Document not found! Please try again

High Nuclearity Assemblies and One-Dimensional (1D) Coordination

Oct 25, 2017 - Department of Chemistry, Purdue University, 560 Oval Drive, West Lafayette, Indiana 47907-2084, United States ... compound 1, or Ln = S...
0 downloads 12 Views 2MB Size
Article pubs.acs.org/IC

Cite This: Inorg. Chem. 2017, 56, 13152-13165

High Nuclearity Assemblies and One-Dimensional (1D) Coordination Polymers Based on Lanthanide−Copper 15-Metallacrown‑5 Complexes (LnIII = Pr, Nd, Sm, Eu) Anna V. Pavlishchuk,*,†,‡ Sergey V. Kolotilov,‡ Matthias Zeller,§,∇ Samuel E. Lofland,∥ Laurence K. Thompson,⊥ Anthony W. Addison,*,# and Allen D. Hunter§,∇ †

Department of Chemistry, Taras Shevchenko National University of Kyiv, Volodymyrska str. 62, Kiev 01601, Ukraine L.V. Pisarzhevskii Institute of Physical Chemistry of the National Academy of Sciences of Ukraine, Prospect Nauki 31, Kiev 03028, Ukraine § STaRBURSTT CyberInstrumentation Consortium and Department of Chemistry, Youngstown State University, One University Plaza, Youngstown, Ohio 44555-3663, United States ∇ Department of Chemistry, Purdue University, 560 Oval Drive, West Lafayette, Indiana 47907-2084, United States ∥ Department of Physics and Astronomy, Rowan University, Glassboro, New Jersey 08028, United States ⊥ Department of Chemistry, Memorial University, St. John’s, Newfoundland, A1B 3X7, Canada # Department of Chemistry, Drexel University, Philadelphia, Pennsylvania 19104-2816, United States ‡

S Supporting Information *

ABSTRACT: Complexes {[LnCu5(GlyHA)5(m-bdc)(H2O)4−x]2[LnCu5(GlyHA)5(SO4)(mbdc)(H2O)4]2}·(30 + 2x)H2O (where GlyHA2− = glycinehydroxamate, m-bdc2− = mphthalate; Ln = Pr and x = 0.21 for compound 1, or Ln = Sm and x = 0.24 for 3) and oned i m e n s i o na l ( 1 D ) c o o r d i n a t i on p ol y me r s { [N d C u 5 ( Gl y H A ) 5 (H 2 O) 5 ( mbdc)]nn[NdCu5(GlyHA)5(H2O)4(μ-CO3)(m-bdc)]}·13nH2O (2) and {[EuCu5(GlyHA)5(H2O)3](m-bdc)2[EuCu5(GlyHA)5(m-bdc)(H2O)3]}n·17nH2O (4) were obtained starting from the 15-metallacrown-5 complexes {[LnCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (Ln = Pr, Nd, Sm, Eu) by the partial or complete metathesis of sulfate anions with m-phthalate. Compounds 1 and 3 contain unprecedented quadrupledecker neutral metallacrown assemblies, where the [LnCu5(GlyHA)5]3+ cations are linked by m-phthalate dianions. In contrast, in complexes 2 and 4, these components assemble into 1D chains of coordination polymers, the adjacent {[NdCu5(GlyHA)5(H2O)5(m-bdc)]+}n 1D chains in 2 being separated by discrete [NdCu5(GlyHA)5(H2O)4(μ-CO3)(m-bdc)]}− complex anions. The crystal lattices of 2 and 4 contain voids filled by solvent molecules. Desolvated 4 is able to absorb up to 0.12 cm3/g of methanol vapor or 0.04 cm3/g of ethanol at 293 K. The isotherm for methanol absorption by compound 4 is consistent with a possible “gate opening” mechanism upon interaction with this substrate. The χMT vs T data for complexes 1−4 and their simpler starting materials {[LnCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (Ln(III) = Pr, Nd, Sm, Eu) were fitted using an additive model, which takes into account exchange interactions between lanthanide(III) and copper(II) ions in the metallamacrocycles via a molecular field model. The exchange interactions between adjacent Cu(II) ions in metallacrown fragments were found to fall in the range of −47 < JCu−Cu < −63 cm−1. These complexes are the first examples of a Ln(III)-Cu(II) 15-metallacrowns-5 (Ln(III) = Pr, Nd, Sm, Eu), for which values of exchange parameters have now been reported.



small-molecule substrates.18−22 It was found that counterion charge dramatically influenced the oligomerization of these 12metallacrown-4 blocks.18−21 In contrast to homometallic pentacopper(II) 12-metallacrown-4 complexes, only a limited number of reports devoted to coordination polymers and supramolecular systems with porous lattices based on lanthanide-containing 15-metallacrown-5 complexes have been published.23,24 We have recently demonstrated that both coordination polymers and discrete complexes can be isolated

INTRODUCTION Polynuclear 3d−4f metal complexes, including metallacrowns, have been attracting attention, because of their interesting properties, such as non-trivial magnetic behavior,1−6 catalytic activity,7 luminescence,8,9 possibility to serve as recognition agents,10,11 and their use in sensors.12,13 Such complexes could also be considered to be promising for the generation of multifunctional materials.14−17 It has previously been shown that pentacopper(II) 12metallacrown-4 complexes can be used as building blocks for the construction of supramolecular assemblies and coordination polymers with porous lattices, which display uptake of various © 2017 American Chemical Society

Received: July 31, 2017 Published: October 25, 2017 13152

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165

Article

Inorganic Chemistry

Scheme 1. Formation of Complexes 1−4 from Glycinehydroxamate Ln(III)-Cu(II) 15-metallacrowns-5 and m-Phthalate

solutions containing a 1:5:5 (equiv) ratio of Ln(NO3)3·xH2O, CuSO4· 5H2O, and NaHGlyHA.30 C, H, N microanalyses were carried out on a Carlo Erba 1106 instrument. Infrared (IR) spectra (in KBr pellets, 400−4000 cm−1) were obtained with a PerkinElmer Spectrum BX FTIR instrument. Single-crystal X-ray diffraction was performed at 100 K on a Bruker Smart Apex diffractometer with Smart or Apex2 software, using graphite-monochromated Mo Kα radiation with a wavelength of 0.71073 Å, while powder data were measured on a Bruker D8 Advance diffractometer (radiation wavelength = 1.54056 Å) at room temperature. Crystals suitable for X-ray data collection were taken directly from the reaction mixtures (CCDC 891994, 891995, 1552713, and 1552714; see Table S1 in the Supporting Information). Variable-temperature magnetic data (2−300 K) were obtained using a Quantum Design MPMS5S SQUID magnetometer with field strengths in the range of 0.1−1.0 T. Samples were prepared in gelatin capsules, mounted inside straws, and then fixed to the end of the sample transport rod. Background corrections for the sample holder assembly were applied. Susceptibility data were corrected for diamagnetism using Pascal’s constants,32 with Co[Hg(SCN)4] being used as a calibration standard. Thermogravimetric analysis (TGA) was performed with the samples in air, with an MOM Q1500 instrument at a heating rate of 10 °C min−1, in the range of 20−670 °C. Sorption measurements were performed by gravimetric methods at 293 K. Prior to all sorption measurements, samples were activated by desolvation at 130 °C at 10−2 Torr. Each point on the sorption and desorption isotherms corresponds to equilibrium conditions (constant sample weight at a given P/PS value). General Procedure of Synthesis of Complexes 1−4. A solution of {[LnCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (0.1 mmol) in 1.5 mL of DMF was added to a solution of sodium mphthalate (C8H4O4Na2, 63 mg, 0.3 mmol) in 3 mL of water and the mixture was stirred at 100 °C for 30 min. After filtration to remove some green precipitate that formed, the solution was allowed to cool, and the dark blue crystals that formed after 1 day were filtered off and air-dried. Information about composition, yields, masses of initial salts, and elemental analyses data is presented in Table S2 in the Supporting Information. The compositions of complexes 1−4 obtained from single-crystal Xray structure determination and from the elemental analysis data (Table S2) are slightly different. The latter, reflecting the bulk sample compositions, were used to describe the TG data, sorption, and magnetic properties, while the X-ray data were used for the structure descriptions. Compositions of bulk samples of the complexes calculated from elemental analyses and formulae obtained from single-crystal X-ray structure determination are given below.

as the outcomes of reactions between gadolinium(III)− copper(II) 15-metallacrowns-5 and different polycarboxylates.25 Desolvated samples of a 1D-coordination polymer assembled from Gd(III)−Cu(II) 15-metallacrown-5 and m-phthalate dianions have exhibited methanol sorption, typical of microporous sorbents.25 Special attention to lanthanide-based metallacrowns is justified by the possible occurrence of SMM behavior,26,27 luminescence27,28 and their potential for use in the creation of nonlinear optical materials.29 Because of the non-trivial magnetic behavior of metallacrowns, exchange parameter values have been reported only for gadolinium-containing systems,25 so the generation of models that could describe the magnetic properties of 15-metallacrowns-5 with different lanthanides(III) is another important task in this field. The objective of this study was to determine if there was any dependence of the composition and structure of the reaction products on the nature of the Ln(III) ion in the isostructural 15-metallacrown-5 complexes used as starting compounds, to study the magnetic and sorption properties of the products and make an effort to develop a theoretical model for the description of the magnetic properties of Ln(III)-Cu(II) metallacrowns with the lighter lanthanides. We focused our study on the interaction of isostructural 15-metallacrown-5 complexes {[LnCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O,30 containing Pr(III), Nd(III), Sm(III), and Eu(III) with mphthalate, which resulted in the formation of isomorphous discrete complexes in the case of Pr(III) and Sm(III) but two different 1D polymers for Nd(III) and Eu(III). The magnetic properties of complexes 1−4 and their simpler starting materials {[LnCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (Ln = Nd, Sm, and Eu) and the thermal stability of 4 and sorption properties of desolvated polycrystalline complex 4 are reported. Theoretical models, which allow fitting of the magnetic data to χMT vs T curves, are proposed for the first time for Pr(III), Nd(III), Sm(III), and Eu(III)-containing 15metallacrown-5 complexes.



EXPERIMENTAL SECTION

Materials and Measurements. Commercially available reagents and solvents (Merck and Aldrich) were used without further purification. The sodium salt of glycine hydroxamic acid was synthesized as described previously.31 The starting {[LnCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (Ln = Pr, Nd, Sm, and Eu) were obtained according to previously reported procedures as dark blue polycrystalline powders, via the slow evaporation of aqueous

1 {[PrCu5(GlyHA)5 ]4 (m‐bdc)4 (SO4 )2 (H 2O)46 } from CHN 1 {[PrCu5(GlyHA)5 ]4 (m‐bdc)4 (SO4 )2 (H 2O)46 } from X‐ray 13153

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165

Article

Inorganic Chemistry Table 1. Selected Characteristics of Crystal Structures of 1 and 3 1 PrCu5 in complex cation range of Ln···Cu separations, Å range of Cu···Cu separations, Å range of Ln−Oequat, Å range of Cu−Oequat, Å range of Cu−Nequat, Å range of τ values38 for pentacoordinate Cu(II) ions coordination number of Ln(III) ions average deviation among non-hydrogen atoms from Cu5 plane,a Å largest deviation among non-hydrogen atoms from Cu5 plane,a Å deviation of LnIII ion from Cu5 plane,a Å a

3 PrCu5 in complex anion

3.8748(6)−3.9848(6) 3.8739(6)−3.9242(7) 4.5656(9)−4.6106(6) 4.5674(8)−4.5954(7) 2.447(3)−2.549(3) 2.428(3)−2.488(3) 1.906(3)−1.966(3) 1.918(3)−1.968(3) 1.905(3)−2.041(3) 1.890(3)−2.018(3) 0.02−0.17 8 8 0.23 0.26

SmCu5 in complex cation

SmCu5 in complex anion

3.860(1)−3.965(1) 3.857(1)−3.905(1) 4.546(2)−4.594(1) 4.547(2)−4.581(2) 2.410(6)−2.523(7) 2.398(7)−2.469(7) 1.911(7)−1.945(7) 1.911(7)−1.960(7) 1.907(8)−2.017(9) 1.894(9)−2.018(8) 0.03−0.16 8 8 0.24 0.27

0.94

0.87

0.98

0.87

0.46

0.21

0.43

0.21

Deviation from the Cu1−Cu5 plane for LnCu5 of the complex anion or the Cu6−Cu10 plane for LnCu5 of the complex cation.

were replaced by m-phthalates (compound 4). Given the similarity of the reaction procedures for compounds 1−4, it is concluded that, although the thermodynamic stability of sulfato-Ln(III) complexes in solution is less than that of their carboxylate complexes,35 the isolation of different (e.g., mixed carboxylate/sulfate) products indicates that other factors (such as differing product solubilities) affect the ultimate outcomes. Attempts were made to analyze the reaction mixtures via mass spectrometry techniques (electrospray ionization (ESI), fast atom bombardment (FAB) and matrix-assisted laser desorption ionization (MALDI)) in order to understand which complex particles form during the metatheses in solution. However, because of the low solubility of the final compounds in volatile solvents, such as methanol or acetonitrile, and thus almost instantaneous commencement of crystallization of products, it was not possible to obtain informative mass spectra. Previously, it was noted that, in the case of 12-MC-4 complexes, there was some connection between the degree of the metallacrown blocks’ oligomerization and the charge of the counteranions: 1D chains had anions with a charge of −1, dimers or stacks of weakly bonded pentanuclear cations had anions with a charge of −2, and trimers had anions with a charge of −3.21 These dimers, trimers, or polymers were associated via Cu−O bond bridges, where O was an hydroxamate oxygen. From the present study, we can conclude that 15-MC-5 complexes can also be arranged in stacks, although oligomerization occurs through carboxylate bridges. This difference can be accounted by the higher affinity of LnIII ion for carboxylate, compared to CuII, and probably the higher charge (+3) of 15-metallacrowns-5, in comparison to 12metallacrown-4 complexes (+2). To the best of our knowledge, compounds 1 and 3 are the first examples of “tetramerization” of 15-MC-5 complexes. In contrast with compounds obtained from copper(II) 12-metallacrown-4 blocks and polycarboxylates,18−21 among which only discrete molecules were found, combination of LnCu5 15-metallacrown-5 blocks with carboxylates led to 1D-coordination polymers 2 and 4. Again, the high affinity of Ln(III) for carboxylate seems to be a driving factor for such 1D chain formation. General Structural Features of 15-Metallacrown-5 Units in 1−4. Complexes 1−4 are formed from 15metallacrown-5 [LnCu5(GlyHA)5]3+ building blocks, which are linked together by m-phthalates. Inasmuch as differences

2 {[NdCu5(GlyHA)5 ]2 (H 2O)26 (m‐bdc)2 (μ‐CO3)} from CHN 2 {[NdCu5(GlyHA)5 ]2 (H 2O)22 (m‐bdc)2 (μ‐CO3)} from X‐ray

3 {[SmCu5(GlyHA)5 ]4 (m‐bdc)4 (SO4 )2 (H 2O)44 } from CHN 3 {[SmCu5(GlyHA)5 ]4 (m‐bdc)4 (SO4 )2 (H 2O)46 } from X‐ray 4 {[EuCu5(GlyHA)5 ]2 (H 2O)19 (m‐bdc)3 }n from CHN



4 {[EuCu5(GlyHA)5 ]2 (H 2O)23 (m‐bdc)3 }n from X‐ray

RESULTS AND DISCUSSION Interaction of the previously reported hexanuclear copper lanthanide complexes (Scheme 1) {[LnCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O30 (Ln = Pr, Nd, Sm, and Eu, and GlyHA2− is the dianion of glycinehydroxamic acid, H2N− CH2−C(O)−NHOH) with sodium m-phthalate in a DMF− water solution resulted in the formation of a series of two isostructural quadruple-decker discrete complex pairs {[LnCu5(GlyHA)5(m-bdc)(H2O)4−x]2[LnCu5(GlyHA)5(SO4)(m-bdc)(H2O)4]2}·(30 + 2x)H2O (LnIII = Pr, x = 0.21 for compound 1; or LnIII = Sm, x = 0.24 for 3; m-bdc2− is mphthalate) and two 1D-coordination polymers of different composition and structure: {[NdCu5(GlyHA)5(H2O)5(mbdc)]nn[NdCu5(GlyHA)5(H2O)4(μ-CO3)(m-bdc)]}·13nH2O (2) and {[EuCu5(GlyHA)5(H2O)3](mbdc)2[EuCu5 (GlyHA) 5(m-bdc)(H 2O) 3]} n·17.11nH 2O (4). Syntheses of 1−4 were performed under identical conditions (i.e., the quantities of m-phthalate, the composition of solvent mixture, the temperature, the order of reagents’ combination, and the crystallization conditions were the same). For Ln = Pr or Sm, isostructural compounds 1 and 3 formed, as expected (there was only an insignificant difference in the content of solvating water in these complexes). However, the compositions and structures of complexes 2 and 4 were completely different from 1 (or 3), as well as from each other. Of the six sulfate dianions (per four LnCu5 formula units) in the starting materials containing Pr(III) or Sm(III), only four were replaced by the excess m-phthalate used in the reaction, leading to 1 or 3, respectively. In the case of Nd(III), four of six sulfate anions (per four NdCu5 units) were replaced by m-phthalate, while two others were substituted by carbonate anions formed due to atmospheric CO2 capture, similar to other reported instances.25,33,34 Finally, in the case of Eu(III), all the sulfate anions 13154

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165

Article

Inorganic Chemistry Table 2. Selected Characteristics of Crystal Structures of 2 and 4 4 2 range of Ln···Cu separations, Å range of Cu···Cu separations, Å range of Ln−Oequat, Å range of Cu−Oequat, Å range of Cu−Nequat, Å range of τ values38 for pentacoordinate Cu(II) ions coordination number of Ln(III) ions average deviation among non-hydrogen atoms from Cu5 plane,a Å largest deviation among non-hydrogen atoms from Cu5 plane,a Å deviation of LnIII ion from Cu5 plane,a Å a b

3.856(3)−3.9504(6) 4.57(1)−4.610(9) 2.44(2)−2.484(2) 1.920(9)−1.976(9) 1.899(3)−2.03(2) 0.013−0.118 8 0.23/0.24b 0.74/0.86b 0.38/0.56b

1st type of EuCu5

2nd type of EuCu5

3.866(1)−3.930(1) 3.832(1)−3.892(1) 4.507(2)−4.580(2) 4.505(2)−4.570(2) 2.388(6)−2.452(6) 2.403(6)−2.484(5) 1.906(5)−1.953(6) 1.905(5)−1.944(5) 1.842(7)−2.031(7) 1.874(7)−2.015(7) 0.003−0.203 8 8 0.37 0.25 1.12 0.57 0.34 0.47

Deviation from the Cu1−Cu5 plane for NdCu5 in 2 and for 2nd type of EuCu5 in 4 or from the Cu6−Cu10 plane for 1st type of EuCu5 in 4. Deviation from the Cu1BCu2BCu3BCu4Cu5 plane.

among the structures of 15-metallacrown-5 units in 1−4 are minor and the structures of the blocks are similar to those previously reported, 30 the intracrown structures of [LnCu5(GlyHA)5]3+ in 1−4 are only briefly described in this section. The metallacrown cations [LnCu5(GlyHA)5]3+ in 1−4 are formed by five Cu(II) ions bound in a metallamacrocyclic ring by five GlyHA2− dianions and the Ln(III) ion is bound in the center of the fragment via five hydroxamate O atoms. The coordination environment of the Cu(II) ions in the equatorial planes consists of two N atoms (one from the primary amino group and one from the doubly deprotonated hydroxamate group) and two O atoms (one from an N−O group and one hydroxamate carbonyl O). The equatorial Cu−O and Cu−N bond lengths in [LnCu5(GlyHA)5]3+ in 1−4 do not differ significantly, falling in the range 1.842(7)−2.041(3) Å (see Tables 1 and 2), consistent with the Cu−O and Cu−N bond lengths in previously reported Ln(III)-Cu(II) 15-metallacrown5 systems.23b,30,36,37 The Cu(II) ions in 1−4 are tetracoordinate or pentacoordinate. The fifth coordination position is usually occupied by an O atom from a carboxylate group or a water molecule (the lengths of these Cu−Ow bonds range from 2.31(5) Å to 2.85(4) Å), thus forming a N2O3 donor set. The τ38 values for the pentacoordinated Cu(II) ions in 1−4 range from 0 (square pyramidal) to 0.20 (somewhat trigonally distorted). The Ln(III) ions in 1−4 are octacoordinate. The equatorial Ln−O bond lengths in 1−4 fall in the range of 2.388(6) Å to 2.549(3) Å. The Ln···Cu and Cu···Cu separations in the metallacrown [LnCu5(GlyHA)5]3+ fragments in 1−4 are listed in Tables 1 and 2. The observed values are typical for 15metallacrown-5 complexes.23b,30,36,37 Generally, values of equatorial Cu−O, Cu−N, and Ln−O bond distances, as well as Ln···Cu and Cu···Cu separations show a tendency to decrease from 1 to 4, presumably as a consequence of the lanthanide contraction (Figure 1). A similar contraction was observed in the seven isostructural 15-metallacrown-5 complexes {[LnCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (Ln = Pr, Nd, Sm, Eu, Gd, Dy, Ho).30 The metallacrown units [LnCu5(GlyHA)5]3+ in 1−4 are slightly distorted from ideal planarity, the deviation of the Ln(III) ions from the Cu5 mean planes ranging from 0.21 Å to 0.56 Å. Both the largest and average deviations among the nonhydrogen atoms in 1−4 from the metallacrown mean planes Cu5 are in Tables 1 and 2. Selected bond distances and angles

Figure 1. Ranges of some bond distances and interatomic separations in 1−4 and complex {[(GdCu5(GlyHA)5(H2O)2)(GdCu5(GlyHA)5(H2O)3)(m-bdc)3]·16H2O}n,25 which is isostructural with 4.

for 1−4 are summarized in Tables S1−S4 in the Supporting Information. X-ray Structures of Complexes 1 and 3. Complexes 1 and 3 are isostructural (P1̅ space group), so their structures will be described in a general way and only significant differences will be addressed. Some unit-cell parameters (a, b, c, and V; see Table S1) decrease from 1 to 3, which is again attributable to the lanthanide contraction. Compounds 1 and 3 are centrosymmetric quadruple-decker neutral molecules, containing four [LnCu5(GlyHA)5]3+ metallacrown cations linked by m-phthalates (see Figures S1−S5 in the Supporting Information). The central cationic [LnCu5(GlyHA)5(m-bdc)(H2O)4−x]22+ “cores” of 1 and 3 are formed by two approximately planar metallacrown [LnCu5(GlyHA)5(H2O)4−x]3+ fragments, which are bridged by two m-phthalate dianions (see below). This pair of mphthalate bridges is coordinated to the Ln(III) ions through O atoms (see Figure 2) giving a 16-membered Ln2O4C10 ring. On the molecular periphery, the two approximately planar (see below) capping [LnCu 5 (GlyHA) 5 (SO 4 )(m-bdc)(H2 O) 4] − complex anions are each connected to the core [LnCu5(GlyHA)5(m-bdc)(H2O)4−x]22+ cationic unit by a quite nonsymmetric m-phthalate bridge. One of the O atoms on one end of the m-phthalate is tightly bound to the lanthanide atom of the capping layer, while the other O atom of the same carboxylate has a long Cu−O contact (2.925(4) Å for 1 and 2.93(1) Å for 3). 13155

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165

Article

Inorganic Chemistry

phthalate anions have occupancy factors of 0.5, so that each of these ligands coordinates to metal ions in only 50% of the metallacrown units; these ligations are mutually exclusive, with respect to any individual lanthanide ion. Half of the metallacrown units in 2 are involved in a 1D polymer, formed by adjacent metallacrown units linking by bridging m-phthalates (bound to Nd1 ions; see Figure 3) while the other 50% of the metallacrown cations are independent crowns, in which carbonate and m-phthalate are bound to Nd1 ions in nonbridging mode (see Figure 4). Figure 2. Fragment of the molecular structure of 1, showing one quadruple-decker unit. For clarity of presentation, H atoms and both coordinated and lattice water molecules are omitted from the figure. Broken lines indicate the 2.925(4) Å Cu−O contacts.

The coordination environments of the copper ions Cu1, Cu3, Cu4, Cu5, Cu6, Cu9, and Cu10 are completed by O atoms of the axially coordinated water molecules, thus forming N2O3 donor sets. The water oxygen atom O66w is disordered between two positionsO66w and O66Bw (with occupation factors of 0.207 and 0.236 in 1 and 3, respectively)so that the Cu9 copper ions possess both square-pyramidal N2O3 and square-planar N2O2 environments (20.7% and 23.6% of Cu9 ions in 1 and 3, respectively). The water molecules O66Bw are not coordinated to the Cu9 copper ions ((Cu9−O66Bw) = 3.64(2) Å in 1; 3.61(8) Å in 3). A similar solvent disorder situation was found previously for the complexes {[LnCu 5(GlyHA) 5(SO4 )(H2 O) 6.5 ]}2 (SO 4 )·6H 2 O,30 where 50% of a particular Cu2+ ion had axially coordinated water, and 50% did not. The length of these O−Cu bonds presumably betrays a weakness that leads to a frequent absence of the water molecule. Since there is no steric obstacle to coordination, this likely reflects a “cooperativity” between the strong equatorial and weak axial ligation and also invokes the question about the distinction between a “donor” atom being coordinated versus just occupying a location of minimum electrostatic energy. For the O26 of m-phthalate, which provides the connection between “central” and “peripheral” metallacrowns, at 2.925(4) Å from Cu8 for 1 and at 2.93(1) Å for 3, the latter situation would appear to be the case. The coordination environments of the ions Ln2 are completed by two carboxylate oxygen atomsO42 and O44 ((Pr2−O42) = 2.400(3) Å and (Pr2−O44) = 2.410(3) Å in 1; 2.362(7) Å and 2.363(6) Å in 3)from the two bridging mphthalate dianions and O46w atoms of the coordinated water molecules ((Pr2−O46w) = 2.461(3) Å in 1; 2.404(6) Å in 3). One of the carboxyl groups of each m-phthalate dianion is coordinated monodentately to the Ln2 ion of the first 15metallacrown-5 unit, while the second carboxyl group is similarly coordinated to the lanthanide ion of the neighboring “central” metallacrown, thus connecting two metallacrown fragments. Two coordination positions of Ln1 are occupied by two O atoms O21 and O22 of the bidentate sulfate dianions ((Pr1−O21) = 2.572(3) Å and (Pr1−O22) = 2.486(3) Å in 1; 2.517(7) Å and 2.422(7) Å in 3) and one is occupied by a carboxyl oxygen atom from the monodentate m-phthalate dianion ((Pr1−O25) = 2.395(3) Å in 1 and 2.343(6) Å in 3). X-ray Structure of Complex 2. The structure of compound 2 is significantly disordered. Part of the 15metallacrown-5 unit and some of the water molecules that are coordinated to the Cu(II) ions are disordered between two equally occupied positions. In addition, the carbonate and m-

Figure 3. Fragment of the structure of 2, showing the 1D chain {[NdCu5(GlyHA)5(H2O)5(m-bdc)]nn+, together with discrete complex anions [NdCu5(GlyHA)5(H2O)4(μ-CO3)(m-bdc)]−. For clarity of presentation, H atoms and both coordinated and lattice water molecules are omitted from the figure.

Figure 4. Fragment of the structure of 2, showing the chain formed by discrete complex anions {[NdCu5(GlyHA)5(H2O)4(μ-CO3)(mbdc)]}−. For clarity of presentation, H atoms and both coordinated and lattice water molecules are omitted from the figure.

The structure of compound 2 consists of positively charged 1D chains [NdCu5(GlyHA)5(H2O)5(m-bdc)]nn+ (Figure 3), separated by discrete anionic complexes [NdCu5(GlyHA)5(H2O)4(μ-CO3)(m-bdc)]− and noncoordinated water molecules. The [NdCu5(GlyHA)5(H2O)5(mbdc)]nn+ 1D chain is constructed from 15-metallacrown-5 cations [NdCu5(GlyHA)5(H2O)5]3+ (see Figure S6 in the Supporting Information) connected by bridging m-phthalates. 13156

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165

Article

Inorganic Chemistry One of the carboxylate groups of the m-phthalate is bidentately coordinated to Nd(III) ions via carboxyl oxygen atoms O18 and O19 ((Nd1−O18) = 2.603(5) Å and (Nd1−O19) = 2.854(6) Å), while the second carboxyl group is monodentately coordinated to Nd(III) ions from the adjacent 15-metallacrown-5 fragment via carboxyl oxygen atoms O16 ((Nd1− O16) = 2.37(1) Å). The octacoordination environment of Nd1 ions in the 1D chain compound {[NdCu5(GlyHA)5(H2O)5(mbdc)]}nn+ is completed by the oxygen atoms O28w ((Nd1− O28w) = 2.589(6) Å) from the coordinated water molecules. Each discrete anion [NdCu5(GlyHA)5(H2O)4(μ-CO3)(mbdc)] − consists of a 15-metallacrown-5 fragment [NdCu5(GlyHA)5(H2O)4]3+ (see Figure S7 in the Supporting Information), carbonate anion, which is bidentately bound to Nd1 via oxygen atoms O11 and O12 ((Nd1−O11) = 2.415(5) Å and (Nd1−O12) = 2.370(5) Å) and m-phthalate, which is coordinated to Nd1 ion through carboxyl oxygen atom O16B ((Nd1−O16B) = 2.54(1) Å). The Nd1 ions in the discrete complex anions [NdCu5(GlyHA)5(H2O)4(μ-CO3)(m-bdc)]− are also octacoordinated. The metallacrown units are partially disordered between two positions with occupancy factors of 0.5 (groups of adjacent atoms Cu1−Cu3; O1, O2, O4, O5; C1−C4; N1−N4 and N10) both in [NdCu 5 (GlyHA) 5 (H 2 O) 5 (m-bdc)] n n + and [NdCu5(GlyHA)5(H2O)4(μ-CO3)(m-bdc)]−. All the copper ions in the 15-metallacrown-5 units in 2 have square-pyramidal coordination (τ38 = 0.01−0.12) formed by two nitrogen atoms (from amino and hydroxamate groups) and two oxygen atoms (from carboxylate and hydroxamate groups) in the basal plane and oxygen atoms of water (O22w−O26w) from apically coordinated water molecules. Water oxygen atoms O22w, O23w, and O25w are disordered between two positions with occupancy factors of 0.5. Half of the O23w atoms are coordinated to Cu2 (d(Cu2−O23w) = 2.40(4) Å), while the rest are located 2.68(4) Å from Cu2B. In the case of O22w, located close to the disordered Cu1, the situation is more complicated: the distances between both positions occupied by disordered O22w and Cu1 are ∼2.3 and 2.6 Å, while Cu1B is more remote from both positions of O22w; d(Cu1B−O22w) and d(Cu1B−O22Bw) are ∼2.5 and 2.9 Å, respectively. Both of these “short” and “long” contacts exist in 2. The existence of such disorder evidence that coordination of water to Cu1 and Cu2 is not the main factor that governs localization of this O atom, similarly to a previously reported case.30 A PLATON39 estimation of solvent-accessible volume for the cell, from which all noncoordinated and coordinated solvent molecules are removed, gives a value of 17.3% (for a probe molecule with r = 1.4 Å). The space occupied by solvent molecules can be represented as 7.8 Å × 10.1 Å channels directed along the a-axis (estimated by taking into account the van der Waals radii of the corresponding atoms; see Figure S9 in the Supporting Information). X-ray Structure of Complex 4. Compound 4, which is a 1D-coordination polymer (Figure 5), is isostructural with the previously reported complex {[(GdCu5(GlyHA)5(H2O)2)(GdCu5(GlyHA)5(H2O)3)(m-bdc)3]·16H2O}n.25 1D chains of 4 are constructed from two types of alternating metallacrown cations [EuCu5(GlyHA)5(m-bdc)(H2O)3]+ (1st type of cation; see Figure S10 in the Supporting Information) and [EuCu5(GlyHA)5(H2O)3]3+ (2nd type of cation; see Figure S11 in the Supporting Information) linked by two different mphthalates. In addition to two m-phthalates, which connect two neighboring metallacrown units (hereinafter referred to as

Figure 5. Fragment of the structure of 4. For clarity of presentation, hydrogen atoms and both coordinated and lattice water molecules are omitted from the figure.

“bridging phthalates”), one more m-phthalate anion is bound to Eu1 and Cu5 ions from the same metallacrown (of the first type). One of the bridging m-phthalates is coordinated to the Eu1 and Eu2 ions in the [EuCu 5 (GlyHA) 5 (H 2 O) 3 ] 3+ and [EuCu5(GlyHA)5(m-bdc)(H2O)3]+ via two carboxylate oxygen atoms O25 and O28, respectively. Another bridging mphthalate also links Eu1 and Eu2 ions from two different metallacrown cations via O21 and O36 oxygen atoms and, in addition, is coordinated via O37 to the Cu6 ion in [EuCu5(GlyHA)5(H2O)3]3+, acting thus as a tridentate ligand. Apical positions of Eu1 in [EuCu5(GlyHA)5(m-bdc)(H2O)3]+ are occupied by oxygen atoms O21 and O25 ((Eu1−O21) = 2.346(6) Å and (Eu1−O25) = 2.403(6) Å) from two bridging m-phthalate dianions and by oxygen atom O29 ((Eu1−O29) = 2.398(6) Å) from m-phthalate dianion, which is bound to the Eu1 and Cu5 ions from the same metallacrown cation of the first type. Apical positions of Eu2 in [EuCu5(GlyHA)5(H2O)3]3+ cation are occupied by oxygen atoms O28 and O36 ((Eu2−O28) = 2.340(6) Å and (Eu2−O36) = 2.402(6) Å) from two different bridging m-phthalate dianions, while the third position is filled by oxygen atom O35w of the coordinated water molecule. The copper ions Cu2, Cu3, Cu7, and Cu10 possess square-planar N2O2 donor sets. The squarepyramidal coordination environments of the copper ions Cu1, Cu4, Cu8, and Cu9 are completed by oxygen atoms O24w, O23w, O33w, and O34w of the axially coordinated water molecules, thus possessing N2O3 donor sets. Copper ions Cu5 and Cu6 are also located in square-pyramidal N2O3 donor sets, but their apical positions are occupied by carboxyl oxygen atoms O30 and O37 ((Cu5−O30) = 2.629(7) Å and (Cu6− O37) = 2.422(6) Å) of m-phthalate dianions. Thus, in 4, there are three different m-phthalate dianions (see Figure S12 in the Supporting Information). Two of them act as linkers between adjacent metallacrown units, forming a 1D chain. Both these m-phthalates are coordinated to Eu1 and Eu2 ions, but one of them is additionally coordinated to a Cu6 ion from the [EuCu5(GlyHA)5(H2O)3]3+ cation. The third mphthalate is bidentately coordinated via one of the carboxylic groups to the Cu5 and Eu1 ions from same [EuCu5(GlyHA)5(m-bdc)(H2O)3]+ unit. The second carboxylate group of this m-phthalate remains noncoordinated, so that, of these three phthalates in 4, one is tridentate and two are bidentate. The voids (Figure S13 in the Supporting Information) between the 1D chains in 4 are filled by solvent water molecules. The PLATON-estimated39 total volume of solvent13157

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165

Article

Inorganic Chemistry accessible voids in 4 would be 18.6%, assuming that elimination of all solvent molecules does not lead to structure collapse. Since the 1D chains in the crystal lattice of 4 are not anchored by covalent bonds (but only hydrogen bonds), the desolvation of 4 should, however, result in collapse of its crystal structure. Thermal Analysis and Sorption Properties of Compound 4. The desolvation (activation) of polycrystalline 4 prior to sorption measurements was performed at 130 °C (10−2 Torr). Thermogravimetric analysis showed that it is possible to eliminate almost all the water (9.5% weight loss at 760 Torr, corresponding to 14 of the 19 water molecules) from 4 without chemical decomposition at this temperature. For desolvation of 4 in vacuo, we decreased the temperature from 240 °C (the temperature for complete water elimination from 4 at 760 Torr) to 130 °C. According to the thermogravimetric analysis of 4 (see Figure 6), weight loss is detectable, even at room temperature, rising

Figure 8. Isotherms (293 K) for (○) ethanol adsorption and (□) methanol sorption−desorption by desolvated complex 4.

rearrangement of the crystal structure of 4 upon interaction with this substrate. Similar isotherms have been observed previously,18−21,40 and the hysteresis in the sorption curves suggests structural rearrangements. The ethanol sorption capacity was only about one-third of that for MeOH (ca. 0.04 cm3/g at P/PS ≈ 1; 0.8 EtOH per EuCu5 unit). Unlike methanol sorption, ethanol sorption resembled a BDDT type I isotherm,41 implying micropore filling without lattice rearrangement. However, comparison of the isotherms for methanol and ethanol sorption by activated 4 leads to the conclusion that ethanol sorption is also associated with structural rearrangement of the framework of 4, as all rigid pores accessible to ethanol must have been easily filled by methanol. The difference between the methanol and ethanol sorption avidities can thus be taken into account for both via the difference in molecular size (kinetic diameters 3.6 and 4.5 Å, respectively42) and their interaction energy with 4, leading to different abilities to induce structural rearrangement. In contrast, ethanol sorption isotherms for supramolecular systems based on pentacopper(II) 12-metallacrowns-4 were similar to isotherms of methanol sorption.18−21 The isotherm for methanol sorption by the previously reported Gd(III)− Cu(II) 15-metallacrown-5 complex,25 which is isostructural to 4, was more akin to BDDT type I behavior.41 No conclusion about the absence of rigid pores in that case could be made,25 because the difference between methanol and ethanol sorption was consistent simply with their molecular sizes. Magnetic Properties. Despite numerous qualitative studies of lanthanide-copper 15-metallacrown-5 complexes, the quantitative magnetic properties of these compounds are still insufficiently defined, exchange parameters between paramagnetic centers having been determined only for gadolinium(III) 15-metallacrown-5 complexes.25 Fitting of the χMT vs T curves for polynuclear complexes for almost all lanthanides (except Gd(III), La(III), and Lu(III)) is significantly hindered by large magnetic anisotropy, spin−orbit coupling and/or population of higher energy levels (for Sm(III) and Eu(III)). We have previously demonstrated25 that two approaches may be used for estimation of magnetic parameters of gadolinium(III)−copper(II) metallacrowns. The first one is based on the isotropic spin Hamiltonian,

Figure 6. TG curve for compound 4 at atmospheric pressure.

to 2.8% at 67 °C, corresponding to the elimination of 4 water molecules per {EuCu5}2 unit. Further heating led to weight loss (10.8%) equivalent to 12 water molecules and the weight then plateaued between 156 °C and 240 °C (12.9% weight loss, corresponding to elimination of three additional H2O). Beyond 240 °C, abrupt weight loss (34% in total at 270 °C), is probably associated with the decomposition of 4. Drying polycrystalline 4 in vacuo at 130 °C resulted in a substantial decrease in crystallinity (Figure 7). Activated 4

Figure 7. Powder X-ray diffraction (XRD) (λ = 1.54056 Å) patterns for 4: (a) calculated from the single-crystal X-ray structure, (b) polycrystalline sample, and (c) the same sample after drying at 130 °C in vacuo.

absorbed up to 0.12 cm3/g of MeOH vapor (3.5 MeOH molecules per EuCu5 formula unit) at P/PS ≈ 1 at 293 K (Figure 8). The methanol sorption isotherm grew to 0.04 cm3/ g at P/PS ≈ 0.2 (where P is the given methanol pressure and PS its saturation vapor pressure at 293 K). Increasing P/PS to ∼0.5 did not lead to marked methanol sorption, but further growth of the sorption curve was indeed observed for 0.5 ≤ P/PS ≤ 1.0. The shape of the methanol sorption isotherm was not typical of classical physical adsorption but seems more consistent with

Ĥ (GdCu5) = − 2J1(S1·SGd + S2·SGd + S3·SGd + S4 · SGd + S5· SGd) − 2J2 (S1·S2 + S2·S3 + S3·S4 + S4 ·S5 + S5·S1)

(1)

where J1 is the exchange integral between Cu(II) and Gd(III), J2 is the exchange integral between each pair of adjacent peripheral Cu(II) ions in the metallacrown unit, SGd is the spin 13158

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165

Article

Inorganic Chemistry

Figure 9. χMT vs T plots for complexes (A) 1, (B) 2, and (C) {[NdCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (complex A in Tables 3 and 4). Data points (○) represent experimental data, line represents the best fit with parameters (from Table 4).

χMT with decreasing T can be associated with the depopulation of Stark sublevels (splitting of mJ sublevels of Ln(III) ions by weak crystal fields is often referred to as Stark sublevel generation; however, the use of this terminology is inconsistent)46 in the cases of Nd(III) and Pr(III) complexes or depopulation of low-lying excited states for Eu(III) and Sm(III). Antiferromagnetic interactions dominate between adjacent copper(II) ions in Gd(III)−Cu(II) metallacrowns (as in copper(II) 12-metallacrown-4 complexes).18−20,25 The experimental χMT vs T curves for complexes 1−4, and {[LnCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (Ln = Nd (A), Sm (B), Eu (C)), were fitted using the additive model described by eq 2:

operator for the Gd(III) ion, and Si are the spin operators for the Cu(II) ions. In the second approach, exchange interactions between adjacent copper(II) ions in the 15-metallacrown-5 unit were assumed to be dominant, while weak copper−gadolinium exchange interactions were taken into account via a molecular field model. Inasmuch as lanthanide ions (except Gd(III)) possess spin−orbit coupling, the former model based on that implicitly isotropic spin Hamiltonian could not be used for other lanthanide ions. Thus, the objective of this study was to see if the proposed additive model could be used to describe the magnetic properties of 15-metallacrown-5 with lanthanides possessing spin−orbit coupling. Magnetic susceptibility measurements in the temperature range of 3−300 K were performed for polycrystalline samples of complexes 1−4 and the initial complexes {[LnCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (Ln = Nd (A), Sm (B), Eu (C)) The χMT vs T curves for all these compounds are presented per LnCu5 unit (see Figures 9 and 10). The room-temperature values of χMT for complexes 1−4

χM T (LnCu5) = χM T (Ln) + χM T (Cu5)

(2)

Exchange interactions between neighboring Cu(II) ions were assumed to be dominant, while the lesser-magnitude exchange interactions between Ln(III) and Cu(II) ions were taken into account by a molecular field model:25,32 Ĥ (Cu 5II) = − 2JCu−Cu (S1̂ ·S2̂ + S2̂ ·S3̂ + S3̂ ·S4̂ + S4̂ ·S5̂ + S5̂ ·S1̂ ) (3)

χMF T =

χM T 1−

2zJ ′ χM NAg 2β 2

(4)

where JCu−Cu is the exchange integral between adjacent Cu(II) ions in the metallacrown unit, and Si is the spin operator for the corresponding Cu(II) ion. Equation 4 treats all the intermolecular and intramolecular weak interactions that are not directly addressed in the Hamiltonian described in eq 3. For these Ln(III)−Cu(II) 15-metallacrowns-5, zJ′ would include both Ln(III)−Cu(II) exchange interactions ions within the metallacrown units and between neighboring metallacrowns. For lanthanide ions such as Pr(III) and Nd(III), the spin− orbit coupling is significant and the J quantum number should be used. The ground terms for Pr(III) and Nd(III) (9-fold degenerate 3H4 and 10-fold degenerate 4I9/2) are well separated from the next levels (∼2200 cm−1 from 3H5 for Pr(III) and 1800 cm−1 from 4I11/2 for Nd(III)).1b,47 Since ligand field effects for lanthanide ions are usually quite small, a free ion approximation can be used in the majority of cases for the simulation of the magnetic properties of lanthanide-containing complexes. However, the magnetic properties of Pr(III) and Nd(III) complexes are governed by the gradual depopulation of Stark sublevels as the temperature decreases (these mJ sublevels arise from crystal field splitting of the ground level). The magnetic properties for complexes 1, 2, and {[NdCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (A) were

Figure 10. χMT vs T plots for complexes (A) {[SmCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (complex B in Tables 3 and 4), (B) 3, and (C) 4, and (D) {[EuCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (complex C in Tables 3 and 4). Data points (○) represent experimental data, line represents the best fit with parameters from Table 4.

and {[LnCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (Ln = Nd, Sm, Eu) are close to or lower than the calculated values for five non-interacting Cu2+ (S = 1/2) ions with gCu = 2.0 and one Ln(III) ion (see Table 3).1b,43−45 With the temperature lowered to 3 K, χM T values for all the studied complexes fell (Table 3), indicating that antiferromagnetic interactions dominate in these complexes. In addition, the decrease of 13159

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165

Article

Inorganic Chemistry Table 3. Experimental (at RTa and LTb K) and Calculated (at RT) Values of χMT for Complexes 1−4 and {[LnCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (Ln = Nd (A), Sm (B), Eu (C))c Calculated χT at RT

Experimental χT at LT 1 2 A 3 B 4 C

PrCu5 NdCu5 NdCu5 SmCu5 SmCu5 EuCu5 EuCu5

0.43 0.78 0.61 0.31 0.24 0.41 0.39

(3 (3 (2 (3 (2 (2 (3

K) K) K) K) K) K) K)

χT at RT

LnIII

Cu5

LnCu5

2.31 2.66 3.68 2.51 2.40 3.05 2.56

1.60

1.875

3.475

1.64

1.875

3.515

0.18−0.42d

1.875

2.055−2.295

1.23−2.00e

1.875

3.105−3.875

a RT = 298 K. bLT = 2 or 3 K. cA = {[NdCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O; B = {[SmCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O; C = {[SmCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O; D = {[(GdCu5(GlyHA)5(H2O)2) (GdCu5(GlyHA)5(H2O)3)(m-bdc)3]·16H2O}n.25 dData taken from refs 46 and 47. eData taken from refs 47 and 48.

thus fitted with the additive model described by eq 2, taking into account axial splitting in the crystal field for the Pr(III) and Nd(III) ions: 2 ̂ ĤLn = ΔJẑ + μB gJ J H

(described by eqs 6−8) for Pr(III) ions48 allowed us to perform a satisfactory fit of the χMT vs T curve. χ (Pr) =

(5)

where Δ is the axial splitting parameter and J ̂ the total angular momentum operator for the lanthanide(III) ion. The previously obtained expressions for magnetic susceptibility

χ⊥ (Pr) =

χ (Pr) =

x=

(6)

x=

Δ kT

(7)

6H2O (A). The parameters obtained (see Table 4) are in good agreement with values obtained previously with the same approach,49,50 but are not comparable to data obtained from analytical expressions for Nd(III) magnetic susceptibility.51a,54 In contrast to the other lanthanides, Eu(III) and Sm(III) ions have low-lying excited states (7F1 and 6H7/2), which are rather closer to the ground states (ca. 350 cm−1 from 7F0 and ca. 700 cm−1 from 6H5/2, for Eu(III) and Sm(III) ions, respectively). Thus, the thermal population of low-lying excited states of Eu(III) and Sm(III) ions could be relatively high at room temperature, and therefore, could be able to noticeably influence the magnetic properties of these ions.1b,47,48,52,55 Simulations of the magnetic properties of Eu(III)-containing 15-metallacrowns-5 fragments were carried out using the additive model described by eq 2 and the equation for molar magnetic susceptibility of Eu(III) ions deduced with a free ion approximation (eq 9), which takes into account the population of both the nonmagnetic ground state 7F0 (split by spin−orbit interaction into seven Stark sublevels 7FJ; J = 0, 1, 2, 3, 4, 5, and 6) and the first low-lying excited state 7F1:47,48,52,55,56

(8)

The values obtained for Δ and gPr for 1 (Table 4) correspond well with those reported previously.48−50 However, in many prior cases, the χ⊥ term was neglected and the fitting procedure for Pr(III)-containing complexes was performed only with eq 6,51,52 so that the cited Δ values are comparable to zJ′, but the assumed condition |zJ′| ≪ |Δ| was actually not implemented.53 An attempt to fit χ M T vs T curves for 2 and {[NdCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (A), using the additive model described by eq 2 and a previously reported expression for magnetic susceptibility for Nd(III) ions derived from the Hamiltonian presented in eq 5 was not successful.51a,54 The energy levels of the Nd(III) ions for the Hamiltonian presented in eq 5 were obtained by full matrix diagonalization, which, in combination with the additive model described by eq 2, allowed us to perform satisfactory fits of χMT vs T data for 2 and {[NdCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)· ⎡ Nβ 2 ⎢ 24 + χEu = 3kTx ⎢ ⎣ λ x= kT

Δ kT

NgPr2 β 2 ⎡ 10(1 − e−x) + 3(e−x − e−4x) + 7/5(e−4x − e−9x) + 4/7(e−9x − e−16x) ⎤ ⎢ ⎥, Δ ⎣ ⎦ 3 + 4e−7x /2 + 5e−3x + 7e−8x + 6e−27x /2 + 7e−20x + 8e−55x /2

χ (Pr) + 2χ⊥ (Pr) 3

Ng 2β 2 ⎡ 2e−x + 8e−4x + 18e−9x + 32e−16x ⎤ ⎢ ⎥, kT ⎣ 1 + 2e−x + 2e−4x + 2e−9x + 2e−16x ⎦

( 272 x − 23 )e−x + ( 1352 x − 25 )e−3x + (189x − 72 )e−6x + (405x − 29 )e−10x + ( 14852 x − 112 )e−15x + ( 24572 x − 132 )e−21x ⎤⎥, 1 + 3e−x + 5e−3x + 7e−6x + 9e−10x + 11e−15x + 13e−21x

⎥ ⎦

(9)

where λ is the spin−orbit coupling parameter. Its values for Eu(III)-based 15-metallacrown-5 complexes (385(29) cm−1 and 352(7) cm−1 for the initial block and for complex 4, respectively; see Table 4) are in good agreement with

previously reported values obtained both from magnetic and spectroscopic studies for various Eu(III) complexes.47,48,52,55−57 Similarly, the free ion approximation equation for Sm(III) magnetic susceptibility (eq 10) takes into account the 13160

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165

Article

Inorganic Chemistry Table 4. Magnetic Properties of Ln(III)-Cu(II) 15-metallacrowns-5 (Ln = Pr, Nd, Sm, Eu, and Gd)a PrCu5 NdCu5 NdCu5 SmCu5 SmCu5 EuCu5 EuCu5 GdCu525

1 2 A 3 B 4 C D

JCu−Cu, cm−1

gCu

gLn

Δ, cm−1

−47(1) −50.2(2) −49.4(2) −49.4(5) −50.6(5) −63.0(8) −49(2) −68(4)

2.00(1) 2.0 (fixed) 2.0 (fixed) 2.23(2) 2.09(2) 2.31(1) 2.02(2) 2.1 (fixed)

0.80(1) 0.68(1) 0.69(2) 0.29 (fixed) 0.29 (fixed)

32(2) 122(1) 108(3)

λ, cm−1

220 (fixed) 220 (fixed) 352(7) 385(29)

2.036(3)

zJ′, cm−1 −0.3(1) −0.40(2) −0.59(1) −2.1(2) −1.8(1) −1.48(7) −0.4(1) 0.030(2)

TIP

3.1(1) × 10−3 1.4(1) × 10−3 1.65(7) × 10−3 5.6(6) × 10−4 1.3(1) × 10−3

R258 1.5 1.65 2.9 1.3 2.34 1.26 5.69 2.13

× × × × × × × ×

10−4 10−4 10−4 10−4 10−5 10−5 10−5 10−5

a

A = {[NdCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O; B = {[SmCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O; C = {[SmCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O; and D = {[(GdCu5(GlyHA)5(H2O)2)(GdCu5(GlyHA)5(H2O)3)(m-bdc)3]·16H2O}n.25

population of the 6H ground term, which is split by spin−orbit interaction into six states 6HJ (J = 5/2, 7/2, 9/2, 11/2, 13/2, 15/2), and that of the first excited state 6H7/2.47,48,52,55,56 χSm =

Nβ 2 × 3kTx

⎡ 2.143x + 7.347 + (42.92x + 1.641)e−7x /2 + (283.7x − 0.6571)e−8x + (620.6x − 1.94)e−27x /2 + (1122x − 2.835)e−20x + (1813x − 3.556)e−55x /2 ⎤ ⎢ ⎥, ⎣ ⎦ 3 + 4e−7x /2 + 5e−3x + 7e−8x + 6e−27x /2 + 7e−20x + 8e−55x /2 x=

λ kT

(10)

During data fitting for {[SmCu 5 (GlyHA) 5 (SO 4 )(H2O)6.5]}2(SO4)·6H2O (B) and 3 with the additive model described by eq 2 and eq 10, the Landé g-factor and spin−orbit coupling parameter for Sm(III) were fixed using typical values previously reported in the literature (0.29 and 220 cm−1, respectively; see Table 4).47,48,52,55−57 The additive model allowed us to obtain satisfactory fits of the χMT vs T data for 1−4 and the initial complexes {[LnCu5(GlyHA)5(SO4)(H2O)6.5]}2(SO4)·6H2O (Ln = Nd (A), Sm (B), Eu (C)). These models for fitting magnetic data for Pr-, Nd-, Sm-, and Eu-containing 15-metallacrowns-5 are proposed here for the first time. The fitted values for exchange parameters between neighboring Cu(II) ions in the 15-metallacrown-5 block range from −49 cm−1 to −63 cm−1, in good correspondence with previously obtained values for JCu−Cu in Gd(III)−Cu(II) 15-metallacrown-5.25 The exchange interaction between adjacent Cu(II) ions in 15-metallacrown-5 systems and between adjacent peripheral Cu(II) ions in 12metallacrown-4 complexes are provided through hydroxamate N−O bridges. The JCu−Cu between adjacent peripheral Cu(II) ions in the majority of Cu(II) crowns fall in the range between −60 cm−1 and −80 cm−1.18−21,27a However, there are no immediately apparent correlations between the J-values and structural parameters for the 15-metallacrown-5 or 12-metallacrown-4 systems.

involved. The 1D coordination polymer with Eu(III) displayed the sorption of alcohols. Models for the simulation of magnetic properties of Pr(III)-, Nd(III)-, Sm(III)-, and Eu(III)-containing Ln(III)-Cu(II) 15metallacrown-5 complexes were proposed for the first time. An additive model, which takes into account exchange interactions between Cu(II) and Ln(III) ions in the metallacrown unit via only a molecular field model allowed us to obtain satisfactory fits of experimental data for the systems containing discrete 15metallacrown-5 units, as well as for phthalate-bridged oligomeric or polymeric systems based thereon. The observed exchange interaction between neighboring Cu(II) ions is antiferromagnetic and is weaker, in comparison to pentacopper(II) 12-metallacrown-4 systems. The best-fit values for the Δ, λ, and g-factors for Ln(III) ions are comparable to those of other lanthanide complexes.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.7b01944. Additional figures illustrating molecular and crystal structures of compounds 1−4 (Figure S1−S13), data on X-ray crystal structure refinement (Table S1), additional details regarding the synthesis of compounds 1−4 (Table S2), tables with selected bond length and angles (Tables S3−S6) (PDF)



CONCLUSIONS Metatheses of sulfate anions in the isomorphous Ln(III)-Cu(II) 15-metallacrown-5 complexes (Ln(III) = Pr, Nd, Sm, Eu) with m-phthalate resulted in the formation of several different products, despite the same reaction conditions. Two 1Dcoordination polymers were isolated for Eu(III)- and Nd(III)containing 15-metallacrown-5, whereas, in the cases of Pr(III)and Sm(III)-based initial 15-metallacrowns-5, isostructural discrete complexes containing four metallamacrocycle units were formed. The nature of the oligomerization induced by mphthalate was dependent upon the particular Ln(III) ion

Accession Codes

CCDC Nos. 1552713 and 1552714, and 891994 and 891995, contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc. cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, U.K.; fax: +44 1223 336033. 13161

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165

Article

Inorganic Chemistry



Langley, S. K.; Murray, K. S.; Rajaraman, G. Exploring the Influence of Diamagnetic Ions on the Mechanism of Magnetization Relaxation in {CoIII2LnIII2} (Ln = Dy, Tb, Ho) “Butterfly” Complexes. Inorg. Chem. 2017, 56, 2518−2532. (5) (a) Karotsis, G.; Kennedy, S.; Teat, S. J.; Beavers, C. M.; Fowler, D. A.; Morales, J. J.; Evangelisti, M.; Dalgarno, S. J.; Brechin, E. K. [MnIII4LnIII4] Calix[4]arene Clusters as Enhanced Magnetic Coolers and Molecular Magnets. J. Am. Chem. Soc. 2010, 132, 12983−12990. (b) Moreno Pineda, E.; Heesing, C.; Tuna, F.; Zheng, Y.-Z.; McInnes, E. J. L.; Schnack, J.; Winpenny, R. E. P. Copper Lanthanide Phosphonate Cages: Highly Symmetric {Cu3Ln9P6} and {Cu6Ln6P6} Clusters with C3v and D3h Symmetry. Inorg. Chem. 2015, 54, 6331− 6337. (c) Rinck, J.; Lan, Y.; Anson, C. E.; Powell, A. K. Coordination Cluster Nuclearity Decreases with Decreasing Rare Earth Ionic Radius in 1:1 Cr/Ln N-Butyldiethanolamine Compounds: A Journey across the Lanthanide Series from Cr4IIILa4−Cr4IIITb4 via Cr3IIIDy3 and Cr3IIIHo3 to Cr2IIIEr2−Cr2IIILu2. Inorg. Chem. 2015, 54, 3107−3117. (d) Pineda, E. M.; Tuna, F.; Zheng, Y.-Z.; Teat, S. J.; Winpenny, R. E. P.; Schnack, J.; McInnes, E. J. L. Iron Lanthanide Phosphonate Clusters: {Fe6Ln6P6} Wells-Dawson-like Structures with D3d Symmetry. Inorg. Chem. 2014, 53, 3032−3038. (e) Xue, S.; Guo, Y.-N.; Zhao, L.; Zhang, H.; Tang, J. Molecular Magnetic Investigation of a Family of Octanuclear [Cu6Ln2] Nanoclusters. Inorg. Chem. 2014, 53, 8165− 8171. (6) (a) Benelli, C.; Blake, A. J.; Milne, P. E. Y.; Rawson, J. M.; Winpenny, R. E. P. Magnetic and Structural Studies of Copper− Lanthanoid Complexes; the Synthesis and Structures of New Cu3Ln Complexes of 6-chloro-2-Pyridone (Ln = Gd, Dy and Er) and Magnetic Studies on Cu2Gd2, Cu4Gd2 and Cu3Gd Complexes. Chem. Eur. J. 1995, 1, 614−618. (b) Kajiwara, T.; Nakano, M.; Takahashi, K.; Takaishi, S.; Yamashita, M. Structural Design of Easy-Axis Magnetic Anisotropy and Determination of Anisotropic Parameters of LnIII− CuII Single-Molecule Magnets. Chem. - Eur. J. 2011, 17, 196−205. (c) Mishra, A.; Wernsdorfer, W.; Abboud, K. A.; Christou, G. Initial Observation of Magnetization Hysteresis and Quantum Tunneling in Mixed Manganese−Lanthanide Single-Molecule Magnets. J. Am. Chem. Soc. 2004, 126, 15648−15649. (d) Osa, S.; Kido, T.; Matsumoto, N.; Re, N.; Pochaba, A.; Mrozinski, J. A. Tetranuclear 3d−4f Single Molecule Magnet: [CuIILTbIII(hfac)2]2. J. Am. Chem. Soc. 2004, 126, 420−421. (e) Harriman, K. L. M.; Brosmer, J. L.; Ungur, L.; Diaconescu, P. L.; Murugesu, M. Pursuit of Record Breaking Energy Barriers: A Study of Magnetic Axiality in Diamide Ligated DyIII Single-Molecule Magnets. J. Am. Chem. Soc. 2017, 139, 1420−1423. (7) (a) Mikami, K.; Terada, M.; Matsuzawa, H. Asymmetric Catalysis by Lanthanide Complexes. Angew. Chem., Int. Ed. 2002, 41, 3554− 3572. (b) Li, Y.; Deng, P.; Zeng, Y.; Xiong, Y.; Zhou, H. anti-Selective Asymmetric Henry Reaction Catalyzed by a Heterobimetallic Cu− Sm−Aminophenol Sulfonamide Complex. Org. Lett. 2016, 18, 1578− 1581. (8) (a) Aboshyan-Sorgho, L.; Nozary, H.; Aebischer, A.; Bünzli, J.-C. G.; Morgantini, P.-Y.; Kittilstved, K. R.; Hauser, A.; Eliseeva, S. V.; Petoud, S.; Piguet, C. Optimizing Millisecond Time Scale NearInfrared Emission in Polynuclear Chrome(III)−Lanthanide(III) Complexes. J. Am. Chem. Soc. 2012, 134, 12675−12684. (b) Zhao, B.; Chen, X.-Y.; Cheng, P.; Liao, D.-Z.; Yan, S.-P.; Jiang, Z.-H. Coordination Polymers Containing 1D Channels as Selective Luminescent Probes. J. Am. Chem. Soc. 2004, 126, 15394−15395. (c) Oyarzabal, I.; Artetxe, B.; Rodríguez-Diéguez, A.; García, J. Á .; Seco, J. M.; Colacio, E. A family of acetato-diphenoxo triply bridged dimetallic ZnIILnIII complexes: SMM behavior and luminescent properties. Dalton Trans. 2016, 45, 9712−9726. (9) (a) Wong, W.-K.; Yang, X.; Jones, R. A.; Rivers, J. H.; Lynch, V.; Lo, W.-K.; Xiao, D.; Oye, M. M.; Holmes, A. L. Multinuclear Luminescent Schiff-Base Zn−Nd Sandwich Complexes. Inorg. Chem. 2006, 45, 4340−4345. (b) Palacios, M. A.; Titos-Padilla, S.; Ruiz, J.; Herrera, J. M.; Pope, S. J. A.; Brechin, E. K.; Colacio, E. Bifunctional ZnIILnIII Dinuclear Complexes Combining Field Induced SMM Behavior and Luminescence: Enhanced NIR Lanthanide Emission by 9-Anthracene Carboxylate Bridging Ligands. Inorg. Chem. 2014, 53,

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (A. V. Pavlishchuk). *E-mail: [email protected] (A. W. Addison). ORCID

Anthony W. Addison: 0000-0002-0706-7377 Funding

A.V.P. thanks the European Community’s Seventh Framework Programme (FP7/2007−2013, Grant Agreement No. 611488) and the Fulbright Foundation for a Fellowship. A.W.A. thanks Drexel University for support under the Collaborative Research Agreement between the Drexel University College of Arts & Sciences and the Pisarzhevskii Institute of Physical Chemistry. S.V.K. thanks the National Academy of Sciences of Ukraine for support. The X-ray diffractometer was funded by NSF Grant 0087210, Ohio Board of Regents Grant CAP-491, and by Youngstown State University. We thank Prof. I. O. Fritsky for helpful discussion of the structural descriptions for 1 and 3. Notes

The authors declare no competing financial interest.



REFERENCES

(1) (a) Woodruff, D. N.; Winpenny, R. E. P.; Layfield, R. A. Lanthanide Single-Molecule Magnets. Chem. Rev. 2013, 113, 5110− 5148. (b) Benelli, C.; Gatteschi, D. Magnetism of Lanthanides in Molecular Materials with Transition-Metal Ions and Organic Radicals. Chem. Rev. 2002, 102, 2369−2387. (2) (a) Sessoli, R.; Powell, A. K. Strategies towards single molecule magnets based on lanthanide ions. Coord. Chem. Rev. 2009, 253, 2328−2341. (b) Feltham, H. L. C.; Brooker, S. Review of purely 4f and mixed-metal nd-4f single-molecule magnets containing only one lanthanide ion. Coord. Chem. Rev. 2014, 276, 1−33. (c) Liu, J.-L.; Chen, Y.-C.; Guo, F.-S.; Tong, M.-L. Recent advances in the design of magnetic molecules for use as cryogenic magnetic coolants. Coord. Chem. Rev. 2014, 281, 26−49. (3) (a) Dinca, A. S.; Ghirri, A.; Madalan, A. M.; Affronte, M.; Andruh, M. Dodecanuclear [CuII6GdIII6] Nanoclusters as Magnetic Refrigerants. Inorg. Chem. 2012, 51, 3935−3937. (b) Hooper, T. N.; Inglis, R.; Lorusso, G.; Ujma, J.; Barran, P. E.; Uhrin, D.; Schnack, J.; Piligkos, S.; Evangelisti, M.; Brechin, E. K. Structurally Flexible and Solution Stable [Ln4TM8(OH)8(L)8(O2CR)8(MeOH)y](ClO4)4: A Playground for Magnetic Refrigeration. Inorg. Chem. 2016, 55, 10535− 10546. (c) Canaj, A. B.; Kalofolias, D. A.; Siczek, M.; Lis, T.; McNab, R.; Lorusso, G.; Inglis, R.; Evangelisti, M.; Milios, C. J. Tetradecanuclearity in 3d−4f chemistry: relaxation and magnetocaloric effects in [NiII6LnIII8] species. Dalton Trans. 2017, 46, 3449−3452. (d) Ruiz, J.; Lorusso, G.; Evangelisti, M.; Brechin, E. K.; Pope, S. J. A.; Colacio, E. Closely-Related ZnII2LnIII2 Complexes (LnIII = Gd, Yb) with Either Magnetic Refrigerant or Luminescent Single-Molecule Magnet Properties. Inorg. Chem. 2014, 53, 3586−3594. (4) (a) Moreno Pineda, E.; Chilton, N. F.; Tuna, F.; Winpenny, R. E. P.; McInnes, E. J. L. Systematic Study of a Family of Butterfly-Like {M2Ln2} Molecular Magnets (M = MgII, MnIII, CoII, NiII, and CuII; Ln = YIII, GdIII, TbIII, DyIII, HoIII, and ErIII). Inorg. Chem. 2015, 54, 5930− 5941. (b) Langley, S. K.; Wielechowski, D. P.; Chilton, N. F.; Moubaraki, B.; Murray, K. S. A Family of {CrIII2LnIII2} Butterfly Complexes: Effect of the Lanthanide Ion on the Single-Molecule Magnet Properties. Inorg. Chem. 2015, 54, 10497−10503. (c) Langley, S. K.; Ungur, L.; Chilton, N. F.; Moubaraki, B.; Chibotaru, L. F.; Murray, K. S. Single-Molecule Magnetism in a Family of {CoIII2DyIII2} Butterfly Complexes: Effects of Ligand Replacement on the Dynamics of Magnetic Relaxation. Inorg. Chem. 2014, 53, 4303−4315. (d) Langley, S. K.; Le, C.; Ungur, L.; Moubaraki, B.; Abrahams, B. F.; Chibotaru, L. F.; Murray, K. S. Heterometallic 3d−4f SingleMolecule Magnets: Ligand and Metal Ion Influences on the Magnetic Relaxation. Inorg. Chem. 2015, 54, 3631−3642. (e) Vignesh, K. R.; 13162

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165

Article

Inorganic Chemistry

[CuIILnIIIMV] heterotrimetallic complexes (Ln = La, Gd, Tb; M = Mo, W): model systems to probe exchange interactions and singlemolecule magnet properties. Dalton Trans. 2016, 45, 7642−7649. (18) Pavlishchuk, A. V.; Satska, Yu. A.; Kolotilov, S. V.; Fritsky, I. O. Coordination Polymers and Oligonuclear Systems Based on Oximate or Hydroxamate Building Blocks: Magnetic and Sorption Properties. Curr. Inorg. Chem. 2015, 5, 5−25. (19) Pavlishchuk, A. V.; Kolotilov, S. V.; Zeller, M.; Thompson, L. K.; Fritsky, I. O.; Addison, A. W.; Hunter, A. D. A Triple-Decker Heptadecanuclear (CuII)15(CrIII)2 Complex Assembled from Pentanuclear Metallacrowns. Eur. J. Inorg. Chem. 2010, 4851−4858. (20) Pavlishchuk, A. V.; Kolotilov, S. V.; Zeller, M.; Shvets, O. V.; Fritsky, I. O.; Lofland, S. E.; Addison, A. W.; Hunter, A. D. Magnetic and Sorption Properties of Supramolecular Systems Based on Pentanuclear Copper(II) 12-Metallacrown-4 Complexes and Isomeric Phthalates: Structural Modeling of the Different Stages of Alcohol Sorption. Eur. J. Inorg. Chem. 2011, 4826−4836. (21) Pavlishchuk, A. V.; Kolotilov, S. V.; Zeller, M.; Lofland, S. E.; Kiskin, M. A.; Efimov, N. N.; Ugolkova, E. A.; Minin, V. V.; Novotortsev, V. M.; Addison, A. W. Supramolecular Maleate Adducts of Copper(II) 12-Metallacrown-4: Magnetism, EPR and Alcohol Sorption Properties. Eur. J. Inorg. Chem. 2017, in press (DOI: 10.1002/ejic.201700976). (22) Atzeri, C.; Marchiò, L.; Chow, C. Y.; Kampf, J. W.; Pecoraro, V. L.; Tegoni, M. Design of 2D Porous Coordination Polymers Based on Metallacrown Units. Chem.Eur. J. 2016, 22, 6482−6486. (23) (a) Cutland-Van Noord, A. D.; Kampf, J. W.; Pecoraro, V. L. Preparation of Resolved Fourfold Symmetric Amphiphilic Helices Using Chiral Metallacrown Building Blocks. Angew. Chem., Int. Ed. 2002, 41, 4667−4670. (b) Cutland, A. D.; Halfen, J. A.; Kampf, J. W.; Pecoraro, V. L. Chiral 15-Metallacrown-5 Complexes Differentially Bind Carboxylate Anions. J. Am. Chem. Soc. 2001, 123, 6211−6212. (24) Lim, C.-S.; Jankolovits, J.; Kampf, J. W.; Pecoraro, V. L. Chiral Metallacrown Supramolecular Compartments that Template Nanochannels: Self-Assembly and Guest Absorption. Chem.Asian J. 2010, 5, 46−49. (25) Pavlishchuk, A. V.; Kolotilov, S. V.; Zeller, M.; Thompson, L. K.; Addison, A. W. Formation of Coordination Polymers or Discrete Adducts via Reactions of Gadolinium(III)−Copper(II) 15-Metallacrown-5 Complexes with Polycarboxylates: Synthesis, Structures and Magnetic Properties. Inorg. Chem. 2014, 53, 1320−1330. (26) (a) Li, Q.-W.; Wan, R.-C.; Chen, Y.-C.; Liu, J.-L.; Wang, L.-F.; Jia, J.-H.; Chilton, N. F.; Tong, M.-L. Unprecedented hexagonal bipyramidal single-ion magnets based on metallacrowns. Chem. Commun. 2016, 52, 13365−13368. (b) Boron, T. T., III; Lutter, J. C.; Daly, C. I.; Chow, C. Y.; Davis, A. H.; Nimthong-Roldan, A.; Zeller, M.; Kampf, J. W.; Zaleski, C. M.; Pecoraro, V. L. The Nature of the Bridging Anion Controls the Single-Molecule Magnetic Properties of DyX4M 12-Metallacrown-4 Complexes. Inorg. Chem. 2016, 55, 10597−10607. (c) Boron, T. T., III; Kampf, J. W.; Pecoraro, V. L. A Mixed 3d−4f 14-Metallacrown-5 Complex That Displays Slow Magnetic Relaxation through Geometric Control of Magnetoanisotropy. Inorg. Chem. 2010, 49, 9104−9106. (d) Cao, F.; Wang, S.; Li, D.; Zeng, S.; Niu, M.; Song, Y.; Dou, J. Family of Mixed 3d−4f Dimeric 14-Metallacrown-5 Compounds: Syntheses, Structures, and Magnetic Properties. Inorg. Chem. 2013, 52, 10747−10755. (27) (a) Ostrowska, M.; Fritsky, I. O.; Gumienna-Kontecka, E.; Pavlishchuk, A. V. Metallacrown-based compounds: Applications in catalysis, luminescence, molecular magnetism, and adsorption. Coord. Chem. Rev. 2016, 327−328, 304−332. (b) Li, Q.-W.; Liu, J.-L.; Jia, J.H.; Chen, Y.-C.; Liu, J.; Wang, L.-F.; Tong, M. L. “Half-Sandwich” YbIII Single-Ion Magnets with Metallacrowns. Chem. Commun. 2015, 51, 10291−10294. (28) (a) Jankolovits, J.; Andolina, C. M.; Kampf, J. W.; Raymond, K. N.; Pecoraro, V. L. Assembly of Near-Infrared Luminescent Lanthanide Host (Host−Guest) Complexes With a Metallacrown Sandwich Motif. Angew. Chem., Int. Ed. 2011, 50, 9660−9664. (b) Trivedi, E. R.; Eliseeva, S. V.; Jankolovits, J.; Olmstead, M. M.; Petoud, S.; Pecoraro, V. L. Highly Emitting Near-Infrared Lanthanide

1465−1474. (c) Zhang, L.; Zhao, L.; Zhang, P.; Wang, C.; Yuan, S.W.; Tang, J. Nanoscale {LnIII24ZnII6} Triangular Metalloring with Magnetic Refrigerant, Slow Magnetic Relaxation, and Fluorescent Properties. Inorg. Chem. 2015, 54, 11535−11541. (d) Maity, M.; Majee, M. C.; Kundu, S.; Samanta, S. K.; Sañudo, E. C.; Ghosh, S.; Chaudhury, M. Pentanuclear 3d−4f Heterometal Complexes of MII3LnIII2 (M = Ni, Cu, Zn and Ln = Nd, Gd, Tb) Combinations: Syntheses, Structures, Magnetism, and Photoluminescence Properties. Inorg. Chem. 2015, 54, 9715−9726. (e) Griffiths, K.; Mayans, J.; Shipman, M. A.; Tizzard, G. J.; Coles, S. J.; Blight, B. A.; Escuer, A.; Kostakis, G. E. Four New Families of Polynuclear Zn-Ln Coordination Clusters. Synthetic, Topological, Magnetic, and Luminescent Aspects. Cryst. Growth Des. 2017, 17, 1524−1538. (10) Stemmler, A. J.; Kampf, J. W.; Pecoraro, V. L. A Planar [15]Metallacrown-5 That Selectively Binds the Uranyl Cation. Angew. Chem., Int. Ed. Engl. 1996, 35, 2841−2843. (11) Cutland, A. D.; Malkani, R. G.; Kampf, J. W.; Pecoraro, V. L. Lanthanide [15]Metallacrown-5 Complexes Form Nitrate-Selective Chiral Cavities. Angew. Chem., Int. Ed. 2000, 39, 2689−2691. (12) (a) Zhao, B.; Chen, X.-Y.; Chen, Z.; Shi, W.; Cheng, P.; Yan, S.P.; Liao, D.-Z. A porous 3D heterometal-organic framework containing both lanthanide and high-spin Fe(II) ions. Chem. Commun. 2009, 3113−3115. (b) Wang, Y.; Wang, X.-G.; Yuan, B.; Shao, C.-Y.; Chen, Y.-Y.; Zhou, B.-B.; Li, M.-S.; An, X.-M.; Cheng, P.; Zhao, X.-J. CationExchange Porosity Tuning in a Dynamic 4d−4f−3d Framework for NiII Ion-Selective Luminescent Probe. Inorg. Chem. 2015, 54, 4456− 4465. (13) Mahata, P.; Ramya, K. V.; Natarajan, S. Reversible Water Intercalation Accompanied by Coordination and Color Changes in a Layered Metal−Organic Framework. Inorg. Chem. 2009, 48, 4942− 4951. (14) Feng, X.; Feng, Y.; Guo, N.; Sun, Y.; Zhang, T.; Ma, L.; Wang, L. Series d-f Heteronuclear Metal-Organic Frameworks: Color Tunability and Luminescent Probe with Switchable Properties. Inorg. Chem. 2017, 56, 1713−1721. (15) Sapianik, A. A.; Zorina-Tikhonova, E. N.; Kiskin, M. A.; Samsonenko, D. G.; Kovalenko, K. A.; Sidorov, A. A.; Eremenko, I. L.; Dybtsev, D. N.; Blake, A. J.; Argent, S. P.; Schröder, M.; Fedin, V. P. Rational Synthesis and Investigation of Porous Metal−Organic Framework Materials from a Preorganized Heterometallic Carboxylate Building Block. Inorg. Chem. 2017, 56, 1599−1608. (16) (a) Dhers, S.; Feltham, H. L. C.; Clérac, R.; Brooker, S. Design of One-Dimensional Coordination Networks from a Macrocyclic {3d4f} Single-Molecule Magnet Precursor Linked by [W(CN)8]3− Anions. Inorg. Chem. 2013, 52, 13685−13691. (b) Dhers, S.; Feltham, H. L. C.; Rouzières, M.; Clérac, R.; Brooker, S. Macrocyclic {3d−4f} SMMs as building blocks for 1D-polymers: selective bridging of 4f ions by use of an O-donor ligand. Dalton Trans. 2016, 45, 18089−18093. (c) Dhers, S.; Feltham, H. L.C.; Brooker, S. A toolbox of building blocks, linkers and crystallisation methods used to generate single-chain magnets. Coord. Chem. Rev. 2015, 296, 24−44. (d) Kolotilov, S. V.; Kiskin, M. A.; Eremenko, I. L.; Novotortsev, V. M. The Influence of Diamagnetic Substrates Absorption on Magnetic Properties of Porous Coordination Polymers. Curr. Inorg. Chem. 2013, 3, 144−160. (17) (a) Vieru, V.; Pasatoiu, T. D.; Ungur, L.; Suturina, E.; Madalan, A. M.; Duhayon, C.; Sutter, J.-P.; Andruh, M.; Chibotaru, L. F. Synthesis, Crystal Structures, Magnetic Properties, and Theoretical Investigation of a New Series of NiII−LnIII−WV Heterotrimetallics: Understanding the SMM Behavior of Mixed Polynuclear Complexes. Inorg. Chem. 2016, 55, 12158−12171. (b) Hu, K.-Q.; Jiang, X.; Wu, S.Q.; Liu, C.-M.; Cui, A.-L.; Kou, H.-Z. Slow Magnetization Relaxation in NiIIDyIIIFeIII Molecular Cycles. Inorg. Chem. 2015, 54, 1206−1208. (c) Dhers, S.; Costes, J.-P.; Guionneau, P.; Paulsen, C.; Vendier, L.; Sutter, J.-P. On the importance of ferromagnetic exchange between transition metals in field-free SMMs: examples of ring-shaped heterotrimetallic [(LnNi2){W(CN)8}]2 compounds. Chem. Commun. 2015, 51, 7875−7878. (d) Visinescu, D.; Alexandru, M.-G.; Madalan, A. M.; Jeon, I.-R.; Mathonière, C.; Clérac, R.; Andruh, M. A new family of 13163

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165

Article

Inorganic Chemistry “Encapsulated Sandwich” Metallacrown Complexes with Excitation Shifted Toward Lower Energy. J. Am. Chem. Soc. 2014, 136, 1526− 1534. (29) Mezei, G.; Kampf, J. W.; Pan, S.; Poeppelmeier, K. R.; Watkins, B.; Pecoraro, V. L. Metallacrown-based compartments: selective encapsulation of three isonicotinate anions in non-centrosymmetric solids. Chem. Commun. 2007, 1148−1150. (30) Pavlishchuk, A. V.; Kolotilov, S. V.; Fritsky, I. O.; Zeller, M.; Addison, A. W.; Hunter, A. D. Structural trends in a series of isostructural lanthanide-copper metallacrown sulfates (LnIII = Pr, Nd, Sm, Eu, Gd, Dy and Ho): hexaaquapentakis[μ3-glycine hydroxamate(2-)] sulfatopentacopper(II) lanthanide(III) heptaaquapentakis[μ3glycine hydroxamato(2-)]sulfatopentacopper(II) lanthanide(III) sulfate hexahydrate. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 2011, 67, m255−m265. (31) Smissman, E. E.; Warner, V. D. Specificity in enzyme inhibition. 2. alpha.-Aminohydroxamic acids as inhibitors of histidine decarboxylase and 3,4-dihydroxyphenylalanine decarboxylase. J. Med. Chem. 1972, 15, 681−682. (32) Kahn, O. In Molecular Magnetism; VCH Publishers, Inc.: Weinheim, Germany, 1993; Chapter 2, p 26. (33) Youngme, S.; Chaichit, N.; Kongsaeree, P.; van Albada, G. A.; Reedijk, J. Synthesis, structure, spectroscopy, and magnetism of two new dinuclear carbonato-bridged Cu(II) complexes. Inorg. Chim. Acta 2001, 324, 232−240. (34) Majumder, A.; Choudhury, C. R.; Mitra, S.; Rosair, G. M.; El Fallah, M. S.; Ribas, J. Isolation of a new two-dimensional honeycomb carbonato-bridged copper(II) complex exhibiting long-range ferromagnetic ordering. Chem. Commun. 2005, 2158−2160. (35) Vercouter, T.; Amekraz, B.; Moulin, C.; Giffaut, E.; Vitorge, P. Sulfate Complexation of Trivalent Lanthanides Probed by Nanoelectrospray Mass Spectrometry and Time-Resolved Laser-Induced Luminescence. Inorg. Chem. 2005, 44, 7570−7581. (b) Yun, S. S.; Bae, S. H.; Hong, S. W.; Kang, S. K.; Kim, I. H.; Park, J. T. Thermodynamics of complexation of lanthanides by various dimethoxybenzoates in aqueous solution. Thermochim. Acta 1994, 246, 39−47. (36) (a) Stemmler, A. J.; Kampf, J. W.; Kirk, M. L.; Atasi, B. H.; Pecoraro, V. L. The Preparation, Characterization, and Magnetism of Copper 15-Metallacrown-5 Lanthanide Complexes. Inorg. Chem. 1999, 38, 2807−2817. (b) Zaleski, C. M.; Depperman, E. C.; Kampf, J. W.; Kirk, M. L.; Pecoraro, V. L. Using LnIII[15-MCCuII(N)(S)‑pheHA-5]3+ Complexes To Construct Chiral Single-Molecule Magnets and Chains of Single-Molecule Magnets. Inorg. Chem. 2006, 45, 10022−10024. (37) (a) Jankolovits, J.; Lim, C.-S.; Mezei, G.; Kampf, J. W.; Pecoraro, V. L. Influencing the Size and Anion Selectivity of Dimeric Ln3+[15Metallacrown-5] Compartments through Systematic Variation of the Host Side Chains and Central Metal. Inorg. Chem. 2012, 51, 4527− 4538. (b) Zaleski, C. M.; Lim, C.-S.; Cutland-Van Noord, A. D.; Kampf, J. W.; Pecoraro, V. L. Effects of the Central Lanthanide Ion Crystal Radius on the 15-MCCuII(N)pheHA-5 Structure. Inorg. Chem. 2011, 50, 7707−7717. (38) Addison, A. W.; Rao, T. N.; Reedijk, J.; van Rijn, J.; Verschoor, G. C. Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen−sulphur donor ligands; the crystal and molecular structure of aqua[1,7-bis(N-methylbenzimidazol-2′-yl)2,6-dithiaheptane]copper(II) perchlorate. J. Chem. Soc., Dalton Trans. 1984, 1349−1356. (39) (a) Spek, A. L. PLATON, An Integrated Tool for the Analysis of the Results of a Single Crystal Structure Determination. Acta Crystallogr., Sect. A 1990, 46, C34. (b) Spek, A. L. PLATON−A Multipurpose Crystallographic Tool; Utrecht University: Utrecht, The Netherlands, 2006. (40) (a) Reineke, T. M.; Eddaoudi, M.; O’Keeffe, M.; Yaghi, O. M. A Microporous Lanthanide Organic Framework. Angew. Chem., Int. Ed. 1999, 38, 2590−2594. (b) Kitaura, R.; Seki, K.; Akiyama, G.; Kitagawa, S. Porous Coordination-Polymer Crystals with Gated Channels Specific for Supercritical Gases. Angew. Chem., Int. Ed. 2003, 42, 428−431. (c) Tanaka, D.; Nakagawa, K.; Higuchi, M.; Horike, S.;

Kubota, Y.; Kobayashi, T. C.; Takata, M.; Kitagawa, S. Kinetic GateOpening Process in a Flexible Porous Coordination Polymer. Angew. Chem., Int. Ed. 2008, 47, 3914−3918. (41) Gregg, S. J.; Sing, K. S. W. In Adsorption, Surface Area and Porosity, 2nd Edition; Academic Press: London, 1982; p 4. (42) Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Selective gas adsorption and separation in metal−organic frameworks. Chem. Soc. Rev. 2009, 38, 1477−1504. (43) (a) Zhao, L.; Lin, S.; Shen, S.; Tang, J. A novel two dimensional samarium(III) coordination framework with N-(2-Hydroxyethyl)iminodiacetic acid and oxalate ligands: Synthesis, crystal structure and magnetic property. Inorg. Chem. Commun. 2011, 14, 1928−1931. (b) Fomina, I. G.; Dobrokhotova, Zh. V.; Ilyukhin, A. B.; Aleksandrov, G. G.; Kazak, V. O.; Gehman, A. E.; Efimov, N. N.; Bogomyakov, A. S.; Zavorotny, Y. S.; Gerasimova, V. I.; Novotortsev, V. M.; Eremenko, I. L. Binuclear samarium(III) pivalates with chelating N-donors: synthesis, structure, thermal behavior, magnetic and luminescent properties. Polyhedron 2013, 65, 152−160. (c) Li, W.; Wang, X.-L.; Song, X.-Y.; Li, L.-C.; Liao, D.-Z.; Jiang, Z.-H. Crystal structures, magnetic and photoluminescent properties of one-dimensional lanthanide complexes with picolinate ligand. J. Mol. Struct. 2008, 885, 1−6. (d) Oczko, G.; Legendziewicz, J.; Wickleder, M. S.; Meyer, G. Crystal structure, magnetism and photophysics of the lanthanide maleates RE(C4O4H3)3·8H2O (RE = Nd, Sm). J. Alloys Compd. 2002, 341, 255−262. (e) 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. 2010, 49, 969−976. (44) (a) Wu, H.-C.; Jhu, Z.-R.; Lee, G.-H.; Peng, S.-M.; Yang, C.-I.; Kuei, C.-H. Lanthanide coordination polymers based on Ln2 cluster: Syntheses, crystal structures photoluminescence and magnetic properties. Polyhedron 2013, 66, 34−38. (b) Feng, X.; Liu, B.; Wang, L.-Y.; Zhao, J.-S.; Wang, J. G.; Weng, N. S.; Shi, X.-G. A series of lanthanide−organic polymers incorporating nitrogen-heterocyclic and aliphatic carboxylate mixed-ligands: structures, luminescent and magnetic properties. Dalton Trans. 2010, 39, 8038−8049. (c) Feng, X.; Zhou, L.-L.; Wang, L.-Y.; Zhou, J.-G.; Shi, Z.-Q.; Shang, J.-J. Two unique lanthanide−organic frameworks based on 1H-2-methyl-4,5imidazole-dicarboxylate and oxalate: Crystal structure, luminescence and magnetic properties. Inorg. Chim. Acta 2013, 394, 696−702. (45) (a) Han, Y.; Li, X.; Li, L.; Ma, C.; Shen, Z.; Song, Y.; You, X. Structures and Properties of Porous Coordination Polymers Based on Lanthanide Carboxylate Building Units. Inorg. Chem. 2010, 49, 10781−10787. (b) Yang, J.; Yue, Q.; Li, G.-D.; Cao, J.-J.; Li, G.-H.; Chen, J.-S. Structures, Photoluminescence, Up-Conversion, and Magnetism of 2D and 3D Rare-Earth Coordination Polymers with Multicarboxylate Linkages. Inorg. Chem. 2006, 45, 2857−2865. (c) Lin, Z.-J.; Xu, B.; Liu, T.-F.; Cao, M.-N.; Lü, J.; Cao, R. A Series of Lanthanide Metal−Organic Frameworks Based on Biphenyl-3,4′,5tricarboxylate: Syntheses, Structures, Luminescence and Magnetic Properties. Eur. J. Inorg. Chem. 2010, 3842−3849. (d) Black, C. A.; Costa, J. S.; Fu, W. T.; Massera, C.; Roubeau, O.; Teat, S. J.; Aromí, G.; Gamez, P.; Reedijk, J. 3-D Lanthanide Metal-Organic Frameworks: Structure, Photoluminescence, and Magnetism. Inorg. Chem. 2009, 48, 1062−1068. (e) Ji, B.; Deng, D.; He, X.; Liu, B.; Miao, S.; Ma, N.; Wang, W.; Ji, L.; Liu, P.; Li, X. Syntheses, Structures, Luminescence, and Magnetic Properties of One-dimensional Lanthanide Coordination Polymers with a Rigid 2,2′-Bipyridine-3,3′,6,6′-tetracarboxylic Acid Ligand. Inorg. Chem. 2012, 51, 2170−2177. (f) Zhang, H.; Li, N.; Tian, C.; Liu, T.; Du, F.; Lin, P.; Li, Z.; Du, S. Unusual High Thermal Stability within a Series of Novel Lanthanide TATB Frameworks: Synthesis, Structure, and Properties (TATB = 4,4′,4″-s-Triazine-2,4,6triyl-tribenzoate). Cryst. Growth Des. 2012, 12, 670−678. (46) Rinehart, J. D.; Long, J. R. Exploiting single-ion anisotropy in the design of f-element single-molecule magnets. Chem. Sci. 2011, 2, 2078−2085 (also see ref 10). (47) Andruh, M.; Bakalbassis, E.; Kahn, O.; Trombe, J. C.; Porcher, P. Structure and Spectroscopic and Magnetic Properties of Rare Earth Metal(III) Derivatives with the 2-Formyl-4-methyl-6-(N-(213164

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165

Article

Inorganic Chemistry pyridylethy1)formimidoyl)phenol Ligand. Inorg. Chem. 1993, 32, 1616−1622. (48) Kahwa, I. A.; Selbin, J.; O’Connor, C. J.; Foise, J. W.; McPherson, G. L. Magnetic and Luminescence Characteristics of Dinuclear Complexes of Lanthanides and a Phenolic Schiff Base Macrocyclic Ligand. Inorg. Chim. Acta 1988, 148, 265−272. (49) (a) Mikhalyova, E. A.; Yakovenko, A. V.; Zeller, M.; Gavrilenko, K. S.; Lofland, S. E.; Addison, A. W.; Pavlishchuk, V. V. Structure, magnetic and luminescence properties of the lanthanide complexes Ln2(Salphen)3·H2O (Ln = Pr, Nd, Sm, Eu, Gd, Tb, Dy; H2Salphen = N,N′-bis(salicylidene)-1,2-phenylenediamine). Inorg. Chim. Acta 2014, 414, 97−104. (b) Mikhalyova, E. A.; Kolotilov, S. V.; Zeller, M.; Thompson, L. K.; Addison, A. W.; Pavlishchuk, V. V.; Hunter, A. D. Synthesis, structure and magnetic properties of Nd3+ and Pr3+ 2D polymers with tetrafluoro-p-phthalate. Dalton Trans. 2011, 40, 10989− 10996. (50) Marinho, M. V.; Reis, D. O.; Oliveira, W. X. C.; Marques, L. F.; Stumpf, H. O.; Déniz, M.; Pasán, J.; Ruiz-Pérez, C.; Cano, J.; Lloret, F.; Julve, M. Photoluminescent and Slow Magnetic Relaxation Studies on Lanthanide(III)-2,5-pyrazinedicarboxylate Frameworks. Inorg. Chem. 2017, 56, 2108−2123. (51) (a) Xu, N.; Shi, W.; Liao, D.-Z.; Yan, S.-P.; Cheng, P. Template Synthesis of Lanthanide (Pr, Nd, Gd) Coordination Polymers with 2Hydroxynicotinic Acid Exhibiting Ferro-/Antiferromagnetic Interaction. Inorg. Chem. 2008, 47, 8748−8756. (b) Li, B.; Gu, W.; Zhang, L.Z.; Qu, J.; Ma, Z.-P.; Liu, X.; Liao, D.-Z. [Ln2(C2O4)2(pyzc)2(H2O)2]n [Ln = Pr (1), Er (2)]: Novel Two-Dimensional Lanthanide Coordination Polymers with 2-Pyrazinecarboxylate and Oxalate. Inorg. Chem. 2006, 45, 10425−10427. (c) Ouyang, Y.; Zhang, W.; Xu, N.; Xu, G.-F.; Liao, D.-Z.; Yoshimura, K.; Yan, S.-P.; Cheng, P. Three-Dimensional 3d−4f Polymers Containing Heterometallic Rings: Syntheses, Structures, and Magnetic Properties. Inorg. Chem. 2007, 46, 8454−8456. (d) Tang, J.-K.; Wang, Q.-L.; Si, S.-F.; Liao, D.-Z.; Jiang, Z.-H.; Yan, S.-P.; Cheng, P. A novel tetranuclear lanthanide(III)− copper(II) complex of the macrocyclic oxamide [PrCu3] (macrocyclic oxamide = 1,4,8,11-tetraazacyclotradecanne-2,3-dione): synthesis, structure and magnetism. Inorg. Chim. Acta 2005, 358, 325−330. (e) Feng, X.; Liu, L.; Wang, L.-Y.; Song, H.-L.; Shi, Z. Q.; Wu, X.-H.; Ng, S.-W. Lanthanide coordination polymers based on multi-donor ligand containing pyridine and phthalate moieties: Structures, luminescence and magnetic properties. J. Solid State Chem. 2013, 206, 277−285. (52) (a) Cepeda, J.; Balda, R.; Beobide, G.; Castillo, O.; Fernandez, J.; Luque, A.; Pérez-Yánez, S.; Román, P.; Vallejo-Sánchez, D. Lanthanide(III)/Pyrimidine-4,6-dicarboxylate/Oxalate Extended Frameworks: A Detailed Study Based on the Lanthanide Contraction and Temperature Effects. Inorg. Chem. 2011, 50, 8437−8451. (b) Liu, Q.-Y.; Wang, W.-F.; Wang, Y.-L.; Shan, Z.-M.; Wang, M.-S.; Tang, J. Diversity of Lanthanide(III)-Organic Extended Frameworks with a 4,8-Disulfonyl-2,6-naphthalenedicarboxylic Acid Ligand: Syntheses, Structures, and Magnetic and Luminescent Properties. Inorg. Chem. 2012, 51, 2381−2392. (53) Ginsberg, A. P.; Martin, R. L.; Brookes, R. W.; Sherwood, R. C. Magnetic exchange in transition metal complexes. IX. Dimeric nickel(II)-ethylenediamine complexes. Inorg. Chem. 1972, 11, 2884− 2889. (54) (a) Feng, X.; Wang, J.; Liu, B.; Wang, L.; Zhao, J.; Ng, S. From Two-Dimensional Double Decker Architecture to Three Dimensional pcu Framework with One-Dimensional Tube: Syntheses, Structures, Luminescence, and Magnetic Studies. Cryst. Growth Des. 2012, 12, 927−938. (b) Feng, X.; Ling, X.-L.; Liu, L.; Song, H.-L.; Wang, L.-Y.; Ng, S.-W.; Su, B. Y. A series of 3D lanthanide frameworks constructed from aromatic multi-carboxylate ligand: Structural diversity, luminescence and magnetic properties. Dalton Trans. 2013, 42, 10292−10303. (55) (a) Pasatoiu, T. D.; Tiseanu, C.; Madalan, A. M.; Jurca, B.; Duhayon, C.; Sutter, J. P.; Andruh, M. Study of the Luminescent and Magnetic Properties of a Series of Heterodinuclear [ZnIILnIII] Complexes. Inorg. Chem. 2011, 50, 5879−5889. (b) Wu, M.-F.; Wang, M.-S.; Guo, S.-P.; Zheng, F.-K.; Chen, H.-F.; Jiang, X.-M.; Liu,

G.-N.; Guo, G.-C.; Huang, J.-S. Photoluminescent and Magnetic Properties of a Series of Lanthanide Coordination Polymers with 1HTetrazolate-5-formic Acid. Cryst. Growth Des. 2011, 11, 372−381. (56) (a) Xu, J.-X.; Ma, Y.; Liao, D.-Z.; Xu, G.-F.; Tang, J.; Wang, C.; Zhou, N.; Yan, S.-P.; Cheng, P.; Li, L.-C. Four New LanthanideNitronyl Nitroxide (LnIII = PrIII, SmIII, EuIII, TmIII) Complexes and a TbIII Complex Exhibiting Single-Molecule Magnet Behavior. Inorg. Chem. 2009, 48, 8890−8896. (b) Xu, N.; Wang, C.; Shi, W.; Yan, S.-P.; Cheng, P.; Liao, D.-Z. Magnetic and Luminescent Properties of Sm, Eu, Tb, and Dy Coordination Polymers with 2-Hydroxynicotinic Acid. Eur. J. Inorg. Chem. 2011, 2387−2393. (57) (a) Liu, C.-M.; Xiong, M.; Zhang, D.-Q.; Du, M.; Zhu, D.-B. Two- and three-dimensional lanthanide−organic frameworks constructed using 1-hydro-6-oxopyridine-3-carboxylate and oxalate ligands. Dalton Trans. 2009, 5666−5672. (b) Feng, X.; Chen, J.-L.; Wang, L.-Y.; Xie, S.-Y.; Yang, S.; Huo, S.-Z.; Ng, S.-W. A series of homonuclear lanthanide complexes incorporating isonicotinic based carboxylate tectonic and oxalate coligand: structures, luminescent and magnetic properties. CrystEngComm 2014, 16, 1334−1343. (58) The measure of the deviation of the fit is R 2 = ∑ [(χM T )obs − (χM T )calc ]2 /(∑ (χM T )obs 2 ).

13165

DOI: 10.1021/acs.inorgchem.7b01944 Inorg. Chem. 2017, 56, 13152−13165