Solvent-Free Synthesis of a Pillared Three ... - ACS Publications

Oct 15, 2015 - M. Eur. J. Inorg. Chem. 2013, 2013, 753−762. (f) Clemente-León, M.;. Coronado, E.; López-Jordà, M.; Waerenborgh, J. C.; Desplanche...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/IC

Solvent-Free Synthesis of a Pillared Three-Dimensional Coordination Polymer with Magnetic Ordering Javier López-Cabrelles,† Mónica Giménez-Marqués,†,‡ Guillermo Mínguez Espallargas,*,† and Eugenio Coronado*,† †

Instituto de Ciencia Molecular (ICMol), Universidad de Valencia, c/Catedrático José Beltrán, 2, 46980 Paterna, Spain Institut Lavoisier CNRS UMR 8180, Université de Versailles St Quentin-en-Yvelines, 45, Avenue des Etats Unis, 78035 Versailles Cedex, France



S Supporting Information *

ABSTRACT: A new magnetic coordination polymer, [Fe(bipy)(im)2] (bipy = 4,4-bipyridine and im = imidazole), has been synthesized in a solvent-free reaction. Structural analysis reveals a pillared 3D coordination polymer composed by neutral layers, formed by iron(II) and imidazolate linkers, interconnected by bipy ligands which serve as pillars. Magnetic measurements show that the material magnetically orders at low temperatures (Tc = 14.5 K) as a weak ferromagnet, likely due to a spin canting.



based solely in imidazolate ligands have been known16 well since before the popularization of ZIFs in 2006.17 Due to the lack of an inversion center along the NCN imidazolate bridge, a spin canting is allowed in these systems, which can lead to the preparation of molecule-based magnets. The synthesis of layered magnets has been attracting an increasing interest in recent years since the discovery of graphene and other two-dimensional (2D) materials.18 Magnetic layered materials based on coordination polymers include anionic layers based on metal tris-oxalate9b−f,19 and anilate metalloligands,20 and neutral layers formed by bimetallic cyanides,21 where the layers are separated by countercations or by van der Waals interactions, respectively. On the contrary, a covalent connection between the layers has been less explored.22 Thus, we decided to combine imidazolates with other coligands in order to prepare a new type of pillar layered magnetic materials based on a metal−imidazolate coordination polymer. In this manner, the neutral layers are covalently connected through a second ligand acting as spacer, therefore implying an increase in the dimensionality of the system and controlling of the organization of the layers. In addition, this approach should maintain an effective magnetic exchange within the neutral layer, while opening the possibility of modulating the interlamellar space by chemical design. In this paper we describe the use of imidazole (im) and 4,4bipyridine (bipy) ligands to synthesize a layered material which can be described as a pillared coordination polymer of Fe(II). The magnetic analysis of the material of formula [Fe(bipy)-

INTRODUCTION Since the pioneering work by Robson,1 coordination polymers and metal−organic frameworks (MOFs)2 have shown great versatility in the design of crystalline materials with a wide range of properties, including gas storage,3 chemical separation,4 catalysis,5 bioapplications,6 and magnetism,7 among others. The tremendous efforts of these years have resulted in the expansion of the initial rigid systems into more flexible ones,7,8 as well as a movement from the presence of a singular physical property to the coexistence or synergy of different physical properties in the same material.9 Within this large area, pillared-layer coordination polymers formed by neutral polymeric layers which are linked in the third direction by additional ligands present an attractive approach for the rapid modulation of the materials through simple modification of the pillar ligand and/or the layer bridge, while keeping the overall structure. Archetypal examples that have been extensively studied include pillared paddle-wheel structures M(O2C−R− CO2)(L), based on carboxylate layers and pyridine-based ligands,10 and pillared Hofmann systems {M(L)[M′(CN)4]}, based on bimetallic cyanide layers and pyridine-based ligands.11 Theses systems are very attractive for the layer-by-layer growth with atomic control,12 and have also been exploited in the formation of core−shell heterostructures.13 More recently, imidazolate-based ligands have been used to incorporate porosity into coordination polymers. Thus, numerous porous solids with zeolitic structures, also known as ZIFs, have been reported.14 Since the NCN bridge can effectively mediate magnetic coupling,15 imidazolates can be suitable building blocks for the preparation of magnetic coordination polymers, as also are cyanides, azide, oxalate, and formate ligands. Indeed, magnetic coordination polymers © XXXX American Chemical Society

Received: August 29, 2015

A

DOI: 10.1021/acs.inorgchem.5b02003 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

charge from The Cambridge Crystallographic Data Centre via www. ccdc.cam.ac.uk/data_request/cif. X-ray Powder Diffraction Measurements (XRD). A polycrystalline sample of 1 was lightly ground in an agate mortar and pestle and filled into a 0.5 mm borosilicate capillary prior to being mounted and aligned on an Empyrean PANalytical powder diffractometer, using Cu Kα radiation (λ = 1.54056 Å). Two repeated measurements were collected at room temperature (2θ = 4−40°) and merged in a single diffractogram. Pawley refinements25 were performed using the TOPAS computer program26 and revealed an excellent fit to a one-phase model for compound 1, indicating the absence of any other detectable crystalline phases. Thermogravimetric Analysis (TGA). Thermogravimetric analysis of 1 was carried out with a Mettler Toledo TGA/SDTA 851 apparatus in the 25−650 °C temperature range under a 10 °C·min−1 scan rate and a N2 flow of 30 mL·min−1. Magnetic Measurements. Variable-temperature (2−300 K) direct current (dc) magnetic susceptibility measurements were carried out in applied fields of 1.0 kOe and 10 Oe and variable field magnetization measurements up to 5 T at 2.0 K. The susceptibility data were corrected from the diamagnetic contributions as deduced by using Pascal’s constant tables. Zero-field-cooled (ZFC)/field-cooled (FC) magnetizations were measured in the 2−20 K interval under an applied field of 10 Oe. Variable-temperature (10−16 K) alternating current (ac) magnetic susceptibility measurements in a ±4.0 G oscillating field at frequencies in the range of 10−997 Hz were carried out in a zero dc field.

(im)2] indicates a spin-canted antiferromagnetic behavior with Tc = 14.5 K.



EXPERIMENTAL SECTION

Synthesis of [Fe(bipy)(im)2] (1). The compound [Fe(bipy)(im)2] was prepared adapting a previously described method for the preparation of iron azolates.23 All reagents were commercially available and used without further purification. Ferrocene (30 mg, 0.16 mmol), 4,4-bipyridine (50 mg, 0.32 mmol), and imidazole (20 mg, 0.30 mmol) were combined and sealed under vacuum in a layering tube (4 mm diameter). The mixture was heated at 150 °C for 12 days to obtain black crystals suitable for X-ray singlecrystal diffraction. The product was allowed to cool to room temperature, and the layering tube was then opened. The unreacted precursors were extracted with acetonitrile and benzene, and the title compound was isolated as black crystals (yield 30%). Phase purity was established by X-ray powder diffraction. X-ray Structural Studies. A single crystal of compound 1 was mounted on a glass fiber using a viscous hydrocarbon oil to coat the crystal. X-ray data were collected at 120 K on a Supernova diffractometer equipped with a graphite-monochromated Enhance (Mo) X-ray source (λ = 0.71073 Å). The program CrysAlisPro, Oxford Diffraction Ltd., was used for unit cell determinations and data reduction. Empirical absorption correction was performed using spherical harmonics, implemented in the SCALE3 ABSPACK scaling algorithm. The crystal structure was solved and refined against all F2 values by using the SHELXTL suite of programs.24 Non-hydrogen atoms were refined anisotropically, and hydrogen atoms were placed in calculated positions that were refined using idealized geometries (riding model) and assigned fixed isotropic displacement parameters. A summary of the data collection and structure refinements is provided in Table 1. CCDC-1420989 contains the supplementary crystallographic data for this paper. This data can be obtained free of



RESULTS AND DISCUSSION The solvent-free reaction of ferrocene with a mixture of 4,4bipy and imidazole results in the formation of black crystals of formula [Fe(bipy)(im)2] (1). Under the above-mentioned synthetic conditions (heating up to 150 °C for 12 days), both 4,4-bipy and imidazole react in liquid phase (melting points of 109−112 and 88−91 °C, respectively), whereas the metal precursor ferrocene reacts in vapor phase (sublimation point of 100 °C), producing a single-phase crystalline sample of 1 (Figure 1). Crystallographic analysis reveals that compound 1 is a pillared three-dimensional coordination polymer, isostructural with the previously reported Cd analogue,27 composed by neutral layers of chemical composition Fe(im)2, which lie parallel to the bc plane and are interconnected by bipy ligands in the a direction (Figure 2). It crystallizes in the monoclinic P21/c space group, with the iron atom lying at the inversion center and coordinated to six nitrogen atoms presenting a distorted octahedral environment with Jahn−Teller distortion. The equatorial positions are occupied by four nitrogen atoms belonging to four imidazolate ligands (Fe−N distances of 2.1863(6) and 2.1905(6) Å), and the axial positions are occupied by two nitrogen ligands from two bipyridine ligands (Fe−N distances of 2.3645(6) Å) as depicted in Figure 3. Imidazolate ligands serve as bridges between Fe atoms within the layers, connecting each metal center with four adjacent ones leading to M···M distances of 6.2486(2) Å. Bipyridine ligands connect metal centers in the third dimension, but with a longer M···M distance of 11.8917(3) Å (Figure 3), thus serving as pillars linking the layers, with a layer−layer distance of 9.069 Å. This reduction in the interlayer distance compared to the M··· M is a result of the tilting of the bipy ligands relative to the layer plane (63.3°). In addition, two different orientations are observed for the bipyridines at alternate metal centers (Figure 4). The coordination polymer [Fe(bipy)(im)2] presents layers of imidazolate-bridged iron ions that give the possibility of a magnetic exchange between metal centers. Magnetic properties

Table 1. Crystallographic Data for Compound 1 empirical formula formula weight crystal color crystal size (mm3) crystal system, Z space group a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) V (Å3) density (mg·m−3) wavelength, λ (Å) temperature (K) μ(Mo Kα) (mm−1) θ range (deg) reflns collected independent reflns (Rint) reflns used in refinement, n L.S. parameters, p no. of restraints, r R1(F),a I > 2σ(I) wR2(F2),b all data S(F2),c all data

C16H14N6Fe 346.18 black 0.15 × 0.12 × 0.04 monoclinic, 2 P21/c 9.0687(3) 7.6924(3) 9.8492(4) 90.00 90.754(3) 90.00 687.02(4) 1.673 0.71073 120(2) 1.106 2.25−41.02 17680 4434 (0.0322) 4434 106 0 0.0250 0.0675 0.945

R1(F) = ∑(|Fo| − |Fc|)/∑|Fo|. bwR2(F2) = [∑w(Fo2 − Fc2)2/ ∑wFo4]1/2. cS(F2) = [∑w(Fo2 − Fc2)2/(n + r − p)]1/2.

a

B

DOI: 10.1021/acs.inorgchem.5b02003 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. Observed (blue) and calculated (red) profiles and difference plot [(Iobs − Icalcd)] (gray) of the Pawley refinement (2θ range 4.0−40.0°).

Figure 3. Iron environment. Iron, nitrogen, and carbon are colored orange, blue, and black, respectively. Hydrogen atoms have been omitted for clarity.

Figure 4. Iron connectivity interlayers. Yellow lines represent the links in the layer (imidazolate bridges), and black lines represent interlayer links (bipyridine pillars).

Figure 2. (top) View of the crystal structure of [Fe(bipy)(im)2] along the b axis showing the neutral layers in blue and pillars connecting these layers in red. (bottom) Projection of the crystal structure onto the bc plane, with the bipy ligands omitted for clarity, showing the Fe(im)2 square-like layer.

−26.3 K. Interestingly, the observed value of the Curie−Weiss temperature is larger (in absolute value) than that of the similar iron(II) compound [Fe3(im)6(Him)2], which presents two different phases (θHT = −6.7 K; θLT = −9.9 K),28 although this difference might be ascribed to the presence of both tetrahedral and octahedral environments in the latter, whereas compound 1 only possesses octahedral iron centers. The abrupt increase of χMT below the χMT minimum indicates the presence of uncompensated magnetic moments in the ground magnetic state. In fact, the behavior below 15 K is strongly fielddependent (Figure 6), suggesting the onset of a long-range antiferromagnetic order with spin canting, which is confirmed by the divergence in the zero-field-cooled (ZFC) and fieldcooled (FC) plots (Figure 7). As can be observed in Figure 8, left, the magnetization vs field plot, M(H), of crushed crystals shows a complex behavior. There is an abrupt increase of the magnetization to 0.10 μB, which steadily increases linearly with the field up to a value of 0.18 μB at 0.9 T. This is in agreement with that expected for a spin-canted antiferromagnet, and a canting angle of 1.4° can be

of 1 are reported in Figure 5 as a plot of the product of the molar magnetic susceptibility times the temperature (χMT) as a function of the temperature under an applied external field of 1000 Oe. At room temperature a χMT value of 3.31 emu·mol−1· K is observed, in agreement with that expected for a high-spin iron(II) ion through the spin-only formula (3.5 emu·mol−1·K where g = 2.0). Upon cooling, the χMT value drops to reach a minimum value of 2.13 emu·mol−1·K at 30 K and a sharp divergence at lower temperatures to reach a maximum value of 7.35 emu·mol−1·K at 13 K followed by a sharp decrease upon further cooling. The decrease in χMT from room temperature to the minimum indicates the presence of dominant antiferromagnetic exchange coupling between the Fe(II) centers. This is confirmed by the plot of 1/χM vs T, which above 40 K follows a Curie−Weiss law, χM = C/(T − θ), with a Curie constant C = 3.59 emu·mol−1·K and a negative Curie−Weiss temperature θ = C

DOI: 10.1021/acs.inorgchem.5b02003 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 5. (left) Thermal dependence of the χMT product in the temperature range 2−300 K. (right) Same set of data as χM−1 vs T together with the fit to the Curie−Weiss law (solid line).

Figure 6. Plots of temperature dependence of χM (left) and χMT (right) for 1 measured at 1 kOe field (blue) and 10 Oe (red).

crystals upon application of the field. To eliminate this artifact, a repetition of the measurement was carried out using eicosane to fix the crystals (Figure 8, right). Under these conditions we observe that the increase of M with H is much softer, as expected for a decoupling of the antiferromagnetic alignment of the spins caused by the magnetic field. In addition, we observe that in the region where the spin canting is dominant (below 1 T) a hysteresis loop with a coercive field of 0.8 T appears, indicating that intrinsically the compound behaves as a hard magnet. The zero-field alternating-current (ac) magnetic susceptibilities were measured at different frequencies (10 Hz−997 Hz) under Hac = 3.95 Oe. Interestingly, we observe that the value at which χ″ becomes different from zero is independent of the frequency (Figure 9). This value is equal to 14.5 K and defines in a precise way Tc. In addition, we observe that both the inphase, χ′, and out-of-phase signals, χ″, are unsymmetrical and have a small frequency dependence (Figure 9). As this behavior is observed below Tc, it occurs in the ordered magnetic state of the material and might be associated with the movement of the domain walls inside this hard magnet. In fact, it follows an Arrhenius law with a value of the activation energy of Δ/kB = 1510 K. This magnetic behavior is analogous to that of the purely two-dimensional systems based on hexacyanometalate anions and polyamine nickel(II) complexes reported previously by some of us,29 where the complex behavior is due to the

Figure 7. Field-cooled (FC) and zero-field-cooled (ZFC) plots of χM vs T at 10 Oe.

estimated through the equation sin(γ) = MR/MS (MS(FeII) = 4 N μB). Upon further increase of the applied field, a second abrupt increase of the magnetization is observed at 0.9 T, with a subsequent smooth linear increase with the field but without reaching saturation. The final value of 1.52 μB at 5 T is well below the expected value for noninteracting Fe(II) centers (4.00 μB for S = 2 and g = 2). Interestingly, no hysteresis loop is observed. A surprising result for a canted antiferromagnet is the abrupt increase of the magnetization observed at intermediate fields. This may be due to an alignment of the anisotropic D

DOI: 10.1021/acs.inorgchem.5b02003 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 8. Magnetic hysteresis loops at 2 K of crushed crystals (left) and embedded in eicosane (right).

Figure 9. Out-of-phase dynamic susceptibility of 1 at different frequencies of crushed crystals (left) and embedded in eicosane (right). Solid lines are included to guide the eyes.

coordination polymers is a highly challenging goal due to the difficulties in connecting paramagnetic metal centers at distances within interacting magnetic range.7 As we have shown, this could be circumvented with the formation of a pillared structure based on predesigned magnetic layers. Through rational modification of the pillar, porosity could be created in this kind of material. In the field of molecule-based magnetism, tremendous effort has been devoted to the design of this type of material given the possibility of tuning the magnetic properties through the sorption of molecules inside the pores.9a,31

pinning of the domain walls resulting from the coexistence of the layered structure and the strong magnetic anisotropy. Notice that a spin-glass-like material also exhibits this kind of behavior. Thus, the calculated value for ϕ (0.020) is in the range of those expected for spin glass like behavior (0.005< ϕ < 0.06), where ϕ has been defined by Mydosh as ϕ= ΔTf/ [TfΔ(log ω)],30 where ΔTf is the peak temperature shift and ω is the frequency of Hac. However, the present material has not the characteristic features of a spin glass (spin frustration and structural disorder) and therefore this possible origin can be excluded.





CONCLUSIONS In this work we have presented a new type of magnetic coordination polymer with formula Fe(im)2(bipy) which orders at 14.5 K. The use of imidazolate ligands is suitable for the formation of a magnetic layer with a spin-canted antiferromagnetic structure, which is further extended in the third direction by using the coligand bipyridine thus resulting in a three-dimensional coordination polymer. Magnetic analysis has revealed a complex behavior with a spin-canted antiferromagnetic structure in the ordered state. The antiferromagnetic exchange interactions, confirmed by the negative θ value observed in the paramagnetic region, propagate through the imidazolate layer and give rise to the magnetic order. In the future the approach here reported will be exploited to obtain porous magnets. In fact, the synthesis of this class of

ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.5b02003. Crystallographic information for compound 1 (C16H14FeN6) (CIF) TGA analysis of compound 1 (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected] (G.M.E.). Notes

The authors declare no competing financial interest. E

DOI: 10.1021/acs.inorgchem.5b02003 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry



2207−2219. (d) Clemente-León, M.; Coronado, E.; López-Jordà, M.; Desplanches, C.; Asthana, S.; Wang, H.; Létard, J.-F. Chem. Sci. 2011, 2, 1121−1127. (e) Clemente-León, M.; Coronado, E.; López-Jordà, M. Eur. J. Inorg. Chem. 2013, 2013, 753−762. (f) Clemente-León, M.; Coronado, E.; López-Jordà, M.; Waerenborgh, J. C.; Desplanches, C.; Wang, H.; Létard, J.-F.; Hauser, A.; Tissot, A. J. Am. Chem. Soc. 2013, 135, 8655−8667. (10) Kondo, M.; Okubo, T.; Asami, A.; Noro, S.-i.; Yoshitomi, T.; Kitagawa, S.; Ishii, T.; Matsuzaka, H.; Seki, K. Angew. Chem., Int. Ed. 1999, 38, 140−143. (11) Muñoz, M. C.; Real, J. A. Coord. Chem. Rev. 2011, 255, 2068− 2093. (12) (a) Bell, C. M.; Arendt, M. F.; Gomez, L.; Schmehl, R. H.; Mallouk, T. E. J. Am. Chem. Soc. 1994, 116, 8374−8375. (b) Molnár, G.; Cobo, S.; Real, J. A.; Carcenac, F.; Daran, E.; Vieu, C.; Bousseksou, A. Adv. Mater. 2007, 19, 2163−2167. (c) Shekhah, O.; Liu, J.; Fischer, R. A.; Wöll, Ch. Chem. Soc. Rev. 2011, 40, 1081−1106. (d) Bétard, A.; Fischer, R. A. Chem. Rev. 2012, 112, 1055−1083. (13) (a) Furukawa, S.; Hirai, K.; Nakagawa, K.; Takashima, Y.; Matsuda, R.; Tsuruoka, T.; Kondo, M.; Haruki, R.; Tanaka, D.; Sakamoto, H.; Shimomura, S.; Sakata, O.; Kitagawa, S. Angew. Chem., Int. Ed. 2009, 48, 1766−1770. (b) Hirai, K.; Furukawa, S.; Kondo, M.; Uehara, H.; Sakata, O.; Kitagawa, S. Angew. Chem., Int. Ed. 2011, 50, 8057−8061. (14) Phan, A.; Doonan, C. J.; Uribe-Romo, F. J.; Knobler, C. B.; O’Keeffe, M.; Yaghi, O. M. Acc. Chem. Res. 2010, 43, 58−67. (15) Coronado, E.; Giménez-Marqués, M.; Mínguez Espallargas, G.; Brammer, L. Nat. Commun. 2012, 3, 828. (16) Tian, Y.-Q.; Cai, C.-X.; Ren, X.-M.; Duan, C.-Y.; Xu, Y.; Gao, S.; You, X.-Z. Chem. - Eur. J. 2003, 9, 5673−5685. (17) Park, K. S.; Ni, Z.; Côté, A. P.; Choi, J. Y.; Huang, R.; UribeRomo, F. J.; Chae, H. K.; O’Keeffe, M.; Yaghi, O. M. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 10186−10191. (18) Ferrari, A. C.; Bonaccorso, F.; Fal’ko, V.; Novoselov, K. S.; Roche, S.; Bøggild, P.; Borini, S.; Koppens, F. H. L.; Palermo, V.; Pugno, N.; Garrido, J. A.; Sordan, R.; Bianco, A.; Ballerini, L.; Prato, M.; Lidorikis, E.; Kivioja, J.; Marinelli, C.; Ryhänen, T.; Morpurgo, A.; Coleman, J. N.; Nicolosi, V.; Colombo, L.; Fert, A.; Garcia-Hernandez, M.; Bachtold, A.; Schneider, G. F.; Guinea, F.; Dekker, C.; Barbone, M.; Sun, Z.; Galiotis, C.; Grigorenko, A. N.; Konstantatos, G.; Kis, A.; Katsnelson, M.; Vandersypen, L.; Loiseau, A.; Morandi, V.; Neumaier, D.; Treossi, E.; Pellegrini, V.; Polini, M.; Tredicucci, A.; Williams, G. M.; Hong, B. H.; Ahn, J.-H.; Kim, J. M.; Zirath, H.; van Wees, B. J.; van der Zant, H.; Occhipinti, L.; Di Matteo, A.; Kinloch, I. A.; Seyller, T.; Quesnel, E.; Feng, X.; Teo, K.; Rupesinghe, N.; Hakonen, P.; Neil, S. R. T.; Tannock, Q.; Löfwander, T.; Kinaret, J. Nanoscale 2015, 7, 4598−4810. (19) (a) Coronado, E.; Galán-Mascarós, J. R.; Gómez-García, C. J.; Laukhin, V. Nature 2000, 408, 447−449. (b) Kida, N.; Hikita, M.; Kashima, I.; Okubo, M.; Itoi, M.; Enomoto, M.; Kato, K.; Takata, M.; Kojima, N. J. Am. Chem. Soc. 2009, 131, 212−220. (c) Train, C.; Gheorghe, R.; Krstic, V.; Chamoreau, L.; Ovanesyan, N. S.; Rikken, G. L. J. A.; Gruselle, M.; Verdaguer, M. Nat. Mater. 2008, 7, 729−734. (d) Okawa, H.; Shigematsu, A.; Sadakiyo, M.; Miyagawa, T.; Yoneda, K.; Ohba, M.; Kitagawa, H. J. Am. Chem. Soc. 2009, 131, 13516− 13522. (20) (a) Atzori, M.; Benmansour, S.; Mínguez Espallargas, G.; Clemente-León, M.; Abhervé, A.; Gómez-Claramunt, P.; Coronado, E.; Artizzu, F.; Sessini, E.; Deplano, P.; Serpe, A.; Mercuri, M. L.; Gómez García, C. J. Inorg. Chem. 2013, 52, 10031−10040. (b) Abhervé, A.; Clemente-León, M.; Coronado, E.; Gómez-García, C. J.; Verneret, M. Inorg. Chem. 2014, 53, 12014−12026. (c) Abhervé, A.; Mañas-Valero, S.; Clemente-León, M.; Coronado, E. Chem. Sci. 2015, 6, 4665−4673. (21) (a) Nakabayashi, K.; Chorazy, S.; Takahashi, D.; Kinoshita, T.; Sieklucka, B.; Ohkoshi, S.-i. Cryst. Growth Des. 2014, 14, 6093−6100. (b) Colacio, E.; Lloret, F.; Navarrete, M.; Romerosa, A.; StoeckliEvans, H.; Suarez-Varela, J. New J. Chem. 2005, 29, 1189−1194. (c) Marvilliers, A.; Parsons, S.; Rivière, E.; Audière, J.-P.; Kurmoo, M.;

ACKNOWLEDGMENTS We are grateful to the Spanish MINECO (Projects MAT-201456143-R and CTQ-2014-59209-P), the EU (ERC Advanced Grant SPINMOL), and the Generalidad Valenciana (Prometeo and ISIC-Nano Programs). M.G.-M. thanks MICINN for a predoctoral FPU grant and the EU for a Marie SklodowskaCurie postdoctoral fellowship (H2020-MSCA-IF-EF-658224). G.M.E. thanks the Spanish MINECO for a Ramón y Cajal ́ Fellowship. We also acknowledge J. M. Martinez-Agudo and G. Agusti ́ (University of Valencia) for their help with the magnetic measurements.



REFERENCES

(1) (a) Hoskins, B. F.; Robson, R. J. Am. Chem. Soc. 1989, 111, 5962−5964. (b) Hoskins, B. F.; Robson, R. J. Am. Chem. Soc. 1990, 112, 1546−1554. (2) See the recent themed issues on MOFs: (a) Chem. Soc. Rev. 2014, 43, 5415−6172. (b) Chem. Rev. 2012, 112, 673−1268. (3) (a) Murray, L. J.; Dincă, M.; Long, J. R. Chem. Soc. Rev. 2009, 38, 1294−1314. (b) Lin, X.; Jia, J.; Hubberstey, P.; Schröder, M.; Champness, N. R. CrystEngComm 2007, 9, 438−448. (c) D'Alessandro, D. M.; Smit, B.; Long, J. R. Angew. Chem., Int. Ed. 2010, 49, 6058− 6082. (d) He, Y.; Zhou, W.; Qian, G.; Chen, B. Chem. Soc. Rev. 2014, 43, 5657−5678. (e) Carrington, E. J.; Vitórica-Yrezábal, I. J.; Brammer, L. Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater. 2014, 70, 404−422. (4) (a) Li, J.-R.; Sculley, J.; Zhou, H.-C. Chem. Rev. 2012, 112, 869− 932. (b) Van de Voorde, B.; Bueken, B.; Denayer, J.; De Vos, D. Chem. Soc. Rev. 2014, 43, 5766−5788. (c) Banerjee, D.; Cairns, A. J.; Liu, J.; Motkuri, R. K.; Nune, S. K.; Fernandez, C. A.; Krishna, R.; Strachan, D. M.; Thallapally, P. K. Acc. Chem. Res. 2015, 48, 211−219. (5) (a) Lee, J. Y.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. Chem. Soc. Rev. 2009, 38, 1450−1459. (b) Ma, L.; Abney, C.; Lin, W. Chem. Soc. Rev. 2009, 38, 1248−1256. (c) Liu, J.; Chen, L.; Cui, H.; Zhang, J.; Zhang, L.; Su, C.-Y. Chem. Soc. Rev. 2014, 43, 6011−6061. (d) Zhang, T.; Lin, W. Chem. Soc. Rev. 2014, 43, 5982−5993. (e) Dhakshinamoorthy, A.; Garcia, H. Chem. Soc. Rev. 2014, 43, 5750−5765. (6) Giménez-Marqués, M.; Hidalgo, T.; Serre, C.; Horcajada, P. Coord. Chem. Rev. 2015, DOI: 10.1016/j.ccr.2015.08.008. (7) Coronado, E.; Mínguez Espallargas, G. Chem. Soc. Rev. 2013, 42, 1525−1539. (8) (a) Ferey, G.; Serre, C. Chem. Soc. Rev. 2009, 38, 1380−1399. (b) Horike, S.; Shimomura, S.; Kitagawa, S. Nat. Chem. 2009, 1, 695− 704. (c) Pigge, F. C. CrystEngComm 2011, 13, 1733−1748. (d) Fletcher, A. J.; Thomas, K. M.; Rosseinsky, M. J. J. Solid State Chem. 2005, 178, 2491−2510. (e) Coronado, E.; Giménez-Marqués, M.; Mínguez Espallargas, G. Inorg. Chem. 2012, 51, 4403−4410. (f) Coronado, E.; Giménez-Marqués, M.; Gómez-García, C. J.; Mínguez Espallargas, G. Inorg. Chem. 2012, 51, 12938−12947. (g) Hawxwell, S. M.; Mínguez Espallargas, G.; Bradshaw, D.; Rosseinsky, M. J.; Prior, T. J.; Florence, A. J.; van de Streek, J.; Brammer, L. Chem. Commun. 2007, 1532−1534. (h) Libri, S.; Mahler, M.; Mínguez Espallargas, G.; Singh, D. C. N. G.; Soleimannejad, J.; Adams, H.; Burgard, M. D.; Rath, N. P.; Brunelli, M.; Brammer, L. Angew. Chem., Int. Ed. 2008, 47, 1693−1697. (i) Vitórica-Yrezábal, I. J.; Mínguez Espallargas, G.; Soleimannejad, J.; Florence, A. J.; Fletcher, A. J.; Brammer, L. Chem. Sci. 2013, 4, 696−708. (j) Vitórica-Yrezábal, I. J.; Libri, S.; Loader, J. R.; Mínguez Espallargas, G.; Hippler, M.; Fletcher, A. J.; Thompson, S. P.; Warren, J. E.; Musumeci, D.; Ward, M. D.; Brammer, L. Chem. - Eur. J. 2015, 21, 8799−8811. (9) (a) Coronado, E.; Giménez-Marqués, M.; Mínguez Espallargas, G.; Rey, F.; Vitórica-Yrezábal, I. J. J. Am. Chem. Soc. 2013, 135, 15986− 15989. (b) Clemente-León, M.; Coronado, E.; Martí-Gastaldo, C.; Romero, F. M. Chem. Soc. Rev. 2011, 40, 473−497. (c) ClementeLeón, M.; Coronado, E.; López-Jordà, M.; Mínguez Espallargas, G.; Soriano-Portillo, A.; Waerenborgh, J. C. Chem. - Eur. J. 2010, 16, F

DOI: 10.1021/acs.inorgchem.5b02003 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Mallah, T. Eur. J. Inorg. Chem. 2001, 2001, 1287−1293. (d) Ohba, M.; Okawa, H.; Ito, T.; Ohto, A. J. Chem. Soc., Chem. Commun. 1995, 1545−1546. (e) Re, N.; Gallo, E.; Floriani, C.; Miyasaka, H.; Matsumoto, N. Inorg. Chem. 1996, 35, 6004−6008. (f) Ferlay, S.; Mallah, T.; Vaissermann, J.; Bartolome, F.; Veillet, P.; Verdaguer, M. Chem. Commun. 1996, 2481−2482. (g) Ohba, M.; Okawa, H.; Fukita, N.; Hashimoto, Y. J. Am. Chem. Soc. 1997, 119, 1011−1019. (h) Miyasaka, H.; Matsumoto, N.; Re, N.; Gallo, E.; Floriani, C. Inorg. Chem. 1997, 36, 670−676. (i) Re, N.; Crescenzi, R.; Floriani, C.; Miyasaka, H.; Matsumoto, N. Inorg. Chem. 1998, 37, 2717−2722. (j) Colacio, E.; Dominguez-Vera, J. M.; Ghazi, M.; Kivekäs, R.; Lloret, F.; Moreno, J. M.; Stoeckli-Evans, H. Chem. Commun. 1999, 987−988. (22) Sapiña, F.; Coronado, E.; Beltrán, D.; Burriel, R. J. Am. Chem. Soc. 1991, 113, 7940−7944. (23) Patrick, B. O.; Reiff, W. M.; Sánchez, V.; Storr, A.; Thompson, R. C. Inorg. Chem. 2004, 43, 2330−2339. (24) Sheldrick, G. M. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112−122. (25) Pawley, G. S. J. Appl. Crystallogr. 1981, 14, 357−361. (26) Coelho, A. A. TOPAS-Academic, version 4.1. 2007. See: http:// www.topas-academic.net. (27) Han, J.-Y.; Fang, J.; Chang, H.-Y.; Dong, Y.; Liang, S. Acta Crystallogr., Sect. E: Struct. Rep. Online 2005, 61, m2667−m2669. (28) (a) Rettig, S. J.; Storr, A.; Summers, D. A.; Thompson, R. C.; Trotter, J. J. Am. Chem. Soc. 1997, 119, 8675−8680. (b) Sánchez, V. Ph.D. Thesis, University of Columbia, 2001. (29) (a) Coronado, E.; Gómez-García, C. J.; Nuez, A.; Romero, F. M.; Rusanov, E.; Stoeckli-Evans, H. Inorg. Chem. 2002, 41, 4615− 4617. (b) Coronado, E.; Gómez-García, C. J.; Nuez, A.; Romero, F. M.; Waerenborgh, J. C. Chem. Mater. 2006, 18, 2670−2681. (30) Mydosh, J. A. Spin Glasses: An Experimental Introduction; Taylor & Francis: London, 1993. (31) (a) Neville, S. M.; Halder, G. J.; Chapman, K. W.; Duriska, M. B.; Moubaraki, B.; Murray, K. S.; Kepert, C. J. J. Am. Chem. Soc. 2009, 131, 12106−12108. (b) Southon, P. D.; Liu, L.; Fellows, E. A.; Price, D. J.; Halder, G. J.; Chapman, K. W.; Moubaraki, B.; Murray, K. S.; Létard, J.-F.; Kepert, C. J. J. Am. Chem. Soc. 2009, 131, 10998−11009. (c) Ohba, M.; Yoneda, K.; Agustí, G.; Muñoz, M. C.; Gaspar, A. B.; Real, J. A.; Yamasaki, M.; Ando, H.; Nakao, Y.; Sakaki, S.; Kitagawa, S. Angew. Chem., Int. Ed. 2009, 48, 4767−4771. (d) Wriedt, M.; Yakovenko, A. A.; Halder, G. J.; Prosvirin, A. V.; Dunbar, K. R.; Zhou, H.-C. J. Am. Chem. Soc. 2013, 135, 4040−4050. (e) Ferrando-Soria, J.; Serra-Crespo, P.; de Lange, M.; Gascon, J.; Kapteijn, F.; Julve, M.; Cano, J.; Lloret, F.; Pasán, J.; Ruiz-Pérez, C.; Journaux, Y.; Pardo, E. J. Am. Chem. Soc. 2012, 134, 15301−15304. (f) Navarro, J. A. R.; Barea, E.; Rodríguez-Diéguez, A.; Salas, J. M.; Ania, C. O.; Parra, J. B.; Masciocchi, N.; Galli, S.; Sironi, A. J. Am. Chem. Soc. 2008, 130, 3978− 3984.

G

DOI: 10.1021/acs.inorgchem.5b02003 Inorg. Chem. XXXX, XXX, XXX−XXX