-Dotti, C.; Escrivà, E.; Lloret, F.; Martínez-Lillo, J. Hexahalorhenate(IV) salts of protonated ciprofloxacin: Antibiotic-based single-ion magnets. CrystEngComm 2021, 23, 8579–8587. [Google Scholar] [CrossRef]
  43. Sanchis-Perucho, A.; Orts-Arroyo, M.; Castro, I.; Lloret, F.; Martínez-Lillo, J. Crystal polymorphism in 2,2’-bipyrimidine-based iridium(III) complexes. J. Coord. Chem. 2022, 75, 2495–2507. [Google Scholar] [CrossRef]
  44. McAdams, S.G.; Ariciu, A.M.; Kostopoulos, A.K.; Walsh, J.P.S.; Tuna, F. Molecular single-ion magnets based on lanthanides and actinides: Design considerations and new advances in the context of quantum technologies. Coord. Chem. Rev. 2017, 346, 216–239. [Google Scholar] [CrossRef] [Green Version]
  45. Zabala-Lekuona, A.; Seco, J.M.; Colacio, E. Single-Molecule Magnets: From Mn12-ac to dysprosium metallocenes, a travel in time. Coord. Chem. Rev. 2021, 441, 213984. [Google Scholar] [CrossRef]
  46. Uysal, Ş. The Synthesis and Characterization of Star Shaped Metal Complexes of Triazine Cored Schiff Bases: Their Thermal Decompositions and Magnetic Moment Values. J. Inorg. Organomet. Pol. Mat. 2011, 21, 291–296. [Google Scholar] [CrossRef]
  47. Batsanov, S.S. Van der Waals Radii of Elements. Inorg. Mater. 2001, 37, 871–885. [Google Scholar] [CrossRef]
  48. Álvarez, S. A cartography of the van der Waals territories. Dalton Trans. 2013, 42, 8617–8636. [Google Scholar] [CrossRef]
  49. Chiozzone, R.; González, R.; Kremer, C.; De Munno, G.; Cano, J.; Lloret, F.; Julve, M.; Faus, J. Synthesis, Crystal Structure, and Magnetic Properties of Tetraphenylarsonium Tetrachloro(oxalato)rhenate(IV) and Bis(2,2‘-bipyridine)tetrachloro(μ-oxalato)copper(II)rhenium(IV). Inorg. Chem. 1999, 38, 4745–4752. [Google Scholar] [CrossRef]
  50. SHELXTL-2013/4; Bruker Analytical X-ray Instruments. Bruker: Madison, WI, USA, 2013.
  51. Diamond, version 4.5.0; Crystal Impact GbR: Bonn, Germany, 2018.
  52. CrystalMaker, version 8.5.1; CrystalMaker Software Ltd.: Oxford, UK.
  53. Bain, G.A.; Berry, J.F. Diamagnetic Corrections and Pascal’s Constants. J. Chem. Educ. 2008, 85, 532–536. [Google Scholar] [CrossRef]
Figure 1. Detail of the mononuclear [ReCl4(bpym)] complex and solvent MeCN molecule in 1. Thermal ellipsoids are depicted at the 50% probability level.
Figure 1. Detail of the mononuclear [ReCl4(bpym)] complex and solvent MeCN molecule in 1. Thermal ellipsoids are depicted at the 50% probability level.
Inorganics 11 00078 g001
Figure 2. Detail of the mononuclear [ReCl4(bpym)] complex and solvent molecules in 2. Thermal ellipsoids are depicted at the 50% probability level. H-bonding interactions are highlighted as dashed lines.
Figure 2. Detail of the mononuclear [ReCl4(bpym)] complex and solvent molecules in 2. Thermal ellipsoids are depicted at the 50% probability level. H-bonding interactions are highlighted as dashed lines.
Inorganics 11 00078 g002
Figure 3. View along the crystallographic b axis of the one-dimensional motifs of [ReCl4(bpym)] complexes (polyhedron model) connected through Cl⋯Cl interactions in 1 (left); View along the crystallographic a axis of the one-dimensional motifs of [ReCl4(bpym)] complexes (polyhedron model) connected through Cl⋯Cl interactions in 2 (right). Solvent molecules and bpym ligand have been omitted for clarity.
Figure 3. View along the crystallographic b axis of the one-dimensional motifs of [ReCl4(bpym)] complexes (polyhedron model) connected through Cl⋯Cl interactions in 1 (left); View along the crystallographic a axis of the one-dimensional motifs of [ReCl4(bpym)] complexes (polyhedron model) connected through Cl⋯Cl interactions in 2 (right). Solvent molecules and bpym ligand have been omitted for clarity.
Inorganics 11 00078 g003
Figure 4. View along the crystallographic b axis of the intermolecular interactions between [ReCl4(bpym)] complexes and between [ReCl4(bpym)] complexes and MeCN molecules in 1 (left); View along the crystallographic a axis of the intermolecular interactions between [ReCl4(bpym)] complexes and CH3COOH and H2O molecules in 2 (right).
Figure 4. View along the crystallographic b axis of the intermolecular interactions between [ReCl4(bpym)] complexes and between [ReCl4(bpym)] complexes and MeCN molecules in 1 (left); View along the crystallographic a axis of the intermolecular interactions between [ReCl4(bpym)] complexes and CH3COOH and H2O molecules in 2 (right).
Inorganics 11 00078 g004
Figure 5. Hirshfeld surface mapped through dnorm function for 1 (left); Full fingerprint plot for the dinuclear Re(IV) compound 1 (right).
Figure 5. Hirshfeld surface mapped through dnorm function for 1 (left); Full fingerprint plot for the dinuclear Re(IV) compound 1 (right).
Inorganics 11 00078 g005
Figure 6. Hirshfeld surface mapped through dnorm function for 2 (left); Full fingerprint plot for the dinuclear Re(IV) compound 2 (right).
Figure 6. Hirshfeld surface mapped through dnorm function for 2 (left); Full fingerprint plot for the dinuclear Re(IV) compound 2 (right).
Inorganics 11 00078 g006
Figure 7. Hirshfeld surface mapped through dnorm function for the non-solvated [ReCl4(bpym)] complex (left); Full fingerprint plot for the non-solvated [ReCl4(bpym)] complex (right).
Figure 7. Hirshfeld surface mapped through dnorm function for the non-solvated [ReCl4(bpym)] complex (left); Full fingerprint plot for the non-solvated [ReCl4(bpym)] complex (right).
Inorganics 11 00078 g007
Figure 8. View along the crystallographic a axis of the grid formed by [ReCl4(bpym)] complexes connected through intermolecular Cl⋯Cl interactions acting as host of MeCN molecules in 1 (left); View along the crystallographic c axis of the grid formed by [ReCl4(bpym)] complexes connected through intermolecular Cl⋯Cl interactions acting as host of CH3COOH and H2O molecules in 2 (right).
Figure 8. View along the crystallographic a axis of the grid formed by [ReCl4(bpym)] complexes connected through intermolecular Cl⋯Cl interactions acting as host of MeCN molecules in 1 (left); View along the crystallographic c axis of the grid formed by [ReCl4(bpym)] complexes connected through intermolecular Cl⋯Cl interactions acting as host of CH3COOH and H2O molecules in 2 (right).
Inorganics 11 00078 g008
Figure 9. Thermal variation of the χMT product for compounds 1 (left) and 2 (right). The solid red line represents the theoretical fit of the experimental data and the inset shows the χM versus T plot.
Figure 9. Thermal variation of the χMT product for compounds 1 (left) and 2 (right). The solid red line represents the theoretical fit of the experimental data and the inset shows the χM versus T plot.
Inorganics 11 00078 g009
Table 1. Summary of the crystal data and structure refinement parameters for 1 and 2.
Table 1. Summary of the crystal data and structure refinement parameters for 1 and 2.
Compound12
CIF2,233,2432,233,244
FormulaC18H15Cl8N9Re2C18H18Cl8N8O3Re2
Fw/g mol−11013.391050.40
Temperature/K120(2)120(2)
Crystal systemmonoclinicmonoclinic
Space groupP21/nP21/c
a13.239(1)11.960(1)
b11.852(1)18.089(1)
c17.666(1)13.433(1)
α/°9090
β/°90.02(1)93.02(1)
γ/°9090
V32771.97(13)2902.30(10)
Z44
Dc/g cm−32.4282.404
μ(Mo-Kα)/mm−19.5269.110
F(000)18881968
Goodness-of-fit on F21.3071.061
R1 [I > 2σ(I)]/all data0.0221/0.02520.0377/0.0424
wR2 [I > 2σ(I)] /all data0.0649/0.07730.0793/0.0816
Table 2. Selected hydrogen-bonding interactions in 2.
Table 2. Selected hydrogen-bonding interactions in 2.
D-H⋯AD-H/ÅH⋯A/ÅD⋯A/Å(DHA)/°
O(1w)-H(1wB)⋯N(4)0.9271.91(1)2.804(1)162.7(1)
O(1w)-H(1wA)⋯N(7a)0.9351.97(1)2.866(1)160.8(1)
O(1w)-H(1wA)⋯N(8a)0.9352.67(1)3.301(1)125.4(1)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sanchis-Perucho, A.; Orts-Arroyo, M.; Moliner, N.; Martínez-Lillo, J. Effect of the Solvent on the Crystallographic and Magnetic Properties of Rhenium(IV) Complexes Based on 2,2′-Bipyrimidine Ligand. Inorganics 2023, 11, 78. https://doi.org/10.3390/inorganics11020078

AMA Style

Sanchis-Perucho A, Orts-Arroyo M, Moliner N, Martínez-Lillo J. Effect of the Solvent on the Crystallographic and Magnetic Properties of Rhenium(IV) Complexes Based on 2,2′-Bipyrimidine Ligand. Inorganics. 2023; 11(2):78. https://doi.org/10.3390/inorganics11020078

Chicago/Turabian Style

Sanchis-Perucho, Adrián, Marta Orts-Arroyo, Nicolas Moliner, and José Martínez-Lillo. 2023. "Effect of the Solvent on the Crystallographic and Magnetic Properties of Rhenium(IV) Complexes Based on 2,2′-Bipyrimidine Ligand" Inorganics 11, no. 2: 78. https://doi.org/10.3390/inorganics11020078

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