Next Article in Journal
Tetraruthenium Metallamacrocycles with Potentially Coordinating Appended Functionalities
Next Article in Special Issue
Exploring High-Symmetry Lanthanide-Functionalized Polyoxopalladates as Building Blocks for Quantum Computing
Previous Article in Journal
Self-Assembly in Polyoxometalate and Metal Coordination-Based Systems: Synthetic Approaches and Developments
Previous Article in Special Issue
Mononuclear Dysprosium(III) Complexes with Triphenylphosphine Oxide Ligands: Controlling the Coordination Environment and Magnetic Anisotropy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chiral, Heterometallic Lanthanide–Transition Metal Complexes by Design

1
Department of Chemistry, University of Copenhagen, Universitetparken 5, 2100 Copenhagen, Danmark
2
Institut für Physikalische Chemie, Universität Stuttgart, Pfaffenwaldring 55, D-70569 Stuttgart, Germany
*
Authors to whom correspondence should be addressed.
Inorganics 2018, 6(3), 72; https://doi.org/10.3390/inorganics6030072
Submission received: 15 June 2018 / Revised: 12 July 2018 / Accepted: 17 July 2018 / Published: 19 July 2018
(This article belongs to the Special Issue Magnetic Lanthanide Complexes)

Abstract

:
Achieving control over coordination geometries in lanthanide complexes remains a challenge to the coordination chemist. This is particularly the case in the field of molecule-based magnetism, where barriers for magnetic relaxation processes as well as tunneling pathways are strongly influenced by the lanthanide coordination geometry. Addressing the challenge of design of 4f-element coordination environments, the ubiquitous Ln(hfac)3 moieties have been shown to be applicable as Lewis acids coordinating transition metal acetylacetonates facially leading to simple, chiral lanthanide–transition metal heterodinuclear complexes. The broad scope of this approach is illustrated by the synthesis of a range of such complexes LnM: LnM(hfac)32-acac-O,O,O′)3 (Ln = La, Pr, Gd; M = Cr, Fe, Ga), with approximate three-fold symmetry. The complexes have been crystallographically characterized and exhibit polymorphism for some combinations of 4f and 3d metal centers. However, an isostructural set of systems spanning several lanthanides which exhibit spontaneous resolution in the orthorhombic Sohncke space group P212121 is presented here. The electronic structure and ensuing magnetic properties have been studied by EPR spectroscopy and magnetometry. The GdFe, PrFe, and PrCr complexes exhibit ferromagnetic coupling, while GdCr exhibits antiferromagnetic coupling. GdGa exhibits slow relaxation of the magnetization in applied static fields.

Graphical Abstract

1. Introduction

Recent years have seen a revival of lanthanide coordination chemistry, partially fueled by their promising magnetic properties [1,2,3]. Thus, mononuclear lanthanide complexes, polynuclear lanthanide complexes, lanthanides coordinating organic radicals, and mixed 3d–4f metal complexes have all yielded single molecule magnets [2,3,4,5,6]. Among them, only few exhibit chirality [7,8,9,10]. In many regards, lanthanide-based systems outperform molecular magnets based on transition metals, but in the context of magnetic properties, the flexibility of the lanthanide ions concerning coordination numbers and coordination geometries complicates matters in more than one sense. Firstly, as the ligand field splitting in lanthanide complexes is energetically subordinate to the interelectronic repulsion and spin-orbit coupling, the coordination geometries and ensuing ligand fields become determining for the magnetic properties. Hence, tuning of magnetic properties requires a degree of tailoring the coordination environments, which is often difficult to realize for lanthanides. Secondly, the pliant geometries around lanthanide ions hamper the design of highly symmetrical systems, which can act as building blocks for extended structures since any symmetry imposed by multidentate ligands is easily broken. Thus, although crystallographically strictly axial, trigonal, and tetragonal [11,12,13,14,15,16,17,18,19,20] lanthanide complexes have been obtained by the use of structure-directing ligands, these systems typically feature coordinatively saturated lanthanide centers with no obvious avenues towards extended structures which preserve the tailored symmetry. Addressing this challenge, we forward a simple route towards heterobinuclear transition metal–lanthanide complexes with near-axial symmetry.
The hexafluoroacetylacetonate (hfac) complexes of the trivalent lanthanide ions are ubiquitous starting materials. While the homoleptic neutral entities Ln(hfac)3 are unknown in the condensed phase, many neutral heteroleptic complexes Ln(hfac)3Ln with lanthanide coordination numbers ranging from 7 to 10 [21,22,23,24] have been structurally characterized. Among these, eightfold coordination dominates as exemplified by the aquo-complexes Ln(hfac)3(H2O)2 [25] (cf. Figure 1a). Indeed, the Lewis acidic fragment Ln(hfac)3 has been employed as a building block for several structures by ligation of bidentate transition metal complexes creating heterometallic bi- and polynuclear systems in a systematic fashion [26,27] where one simple system is illustrated in Figure 1b [28]. However, to date, only a single example of a triply bridged system based on the Ln(hfac)3 fragment, namely CuLa(μ2-acac-O,O,O′)2(μ-H2O) (hfac)3 (acac = acetylacetonate), entailing nine-coordination of the large La3+ ion, has been reported (cf. Figure 1c) [29].
Common to the previously studied systems is the geometric incompatibility of the transition metal and the lanthanide fragments, which inevitably lowers the overall symmetry to C1. Notable is also the system [(NiL)Gd(hfac)2(EtOH)] (L = μ2-1,1,1-tris(N-salicylideneaminomethyl)ethane) [30], where the trifold symmetric Schiff-base ligand provides too much steric encumbrance to preserve three hfac ligands on the lanthanide. Inspired by the above-mentioned CuLa(μ-acac)2(μ-H2O)(hfac)3 and by examples of triple bridging between two Ln(hfac)3 fragments in, e.g., (hfac)3Ln(μ-PyO)3Ln(hfac)3 (Ln = Eu, Dy; cf. Figure 1d) [31], we decided to pursue the possibility of facial coordination of sterically undemanding, threefold symmetric transition metal complexes terminated in hard oxygen donor ligands to Ln(hfac)3 fragments aiming at threefold symmetric dinuclear systems. This approach necessitates the use of non-competing and hence weakly coordinating solvents, imposing some restrictions on the suitable transition metal building blocks. Evidently, uncharged complexes with three mono-anionic, bidentate ligands meet both the symmetry and solubility requirements and pose an obvious starting point. Indeed, this approach proved feasible and focusing on the pervasive tris-acetylacetonates of the trivalent transition metal ions, a subset of such systems which indeed conforms to the designed threefold pseudosymmetry is presented here (Figure 2).

2. Results and Discussion

Synthesis of the face-sharing mixed 3d–4f metal acetylacetonates LnM(hfac)3(μ-acac)3 proceeds in a facile manner by dehydration of the Ln(hfac)3(H2O)2 precursor through azeotropic distillation with, e.g., benzene in the presence of the transition metal acetylacetonate. In this fashion, triply bridged binuclear complexes have been accessed with Ln = La through Tb. The method provides moderate to good yields (30–84%) of well-crystalline products, which, if needed, can be recrystallized from non-coordinating solvents (e.g., chloroform, hexane…). The heterobinuclear complexes exhibit polymorphism and in the present contribution, we have chosen to focus on an orthorhombic phase which has been obtained pure for Ln = La, Pr, Nd, Sm, Eu, Gd, and Tb with M = Cr, Fe, Ga. This phase can be reproducibly obtained by use of the crystallization procedure described in the experimental section. Other phases belonging to the monoclinic and triclinic crystal systems have been observed for some combinations of 3d and 4f metals, but these will not be discussed here.
The heterobinuclear complexes LnM(hfac)3(μ-acac)3 were obtained spontaneously resolved in the Sohncke space group P212121. The 65 Sohncke space groups (often called chiral space groups) contain only rotation or screw axes. They are thus the only groups in which an enantiopure compound crystallize, and hence also the only groups allowing spontaneous resolution. The resolution occurs by conglomerate formation as demonstrated by structure determination of both enantiomorphs of PrFe (cf. Supplementary Materials). Illustrations of the molecular structure and packing of Λ,Δ-GdFe are given in Figure 3. Tables S1–S3 report the relevant crystallographic data for all the described complexes.
Notably, of the more than 850 known structures containing the Ln(hfac)3 fragment only 38 crystallize in one of the Sohncke space groups and only two heterobimetallic Ln(hfac)3 complexes, both of which are Cu(II)-containing chains, have been found to show spontaneous resolution [32,33]. It must be concluded that the weak dispersion forces from the heavily fluorinated Ln(hfac)3 fragments are not efficient in enforcing chiral packings. In line with this conclusion, the majority of short inter-molecular contacts in the packing of LnM involve the acetylacetonate ligands, alone or in addition to the hexafluoroacetylacetonate ligands (see Supplementary Materials, Figure S2 for details). In all cases, the ligand sphere around the d- or p-block metal is ordered, but in some cases the coordination sphere around the lanthanide is disordered with both helicities represented. In the systems LaFe, PrFe, and GdFe, no disorder is observed and the chirality of the two metal centers is invariably opposite. We hypothesize that the size of the light-metal fragment governs the locking of the relative configurations around the metal centers and the concomitant order in the lanthanide coordination sphere. The metric parameters for select combinations of metal ions are collected in Table 1.
In the first column of Table 1, bond length ranges are given for the d-block metals, including Ga. The geometric perturbation of the M(acac)3 complexes upon ligation to the lanthanides is quite moderate as reflected by the minor changes in bond length from the parent systems: 1.9413–1.9645 Å (for Cr(acac)3), 1.986–2.004 Å (for Fe(acac)3), and 1.941–1.964 Å (for Ga(acac)3). Evident is also the higher geometric pliancy of Ga(III), and especially Fe(III) as compared to Cr(III), which is expected on the basis of the strong preference of Cr(III) for regular octahedral coordination. The second and third columns of Table 1 summarize the coordination geometry around the lanthanide ions. The observed contraction of ca. 6% in Ln–Oacac distances upon going from La to Gd is slightly accentuated over the corresponding contraction of the Ln–Ohfac distances. Again, the geometric rigidity of Cr(acac)3 differentiates this metalloligand from the iron and gallium analogs resulting in longer Ln–Oacac bonds for the Cr(III) systems.
Although the binuclear complexes do not have crystallographic threefold symmetry, they approximate this situation quite closely. Thus, angles between centroids of the three-oxygen donor faces of the distant hfac-ligator atoms, the bridging acac-ligator atoms, and the terminal acac-ligator atoms range from 176.10° to 177.59°, indicating near co-linearity between the approximate threefold axes of the metal centers. The intermetallic distances span the interval 3.321–3.514 Å with distinct dependence on the nature of both metal centers as also reflected in the first coordination sphere bond distances (vide supra). Despite the only partial charge on the bridging donor atoms, the intermetallic distances are relatively short compared to other triply bridged lanthanide–transition metal complexes as illustrated in Figure 4. The relative proximity of the metal ions as well as the possibility of obtaining isostructural members of the series with diamagnetic centers at either position renders these systems well-suited for the investigation of magnetic properties, in particular 3d–4f magnetic interactions.
The magnetic properties of all these compounds were investigated using dc and ac magnetometry. In Figure 5a, we report the χT product of all the investigated derivatives. The room temperature values of all the samples are consistent with the expected Curie constants for independent paramagnetic ions: LaCr: 2.04 (1.88); LaFe: 4.33 (4.38); GdCr: 9.55 (9.76); GdFe: 12.27 (12.26); GdGa: 7.76 (7.88); PrCr: 3.40 (3.48); PrFe: 5.89 (5.98); and PrGa: 1.69 (1.60). The χT product of LaFe, LaCr, and GdGa exhibits a small decrease on lowering temperature due to the relatively small magnetic anisotropy of both the trivalent transition metal ions and of the half-filled Gd3+ ion (cf. Supplementary Materials Figures S3 and S4). This allows a simple qualitative evaluation of the nature of the coupling that instead dominates the shape of the χT curves of the GdM (M = Ga, Fe, or Cr) derivatives: ferromagnetic (FM) in GdFe and antiferromagnetic (AFM) in GdCr. The nature of the coupling is more difficult to extract at a glance for the complexes containing Pr3+ due to its pronounced magnetic anisotropy. However, the low-temperature value of the χT product of PrGa (Figure 5a) goes to zero when the temperature is decreased, and the magnetization curves (Figure S5a) almost coincide at all temperatures, suggesting that the effect of the ligands is to stabilize the diamagnetic mJ = 0 ground state of Pr3+. No hysteresis was observed in any of the samples at the lowest accessible temperature (1.8 K).
Simulation and parametrization of the magnetic data was performed using a Hamiltonian with the general expression:
= i = 1 n { μ B g i H · S ^ i + D i S ^ z i 2 } + j   S ^ 1 · S ^ 2
where the first summation runs over all the paramagnetic centers and contains the Zeeman interaction and Zero Field Splitting (ZFS). The second term, only used for complexes containing two paramagnetic ions, is an isotropic ferromagnetic (j < 0) or antiferromagnetic (j > 0) coupling [34]. For Pr3+, the S quantum number in Equation (1) was replaced with the total angular momentum of the ground multiplet J = 4.
The parameters employed in the simulations are reported in Table 2. The errors in the parameter values were estimated to be on the last digit reported in Table 2. The evaluation was obtained varying the relevant parameters in the simulations and monitoring the resultant change in the reproduction of the data.
For LaCr and LaFe, an axial ZFS parameter (D) and an isotropic g factor were sufficient to reproduce the shape of both the χT and magnetization curves. The simulations of LaFe, LaCr, and GdGa did not allow for a unique determination of the sign of the ZFS. However, previous studies on Cr(acac)3 and Fe(acac)3 suggest a negative ZFS [35,36]; thus, we adopted that sign. As expected, the half-filled Fe3+ ion has a lower magnetic anisotropy and a g factor closer to 2 compared to the Cr3+ ion. The small anisotropy expected in GdGa required the use of a more sensitive technique to be accurately unraveled. We thus performed X-band EPR spectroscopy, obtaining the spectrum reported in Figure 6. The simulation provided an estimation of the magnitude of the ZFS but not its sign (the magnetic measurements on GdCr and GdFe suggest a positive ZFS, vide infra).
The ZFS parameters and g factors obtained for the mono-paramagnetic complexes were fixed for the simulation of the poly-paramagnetic species, allowing for the extraction of the coupling constants reported in Table 2. For the GdCr and GdFe derivatives, a positive ZFS of Gd3+ was found to yield an improved goodness of the simulation compared to the negative case. The coupling in LnFe is significantly weaker than the one in LnCr for both Ln = Gd and Pr. The characteristic feature of the AFM coupling in GdCr (the only derivative that exhibits AFM coupling) is the step-like shape of the magnetization curve in Figure 5: indeed, a field of ca. 26.2 kOe is required to force the magnetic moments to be aligned with the field. A similar behavior with a spin rearrangement corresponding to a ferri-to-ferromagnetic transition at a comparable field was observed both by susceptometry and XMCD for the fluoride-bridged [Dy(hfac)3(H2O)–CrF2(py)4–Dy(hfac)3(NO3)] [37]. The magnitude of the coupling constants is well-comparable with other lanthanide–transition metal complexes connected by monoatomic bridges [38,39].
The magnetic anisotropy of Pr3+ could not be fully modelled with the available data; thus, we tentatively simulated it using a large positive D parameter (80 cm−1) to mimic the diamagnetic ground state and an isotropic Landé factor (gJ = 4/5). The introduction of a coupling term allowed for satisfactory reproduction of the magnetic behavior of both PrCr and PrFe, revealing FM coupling in both compounds.
The ac measurements conducted on all the samples evidenced slow relaxation only in GdGa, reported in Figure S6 (B = 0–5000 Oe). The field scan at 1.8 K revealed the presence of a double peak in the imaginary part of the magnetic susceptibility (χ″), corresponding to two active relaxation processes (Figure S6b), as previously observed in a Gd–EDTA complex. [40] The available experimental frequency window allowed us only to monitor the slow process. A temperature scan with an optimum field (Figure S7) of 4000 Oe reveals that the complex exhibits a single relaxation time up to 2.8 K, conversely to Gd–EDTA. [40] The origin of the slow relaxation can be attributed either to quantum tunneling (QT) or to a phonon bottleneck since they have a similar ac response. [41] The field and temperature dependence of the slow process was consequently parametrized using the theoretical field dependence predicted for a single QT process (Figures S8 and S9) [41,42]. The extracted widths of the relaxation time (0.31 < α < 0.62, Figure S10) provide further evidence that the QT regime persists up to 2.8 K since this process is particularly sensitive to strains and disorder [43].

3. Materials and Methods

3.1. Chemicals

All chemicals and solvents were purchased from commercial sources and used without further purification. All the syntheses were performed in open air. [Ln(hfac)3·x(H2O)2] (Ln = La, Pr, Gd) (x = 2, 3) were synthesized starting from the respective lanthanide oxides by modified literature methods [44]. [M(acac)3] (M = Cr, Fe) were synthesized from transition metal salts according to the known procedures [45,46] while an adjusted procedure was used for M = Ga based on the synthesis of [Fe(acac)3]. The bulk samples used for the magnetic characterizations were verified by powder X-ray diffraction to be phase pure with diffractograms matching the theoretical ones for the P212121 structures (Supplementary Materials, Figures S11–S18).

3.2. Synthesis

Synthesis of [La(hfac)3Cr(acac)3] (LaCr). La(hfac)3∙3H2O (203.5 mg, 0.25 mmol) was boiled close to dryness while stirring in 20 mL of toluene in an Erlenmeyer flask. A new portion of 10 mL of toluene was added, and the solution was again boiled close to dryness. This was repeated two times yielding a pale yellow-brown solution. After the last portion of toluene had been boiled almost dry, the heat was turned off and a solution of [Cr(acac)3] (69.9 mg, 0.20 mmol) in 5 mL of dichloromethane was slowly added to the hot (ca. 60 °C) reaction mixture causing an immediate boil off of dichloromethane. The resulting purple solution was allowed to cool down to room temperature (RT) without stirring, whereupon 10 mL of petroleum spirit was added, and the solution was filtered through a glass frit to remove any unreacted compounds. The solution was transferred to a crystallization bowl with a lid and stored in a refrigerator overnight to yield dark purple crystals of LaCr. To enhance the quality of the crystals, the crude product was recrystallized by boiling in 20 mL of n-heptane, filtering the solution, and then storing it in a refrigerator overnight. Yield: ~53%. Anal. Calcd. for LaCr (C30H24CrF18LaO12, molecular weight (MW) = 1109.40 g·mol−1): C, 32.48%; H, 2.18%. Found: C, 32.48%; H, 2.45%.
Synthesis of [La(hfac)3Fe(acac)3] (LaFe). LaFe was synthesized using a procedure similar to LaCr with the replacement of [Cr(acac)3] by [Fe(acac)3] (70.6 mg, 0.20 mmol) producing dark red crystals. Yield: ~63%. Anal. Calcd. for LaFe (C30H24F18FeLaO12, MW = 1116.33 g·mol−1): C, 32.36%; H, 2.17%. Found: C, 32.45%; H, 1.92%.
Synthesis of [Pr(hfac)3Cr(acac)3] (PrCr). PrCr was synthesized using a procedure similar to LaCr with the replacement of La(hfac)3∙3H2O by Pr(hfac)3∙3H2O (204.0 mg, 0.25 mmol) producing dark purple crystals. Yield: ~37%. Anal. Calcd. for PrCr (C30H24CrF18O12Pr, MW = 1111.40 g·mol−1): C, 32.42%; H, 2.18%. Found: C, 32.46%; H, 2.00%.
Synthesis of [Pr(hfac)3Fe(acac)3] (PrFe). PrFe was synthesized using a procedure similar to LaCr with the replacement of La(hfac)3∙3H2O by Pr(hfac)3∙3H2O (204.0 mg, 0.25 mmol) and [Cr(acac)3] by [Fe(acac)3] (70.6 mg, 0.20 mmol) producing dark red crystals. Yield: ~30%. Anal. Calcd. for PrFe (C30H24F18FeO12Pr, MW = 1115.25 g·mol−1): C, 32.31%; H, 2.18%. Found: C, 32.67%; H, 2.26%.
Synthesis of [Pr(hfac)3Ga(acac)3] (PrGa). PrGa was synthesized using a procedure similar to LaCr with the replacement of La(hfac)3∙3H2O by Pr(hfac)3∙3H2O (204.0 mg, 0.25 mmol) and [Cr(acac)3] by [Ga(acac)3] (73.4 mg, 0.20 mmol) producing faintly coloured yellow-green crystals. Yield: 35%. Anal. Calcd. for PrGa (C30H24F18GaO12Pr, MW = 1129.12 g·mol−1): C, 31.91%; H, 2.14%. Found: C, 31.86%; H, 2.28%.
Synthesis of [Gd(hfac)3Cr(acac)3] (GdCr). GdCr was synthesized using a procedure similar to LaCr with the replacement of La(hfac)3∙3H2O by Gd(hfac)3∙2H2O (203.6 mg, 0.25 mmol) producing dark purple crystals. Yield: ~60%. Anal. Calcd. for GdCr (C30H24CrF18GdO12, MW = 1127.74 g·mol−1): C, 31.95%; H, 2.15%. Found: C, 32.50%; H, 1.90%.
Synthesis of [Gd(hfac)3Fe(acac)3] (GdFe). GdFe was synthesized using a procedure similar to LaCr with the replacement of La(hfac)3∙3H2O by Gd(hfac)3∙2H2O (203.6 mg, 0.25 mmol) and [Cr(acac)3] by [Fe(acac)3] (70.6 mg, 0.20 mmol) producing dark red crystals. Yield: ~84%. Anal. Calcd. for GdFe (C30H24F18FeGdO12, MW = 1131.59 g·mol−1): C, 31.84%; H, 2.14%. Found: C, 31.81%; H, 2.30%.
Synthesis of [Gd(hfac)3Ga(acac)3] (GdGa). GdGa was synthesized using a procedure similar to LaCr with the replacement of La(hfac)3∙3H2O by Gd(hfac)3∙2H2O (203.6 mg, 0.25 mmol) and [Cr(acac)3] by [Ga(acac)3] (73.4 mg, 0.20 mmol) producing faintly coloured sandy crystals. Yield: ~52%. Anal. Calcd. for GdGa (C30H24F18GaGdO12, MW = 1145.46 g·mol−1): C, 31.45%; H, 2.11%. Found: C, 31.49%; H, 1.93%.

3.3. Single Crystal X-ray Diffraction Measurements

Single-crystal X-ray diffraction data were collected on a D8 VENTURE diffractometer (Bruker AXS, Karlsruhe, Germany) equipped with Mo Kα high brilliance IμS radiation (λ = 0.71073 Å), a multilayer X-ray mirror, a PHOTON 100 CMOS detector (Bruker AXS, Karlsruhe, Germany), and a low-temperature controller (Oxford Cryosystems, Oxford, UK). Crystals were mounted on loops made of Kapton. The diffractometer was controlled using the SAINT program as implemented in the APEX2 software package. Intensity data were corrected for absorption using the multi-scan method implemented in SADABS. The structures were solved using Olex2 by means of the olex2.structure solution program using a quasi-E charge flipping algorithm refined with the olex2.refine refinement package using Gauss–Newton minimization. Racemic twin refinements were carried out on all structures with a Flack parameter deviating more than 3 esds. from zero; in all cases, the crystals were determined not to be twinned with final Flack parameters equal to zero. For the structures with disordered hfac ligands, Shelx ISOR and DFIX restraints were applied to ensure chemically sensible metrics of the hfac ligands.

3.4. Powder X-ray Diffraction Measurements

Powder X-ray diffraction data were collected at room temperature on a D8 ADVANCE powder diffractometer (Bruker AXS, Karlsruhe, Germany) operating in a 2θ-θ configuration using Cu Kα radiation (λ = 1.5418 Å).

3.5. Magnetic Measurements

EPR spectra of GdGa were recorded on a microcrystalline powder sample using a Bruker Elexsys E500 equipped with a Bruker ER 4116 DM dual-mode cavity, an EIP 538B frequency counter and an ER035M NMR Gauss meter (Bruker Biospin, Rheinstetten, Germany). Effects related to crystallites’ preferential orientation were dimmed mediating 90 spectra recorded at different orientations. Ac and dc magnetometry measurements were performed using a Quantum Design MPMS-XL7 SQUID magnetometer (Quantum Design, Darmstadt, Germany). The samples were measured as pressed pellets wrapped into teflon. The data were corrected for diamagnetism using Pascal’s constants [47].

3.6. Simulations

All the simulations of the dc magnetic data were obtained using the Easy Spin package [48] of MATLAB. The ac susceptibility was fitted using a home-written program in MATLAB.

4. Conclusions and Outlook

A systematic building-block approach towards chiral and nearly axial heterobinuclear 3d–4f complexes has been forwarded. By employing sterically undemanding, threefold symmetric transition metal complexes it was found to be possible to extend the chemistry based on the Ln(hfac)3 fragment to encompass triply bridged systems with nine-coordinate lanthanide centers. The resulting systems provide for relatively short 3d–4f distances (3.320–3.514 Å) and concomitant easily quantifiable magnetic interactions. The class of systems introduced here are amenable to extension towards other transition metal building blocks. Such investigations are currently being undertaken, and it has already been verified that other motifs based on, e.g., Ru(III), Re(V) and Co(III) complexes are applicable. It is noteworthy that the crystallization as conglomerates and ensuing spontaneous resolution provides a rare avenue towards chiral Ln(hfac)3 fragments with potential for complete resolution based on adducts with robust chiral transition metal complexes.

Supplementary Materials

The following are available online at https://www.mdpi.com/2304-6740/6/3/72/s1: Tables S1–S3: Crystallographic data for the La-, Gd- and Pr-complexes, Checkcif for alternate enantiomorph of PrFe, Figure S1: Alternate views of the packing in GdFe, Figure S2: Shortest intermolecular interactions in the packing of GdFe, Figures S3–S5: Magnetization curves of LaM, GdM, and PrM recorded at T = 1.8, 5, and 10 K, Figures S6–S10: Ac susceptibility of GdGa, Figure S11–S18: Powder X-ray diffractograms of LaCr, LaFe, GdCr, GdFe, GdGa, PrCr, PrFe, and PrGa, Check cif files, and cif files.

Author Contributions

A.Ø., J.B., and M.V. synthetized the compounds and performed the chemical characterization. M.K. and J.v.S. facilitated the magnetic characterization. M.P. performed the magnetic characterization and data analysis. J.B. conceived and designed the project. The paper was written by A.Ø., J.B., and M.P. with inputs from all the authors.

Funding

This research was funded by the Independent Research Fund Denmark (#8021-00410B) and the Velux Foundations.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Goodwin, C.A.; Ortu, F.; Reta, D.; Chilton, N.F.; Mills, D.P. Molecular magnetic hysteresis at 60 kelvin in dysprosocenium. Nature 2017, 548, 439–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. McAdams, S.G.; Ariciu, A.-M.; Kostopoulos, A.K.; Walsh, J.P.; 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]
  3. Layfield, R.A.; Murugesu, M. Lanthanides and Actinides in Molecular Magnetism; John Wiley & Sons: Hoboken, NJ, USA, 2015. [Google Scholar]
  4. Sessoli, R.; Powell, A.K. Strategies towards single molecule magnets based on lanthanide ions. Coord. Chem. Rev. 2009, 253, 2328–2341. [Google Scholar] [CrossRef]
  5. Woodruff, D.N.; Winpenny, R.E.; Layfield, R.A. Lanthanide single-molecule magnets. Chem. Rev. 2013, 113, 5110–5148. [Google Scholar] [CrossRef] [PubMed]
  6. Pointillart, F.; Cador, O.; Le Guennic, B.; Ouahab, L. Uncommon lanthanide ions in purely 4f single molecule magnets. Coord. Chem. Rev. 2017, 346, 150–175. [Google Scholar] [CrossRef]
  7. Wang, Y.; Li, X.L.; Wang, T.W.; Song, Y.; You, X.Z. Slow relaxation processes and single-ion magnetic behaviors in Dysprosium-containing complexes. Inorg. Chem. 2009, 49, 969–976. [Google Scholar] [CrossRef] [PubMed]
  8. Li, X.L.; Chen, C.L.; Gao, Y.L.; Liu, X.L.; Feng, X.L.; Gui, Y.H.; Fang, S.M. Modulation of Homochiral DyIII Complexes: Single Molecule Magnets with Ferroelectric Properties. Chem. Eur. J. 2012, 18, 14632–14637. [Google Scholar] [CrossRef] [PubMed]
  9. Lin, P.H.; Korobov, I.; Wernsdorfer, W.; Ungur, L.; Chibotaru, L.F.; Murugesu, M. A rare μ4-O Centred Dy4 Tetrahedron with Coordination Induced Local Chirality and Single Molecule Magnet Behaviour. Eur. J. Inorg. Chem. 2011, 10, 1535–1539. [Google Scholar] [CrossRef]
  10. Mihalcea, I.; Perfetti, M.; Pineider, F.; Tesi, L.; Mereacre, V.; Wilhelm, F.; Rogalev, A.; Anson, C.E.; Powell, A.K.; Sessoli, R. Spin Helicity in Chiral Lanthanide Chains. Inorg. Chem. 2016, 55, 10068–10074. [Google Scholar] [CrossRef] [PubMed]
  11. Flanagan, B.M.; Bernhardt, P.V.; Krausz, E.R.; Lüthi, S.R.; Riley, M.J. A ligand-field analysis of the trensal (H3trensal = 2,2′,2″-tris(salicylideneimino)triethylamine) ligand. An application of the angular overlap model to lanthanides. Inorg. Chem. 2002, 41, 5024–5033. [Google Scholar] [CrossRef] [PubMed]
  12. Pedersen, K.S.; Dreiser, J.; Weihe, H.; Sibille, R.; Johannesen, H.V.; Sørensen, M.A.; Nielsen, B.E.; Sigrist, M.; Mutka, H.; Rols, S. Design of single-molecule magnets: Insufficiency of the anisotropy barrier as the sole criterion. Inorg. Chem. 2015, 54, 7600–7606. [Google Scholar] [CrossRef] [PubMed]
  13. Pedersen, K.S.; Ungur, L.; Sigrist, M.; Sundt, A.; Schau-Magnussen, M.; Vieru, V.; Mutka, H.; Rols, S.; Weihe, H.; Waldmann, O. Modifying the properties of 4f single-ion magnets by peripheral ligand functionalisation. Chem. Sci. 2014, 5, 1650–1660. [Google Scholar] [CrossRef] [Green Version]
  14. Perfetti, M.; Lucaccini, E.; Sorace, L.; Costes, J.P.; Sessoli, R. Determination of magnetic anisotropy in the lntrensal complexes (Ln = Tb, Dy, Er) by torque magnetometry. Inorg. Chem. 2015, 54, 3090–3092. [Google Scholar] [CrossRef] [PubMed]
  15. Lucaccini, E.; Baldoví, J.J.; Chelazzi, L.; Barra, A.-L.; Grepioni, F.; Costes, J.-P.; Sorace, L. Electronic structure and magnetic anisotropy in lanthanoid single-ion magnets with C3 symmetry: The Ln(trenovan) series. Inorg. Chem. 2017, 56, 4728–4738. [Google Scholar] [CrossRef] [PubMed]
  16. Clemente-Juan, J.M.; Coronado, E.; Gaita-Ariño, A. Magnetic polyoxometalates: From molecular magnetism to molecular spintronics and quantum computing. Chem. Soc. Rev. 2012, 41, 7464–7478. [Google Scholar] [CrossRef] [PubMed]
  17. Baldoví, J.J.; Cardona-Serra, S.; Clemente-Juan, J.M.; Coronado, E.; Gaita-Ariño, A.; Palii, A. Rational design of single-ion magnets and spin qubits based on mononuclear lanthanoid complexes. Inorg. Chem. 2012, 51, 12565–12574. [Google Scholar] [CrossRef] [PubMed]
  18. Ishikawa, N.; Sugita, M.; Wernsdorfer, W. Quantum tunneling of magnetization in lanthanide single-molecule magnets: Bis(phthalocyaninato)terbium and bis(phthalocyaninato)dysprosium anions. Angew. Chem. Int. Ed. 2005, 44, 2931–2935. [Google Scholar] [CrossRef] [PubMed]
  19. Sørensen, M.A.; Hansen, U.B.; Perfetti, M.; Pedersen, K.S.; Bartolomé, E.; Simeoni, G.G.; Mutka, H.; Rols, S.; Jeong, M.; Zivkovic, I.; et al. Chemical tunnel-splitting-engineering in a dysprosium-based molecular nanomagnet. Nat. Commun. 2018, 9, 1292. [Google Scholar] [CrossRef] [PubMed]
  20. Sørensen, M.A.; Weihe, H.; Vinum, M.G.; Mortensen, J.S.; Doerrer, L.H.; Bendix, J. Imposing high-symmetry and tuneable geometry on lanthanide centres with chelating Pt and Pd metalloligands. Chem. Sci. 2017, 8, 3566–3575. [Google Scholar] [CrossRef] [Green Version]
  21. Benelli, C.; Caneschi, A.; Gatteschi, D.; Pardi, L.; Rey, P. Structure and magnetic properties of linear chain complexes of rare earth ions (Gadolinium, Europium) with nitronyl nitroxides. Inorg. Chem. 1989, 28, 275–280. [Google Scholar] [CrossRef]
  22. Gleizes, A.N.; Senocq, F.; Julve, M.; Sanz, J.L.; Kuzmina, N.; Troyanov, S.; Malkerova, I.; Alikhanyan, A.; Ryazanov, M.; Rogachev, A.; et al. Heterobimetallic single-source precursors for MOCVD. Synthesis and characterization of volatile mixed ligand complexes of Lanthanides, Barium and Magnesium beta-diketonates with d-element containing ligands. J. Phys. IV 1999, 9, 943–951. [Google Scholar]
  23. Evans, W.J.; Giarikos, D.G.; Johnston, M.A.; Greci, M.A.; Ziller, J.W. Reactivity of the europium hexafluoroacetylacetonate (hfac) complex, Eu(hfac)3(diglyme), and related analogs with potassium: Formation of the fluoride hfac “ate” complexes, [LnF(hfac)3K(diglyme)]2. J. Chem. Soc. Dalton Trans. 2002, 520–526. [Google Scholar] [CrossRef]
  24. Sun, O.; Gao, T.; Sun, J.; Li, G.; Li, H.; Xu, H.; Wang, C.; Yan, P. A series of lanthanide(III) complexes constructed from schiff base and β-diketonate ligands. CrystEngComm 2014, 16, 10460–10468. [Google Scholar] [CrossRef]
  25. Kennedy, F.; Shavaleev, N.M.; Koullourou, T.; Bell, Z.R.; Jeffery, J.C.; Faulkner, S.; Ward, M.D. Sensitised near-infrared luminescence from lanthanide(III) centres using Re(I) and Pt(II) diimine complexes as energy donors in d–f dinuclear complexes based on 2,3-bis(2-pyridyl)pyrazine. Dalton Trans. 2007, 1492–1499. [Google Scholar] [CrossRef] [PubMed]
  26. Gleizes, A.; Julve, M.; Kuzmina, N.; Alikhanyan, A.; Lloret, F.; Malkerova, I.; Sanz, J.L.; Senocq, F. Heterobimetallic d–f metal complexes as potential single-source precursors for mocvd: Structure and thermodynamic study of the sublimation of [Ni(salen)Ln(hfa)3], Ln = Y, Gd. Eur. J. Inorg. Chem. 1998, 8, 1169–1174. [Google Scholar] [CrossRef]
  27. Pointillart, F.; Bernot, K.; Sessoli, R.; Gatteschi, D. Effects of 3d–4f magnetic exchange interactions on the dynamics of the magnetization of Dy(III)–M(II)–Dy(III) trinuclear clusters. Chem. Eur. J. 2007, 13, 1602–1609. [Google Scholar] [CrossRef] [PubMed]
  28. Ramade, I.; Kahn, O.; Jeannin, Y.; Robert, F. Design and magnetic properties of a magnetically isolated GdIIICuII pair. Crystal structures of [Gd(hfa)3Cu(salen)], [Y(hfa)3Cu(salen)], [Gd(hfa)3Cu(salen)(meim)], and [La(hfa)3(H2O)Cu(salen)][hfa = Hexafluoroacetylacetonato, salen = N,N′-Ethylenebis (salicylideneaminato), meim = 1-Methylimidazole]. Inorg. Chem. 1997, 36, 930–936. [Google Scholar]
  29. Rogachev, A.Y.; Mironov, A.V.; Nemukhin, A.V. Experimental and theoretical studies of the products of reaction between Ln(hfa)3 band Cu (acac)2 (Ln = La, Y; acac = acetylacetonate, hfa = hexafluoroacetylacetonate). J. Mol. Struct. 2007, 831, 46–54. [Google Scholar] [CrossRef]
  30. Yamaguchi, T.; Sunatsuki, Y.; Ishida, H.; Kojima, M.; Akashi, H.; Re, N.; Matsumoto, N.; Pochaba, A.; Mrozinski, J. Synthesis, structures, and magnetic properties of face-sharing heterodinuclear Ni(II)–Ln(III) (Ln = Eu, Gd, Tb, Dy) complexes. Inorg. Chem. 2008, 47, 5736–5745. [Google Scholar] [CrossRef] [PubMed]
  31. Yi, X.; Calvez, G.; Daiguebonne, C.; Guillou, O.; Bernot, K. Rational organization of lanthanide-based smm dimers into three-dimensional networks. Inorg. Chem. 2015, 54, 5213–5219. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, C.; Lin, S.-Y.; Shi, W.; Cheng, P.; Tang, J. Exploiting verdazyl radicals to assemble 2p–3d–4f one-dimensional chains. Dalton Trans. 2015, 44, 5364–5368. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, X.; Li, C.; Sun, J.; Li, L. Nitronyl nitroxide based 2p–3d–4f chains with the magnetocaloric effect and slow magnetic relaxation. Dalton Trans. 2015, 44, 18411–18417. [Google Scholar] [CrossRef] [PubMed]
  34. Gatteschi, D.; Sessoli, R.; Villain, J. Molecular Nanomagnets; Oxford University Press: Oxford, UK, 2006. [Google Scholar]
  35. Hansen, S.; Lehmann, G. EPR of Fe3+ in acetylacetonates. Appl. Magn. Reson. 1990, 1, 47–53. [Google Scholar] [CrossRef]
  36. Pritchard, B.; Autschbach, J. Theoretical investigation of paramagnetic NMR shifts in transition metal acetylacetonato complexes: Analysis of signs, magnitudes, and the role of the covalency of ligand–metal bonding. Inorg. Chem. 2012, 51, 8340–8351. [Google Scholar] [CrossRef] [PubMed]
  37. Dreiser, J.; Pedersen, K.S.; Piamonteze, C.; Rusponi, S.; Salman, Z.; Ali, M.E.; Schau-Magnussen, M.; Thuesen, C.A.; Piligkos, S.; Weihe, H. Direct observation of a ferri-to-ferromagnetic transition in a fluoride-bridged 3d–4f molecular cluster. Chem. Sci. 2012, 3, 1024–1032. [Google Scholar] [CrossRef] [Green Version]
  38. Cremades, E.; Gómez-Coca, S.; Aravena, D.; Alvarez, S.; Ruiz, E. Theoretical study of exchange coupling in 3d-Gd complexes: Large magnetocaloric effect systems. J. Am. Chem. Soc. 2012, 134, 10532–10542. [Google Scholar] [CrossRef] [PubMed]
  39. Pedersen, K.S.; Sørensen, M.A.; Bendix, J. Fluoride-coordination chemistry in molecular and low-dimensional magnetism. Coord. Chem. Rev. 2015, 299, 1–21. [Google Scholar] [CrossRef]
  40. Holmberg, R.J.; Ungur, L.; Korobkov, I.; Chibotaru, L.; Murugesu, M. Observation of unusual slow-relaxation of the magnetisation in a Gd–EDTA chelate. Dalton Trans. 2015, 44, 20321–20325. [Google Scholar] [CrossRef] [PubMed]
  41. Stamatatos, T.C.; Wernsdorfer, W.; Christou, G. Enhancing the Quantum Properties of Manganese–Lanthanide Single-Molecule Magnets: Observation of Quantum Tunneling Steps in the Hysteresis Loops of a {Mn12Gd} Cluster. Angew. Chem. Int. Ed. 2009, 48, 521–524. [Google Scholar] [CrossRef] [PubMed]
  42. Lucaccini, E.; Sorace, L.; Perfetti, M.; Costes, J.-P.; Sessoli, R. Beyond the anisotropy barrier: Slow relaxation of the magnetization in both easy-axis and easy-plane Ln(trensal) complexes. Chem. Commun. 2014, 50, 1648–1651. [Google Scholar] [CrossRef] [PubMed]
  43. Bernot, K.; Pointillart, F.; Rosa, P.; Etienne, M.; Sessoli, R.; Gatteschi, D. Single molecule magnet behaviour in robust dysprosium–biradical complexes. Chem. Commun. 2010, 46, 6458–6460. [Google Scholar] [CrossRef] [PubMed]
  44. Richardson, M.F.; Wagner, W.F.; Sands, D.E. Rare-earth trishexafluoroacetylacetonates and related compounds. J. Inorg. Nucl. Chem. 1968, 30, 1275–1289. [Google Scholar] [CrossRef]
  45. Glidewell, C. Inorganic Experiments; Wiley-VCH: Weinheim, Germany, 2003. [Google Scholar]
  46. Liu, R.; Van Rooyen, P.H.; Conradie, J. Geometrical isomers of tris (β-diketonato) metal(III) complexes for M = Cr or Co: Synthesis, X-ray structures and DFT study. Inorg. Chim. Acta 2016, 447, 59–65. [Google Scholar] [CrossRef]
  47. Bain, G.A.; Berry, J.F. Diamagnetic corrections and pascal’s constants. J. Chem. Educ. 2008, 85, 532–536. [Google Scholar] [CrossRef]
  48. Stoll, S.; Schweiger, A. Easyspin, a comprehensive software package for spectral simulation and analysis in EPR. J. Magn. Reson. 2006, 178, 42–55. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic structures of (a) the Ln(hfac)3(H2O)2 precursor; (b) Ln(hfac)3 ligated by bidentate Schiff-base complexes (M = Cu, Ni); (c) the hitherto sole structurally characterized example of a triply bridged Ln(hfac)3—transitional metal binuclear complex; and (d) triple bridging in homodinuclear (hfac)3Dy(μ-PyO)3Dy(hfac)3.
Figure 1. Schematic structures of (a) the Ln(hfac)3(H2O)2 precursor; (b) Ln(hfac)3 ligated by bidentate Schiff-base complexes (M = Cu, Ni); (c) the hitherto sole structurally characterized example of a triply bridged Ln(hfac)3—transitional metal binuclear complex; and (d) triple bridging in homodinuclear (hfac)3Dy(μ-PyO)3Dy(hfac)3.
Inorganics 06 00072 g001
Figure 2. Schematic structure of the targeted mixed lanthanide–transition metal acetylacetonates, LnM(hfac)3(μ-acac)3 (Ln = La, Pr, Gd; M = Cr, Fe, Ga).
Figure 2. Schematic structure of the targeted mixed lanthanide–transition metal acetylacetonates, LnM(hfac)3(μ-acac)3 (Ln = La, Pr, Gd; M = Cr, Fe, Ga).
Inorganics 06 00072 g002
Figure 3. (a) The molecular structure of Λ,Δ-GdFe in the solid state viewed along the approximate threefold axis of the single molecule constituting the asymmetric unit. Thermal ellipsoids represent a probability level of 50%. Disorder in one of the CF3 groups has been omitted for clarity. (b) Packing view along the crystallographic b-axis, emphasizing the four different molecular orientations dictated by the molecules residing at general positions. Hydrogen and fluorine atoms have been omitted for clarity. Color coding: Fe: purple, F: light green, O: red, C: gray, Gd: turquoise. For alternate views cf. Supplementary Materials, Figure S1. For metrics cf. Table 1.
Figure 3. (a) The molecular structure of Λ,Δ-GdFe in the solid state viewed along the approximate threefold axis of the single molecule constituting the asymmetric unit. Thermal ellipsoids represent a probability level of 50%. Disorder in one of the CF3 groups has been omitted for clarity. (b) Packing view along the crystallographic b-axis, emphasizing the four different molecular orientations dictated by the molecules residing at general positions. Hydrogen and fluorine atoms have been omitted for clarity. Color coding: Fe: purple, F: light green, O: red, C: gray, Gd: turquoise. For alternate views cf. Supplementary Materials, Figure S1. For metrics cf. Table 1.
Inorganics 06 00072 g003
Figure 4. Distribution of lanthanide–transition metal distances for structurally characterized, triply bridged polynuclear systems. The bar indicates the range for the present La, Pr, and Gd systems. The cross indicates the average value for all triply bridged La–transition metal structures. The data are based on entries in CSD ver. 5.39.
Figure 4. Distribution of lanthanide–transition metal distances for structurally characterized, triply bridged polynuclear systems. The bar indicates the range for the present La, Pr, and Gd systems. The cross indicates the average value for all triply bridged La–transition metal structures. The data are based on entries in CSD ver. 5.39.
Inorganics 06 00072 g004
Figure 5. (a) Temperature behavior of the χT product of all the synthetized samples in an applied field of 1 kOe. (b) Magnetization curves recorded at 1.8 K. Symbols are experimental data, solid lines are simulations (see text). For additional variable temperature magnetization data see Supplementary Materials, Figures S3–S5.
Figure 5. (a) Temperature behavior of the χT product of all the synthetized samples in an applied field of 1 kOe. (b) Magnetization curves recorded at 1.8 K. Symbols are experimental data, solid lines are simulations (see text). For additional variable temperature magnetization data see Supplementary Materials, Figures S3–S5.
Inorganics 06 00072 g005
Figure 6. X-band EPR spectrum of GdGa at room temperature in the solid state.
Figure 6. X-band EPR spectrum of GdGa at room temperature in the solid state.
Inorganics 06 00072 g006
Table 1. Metric data for select LnM(hfac)3(μ-acac)3 complexes.
Table 1. Metric data for select LnM(hfac)3(μ-acac)3 complexes.
ComplexM–OacacLn–OhfacLn–Oacac<O3>–<LnO3>–<MO3>/° aLn–M/Å
LaCr1.943–1.9712.454–2.5302.649–2.661176.103.420
LaFe1.962–2.0492.457–2.5032.629–2.636177.213.514
PrCr1.940–1.9682.415–2.4702.607–2.628176.053.380
PrFe1.958–2.0452.426–2.4672.587–2.606177.383.471
PrGa1.918–1.9832.406–2.4592.587–2.605176.633.425
GdCr1.936–1.9662.336–2.3992.535–2.565176.183.321
GdFe1.953–2.0512.356–2.4192.513–2.538177.593.406
GdGa1.921–1.9832.336–2.4062.516–2.530176.803.355
a <O3> denotes the centroid defined by the three remote oxygen ligators of the hfac ligands, <LnO3> the centroid defined by the three bridging oxygen ligators of the acetylacetonate ligands, and <MO3> the centroid of the three terminal oxygen ligators of the acetylacetonate ligands.
Table 2. Zero Field Splitting (ZFS) parameters, g factors, and j coupling extracted from the dc magnetic measurements.
Table 2. Zero Field Splitting (ZFS) parameters, g factors, and j coupling extracted from the dc magnetic measurements.
SamplegLngMDLn/cm−1DM/cm−1j/cm−1
LaFe-1.991-−0.75-
LaCr-1.965-−0.98-
GdGa *1.980-0.0465--
GdFe1.9801.9910.0465−0.75−0.38
GdCr1.9801.9650.0465−0.98+0.78
PrGa4/5-80--
PrFe4/51.99180−0.75−0.25
PrCr4/51.96580−0.98−1.2
* Combined fit of EPR and magnetic data.

Share and Cite

MDPI and ACS Style

Øwre, A.; Vinum, M.; Kern, M.; Van Slageren, J.; Bendix, J.; Perfetti, M. Chiral, Heterometallic Lanthanide–Transition Metal Complexes by Design. Inorganics 2018, 6, 72. https://doi.org/10.3390/inorganics6030072

AMA Style

Øwre A, Vinum M, Kern M, Van Slageren J, Bendix J, Perfetti M. Chiral, Heterometallic Lanthanide–Transition Metal Complexes by Design. Inorganics. 2018; 6(3):72. https://doi.org/10.3390/inorganics6030072

Chicago/Turabian Style

Øwre, Anders, Morten Vinum, Michal Kern, Joris Van Slageren, Jesper Bendix, and Mauro Perfetti. 2018. "Chiral, Heterometallic Lanthanide–Transition Metal Complexes by Design" Inorganics 6, no. 3: 72. https://doi.org/10.3390/inorganics6030072

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