Spin-Crossover - ACS Publications

vibration.14,16 The domain model16,17 indicates that the present compound is an almost uncooperative system. ... signals below 90 K are free from exch...
4 downloads 19 Views 3MB Size
Subscriber access provided by Northwestern Univ. Library

Communication

A New S = 0 # S = 2 ‘Spin-Crossover’ Scenario Found in a Nickel(II) Bis(nitroxide) System Yuta Homma, and Takayuki Ishida Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.7b05357 • Publication Date (Web): 10 Jan 2018 Downloaded from http://pubs.acs.org on January 10, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Chemistry of Materials is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 5 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

A New S = 0 ⇄ S = 2 ‘Spin-Crossover’ Scenario Found in a Nickel(II) Bis(nitroxide) System Yuta Homma and Takayuki Ishida* Department of Engineering Science, The University of Electro-Communications, Chofu, Tokyo 182-8585, Japan.

ABSTRACT: Compound [Ni(phpyNO)2Cl2] was prepared as a pseudo-symmetrical 2p-3d-2p heterospin molecule (phpyNO = tert-butyl

5-phenyl-2-pyridyl nitroxide). Variable-temperature X-ray structure analysis recorded the Ni-O-N-C2py torsion angles of 13.0(4) and 25.8(3)° at 400 K and 29.2(2) and 38.9(2)° at 85 K. The magnetic study revealed that the nickel-radical exchange couplings changed from ferro- to antiferromagnetic ones on cooling, affording an equilibrium between the Stotal = 2 and 0 states.

Bistable solid-sate materials without any change in chemical composition, spin-crossover complexes for instance, are of increasing interest for future applications to memory, display, and other devices.1-3 Transition between S = 0 and S = 2 states in discrete iron(II) 3d6 complexes is typical.3 Heterospin systems provide the wide diversity of the characters of paramagnetic centers including the symmetry of magnetic orbitals.4 Direct copper(II)–radical coordination compounds are the best documented among the 3d-2p heterospin molecules,5 and nickel(II)-radical compounds are the second but relatively scarce.6 Recently nickel(II)-iminosemquinonate chelates have been reported to exhibit ferromagnetic coupling, after the pioneering work on the copper(II)-semiquinonate compounds reported by Kahn and co-workers.7 The equatorial coordination in copper(II)-nitroxide radical complexes is well-known to favor strong antiferromagnetic coupling,5 but ferromagnetic interaction actually is available as well.5b,8 The magnitudes of the couplings often exceed the order of 300 K, whether ferroor antiferromagnetic, being sensitive to the geometry. Ferromagnetic coupling can be realized when two magnetic orbitals (radical π* and metal dσ) are strictly orthogonal. On the other hand, severe twist around the Cu-O (or Ni-O) coordination bond would lead to an appreciable overlap between Cu (or Ni) 3dx2-y2 and 3dz2 and O 2pz magnetic orbitals, affording antiferromagnetic coupling.8 Spin-transition-like behavior in threeor four-centered spin systems has been reported owing to the chelate-ring deformation9,10 by using tert-butyl 2-pyridyl nitroxides as a paramagnetic ligand. We will report here the first example of nickel(II)-radical compounds showing spintransition-like behavior. An advantage of the present system resides in the feature where the drastic spin transition occurs between dia- and paramagnetic states. This report includes a mechanistic investigation of the spintransition-like behavior, and the ground high- and low-spin states are directly regulated with the switch of Ni2+-radical exchange couplings. The situation is completely different from those of the known materials having spin-crossover ions and additional exchange coupling.11 A 2p-3d-2p heterospin triad [Ni(phpyNO)2Cl2] (1) was pre-

pared by complex formation of NiCl2•6H2O with a paramagnetic ligand, tert-butyl 5-phenyl-2-pyridyl nitroxide (phpyNO) ligand8c (Srad = 1/2). The spectroscopic and analytic characterizations were satisfactory, and finally the molecular structure was confirmed by X-ray crystallographic study (Fig. 1). On cooling, chelate ring distortion very gradually took place in a single-crystal-to-single-crystal manner, and we determined the molecular structures at any temperature between 85 and 400 K (a)

(b)

Figure 1. (a) X-ray crystal structures of 1 measured at 400 K (light tone) and 85 K (dark tone). Thermal ellipsoids are drawn at the 50% probability levels. Structural formula of 1 is also shown (right). (b) The cell parameters and the torsion angles as a function of temperature.

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(Fig. S1, Supporting Information). The space group retains monoclinic P21/c. The nickel(II) ion has an octahedral geometry, which guarantees the high spin (SNi = 1). The phpyNO ligands are arranged in a cis position; the O1-Ni-O2 angle is 80.44(11)° at 400 K. The molecule has a pseudo-C2 symmetry; namely the two coordination geometries are similar but independent. The five-membered chelate rings are clearly depicted in Fig. 1a. The nitroxide oxygen atom is directly coordinate to the nickel(II) ion with the Ni-O1 and Ni-O2 distances of 2.083(3) and 2.103(3) Å, respectively, at 400 K. The unit cell smoothly shrinks on cooling (Fig. 1b). The NiO1, Ni-O2, and other bond lengths hardly changed. On the other hand, the Ni-O-N-C2py torsion angles (φ) dramatically change from |φ1| = 13.0(4) and |φ2| = 25.8(3)° at 400 K to |φ1| = 29.2(2) and |φ2| = 38.9(2)° at 85 K. In short, the Ni-O bonds seem to rotate, as indicated with small curved arrows in Fig. 1a. We have proposed8 that φ is a useful metric for orthogonal geometry between the nitroxide π* and Ni2+ 3dx2-y2 and 3dz2 orbitals (Scheme 1(i,iii)).5 At low temperatures, the breakdown of the d-π* orthogonality brings about antiferromagnetic coupling (Scheme 1(ii,iv)). Thus, the present geometrical deformation would enhance antiferromagnetic coupling.

Page 2 of 5

found that the data perfectly obeyed the van’t Hoff equation (eq. 1), which is available for spin-crossover compounds.14

x(HS) =

1 1 + exp[(ΔH / R)(1 / T − 1 / Tc)]

(eq. 1)

The molar fraction of the high-spin molecules (x(HS)) is proportional to the χmT value, and accordingly χmT = C0 x(HS) + C1. The optimized parameters were: Tc (transition temperature) = 173.7(8) K and ΔH (enthalpy change of the phase transition) = 4.57(5) kJ mol–1 with the Curie constant for the highspin state (C0) was 2.38(2) cm3 K mol–1 and impurity constant (C1) 0.093(4) cm3 K mol–1. The entropy change of the phase transition was estimated as ΔS = 26.3(1) J K–1 mol–1.

Scheme 1. Mutual geometries between Ni 3dx2-y2 and O 2pz (i,ii) and between Ni 3dz2 and O 2pz (iii,iv)a)

(i)

(ii)

O N N

Ni (iii)

N N

O

Ni

(iv)

N N

Ni

O N

N

Figure 2. Temperature dependence of χmT for polycrystalline 1, measured in the applied magnetic field was 5000 Oe on heating, cooling, and heating again. For the solid line, see the main text.

O Ni

a)

Only radical chelate planes are shown. The oxygen atoms are located on the plane (i,iii) and dislocated out of the plane (ii,iv).

Magnetic susceptibility of 1 was measured on a SQUID magnetometer (Fig. 2). The χmT value was 2.14 cm3 K mol–1 at 400 K, which is larger than the calculated spin-only value of two radicals and one nickel(II) ion (1.96 cm3 K mol–1 from typical isotropic gNi is close to 2.212). The positive slope at 400 K indicates that the high-temperature limit would be ca. 2.38 cm3 K mol–1 (see below), leading to an unrealistic gNi = 2.55. It does not match the value from the EPR results (see below). This finding implies the presence of ferromagnetic interaction observable even around room temperature. On cooling, the χmT value was monotonically decreased, indicating the presence of considerably large antiferromagnetic coupling. Note that the temperature region of this behavior (around 170 K) corresponds to that of the structural change (Fig. 1). Below 50 K the specimen was practically diamagnetic (Stotal = 0). Repeated cycles exhibited complete reproducibility and no thermal hysteresis. The change to the molecular structure causes the change to the intramolecular exchange coupling. Simulation was unsuccessful to fit the data to any van Vleck equations derived from the spin-Hamiltonian on linear or triangular spin systems13 with S = 1/2, 1, and 1/2, and it can be rationalized by noting the temperature-dependent exchange coupling. Instead, we

Figure 3. Temperature dependence of EPR spectra of polycrystalline 1. (top) Simulation and (bottom) observation.

At the low-temperature (LT) phase, the ground Stotal is 0. Both exchange couplings (J1 and J2) are antiferromagnetic, where the spin-Hamiltonian was defined as H = –J1SRad1•SNi – J2SRad2•SNi. Assuming the ground Stotal = 2 in the hightemperature (HT) phase, the entropy change from the spin multiplicity should be ΔSLH = R ln 5 = 13.4 J K–1 mol–1.15 The observed value has an excess entropy change, and the HT phase is assumed to involve other degrees of freedom such as vibration.14,16 The domain model16,17 indicates that the present compound is an almost uncooperative system. The crystal structure analysis indicates that there are a few hydrogen-bond contacts with respect to the chlorine atoms (2.6 – 2.8 Å; Fig. S2, Supporting Information), but they seem not so beneficial to cooperativity. The variable-temperature X-band EPR study on polycrystalline 1 was performed (Fig. 3). In the LT phase, two resolved signals appeared, which are assigned to the nickel(II) and radical spins. The simulation curve was drawn on a WIN-EPR

ACS Paragon Plus Environment

Page 3 of 5 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials SimFonia software18 with gx = 2.358, gy = 2.358, gz = 2.368, |D|/gµB = 18.5 mT, and |E|/gµB = 2.9 mT for the former, and gx = 1.986, gy = 2.005, gz = 2.018 for the latter. The resolved signals below 90 K are free from exchange coupling, suggesting the low concentration of the spin. The paramagnetic species in the LT phase are attributed to the lattice defect because it appeared only after the structural and spin transition. Namely, the majority of the molecules are EPR-silent below 90 K. This finding is consistent with the ground diamagnetic state for this morph. (a)

(b)

(c)

(d)

(e)

convert the roles of axial and equatorial sites.22 It should be stressed that the present spin-transition phenomenon takes place in a completely different way from theirs. We have found an advantage in the nickel(II) system; the spin-transition occurs regardless of the coordination sites, axial or equatorial, and furthermore two spin-transitions simultaneously occur in a molecule. The trigger of the transition is not elongation or shortening of coordination bonds. The angular torsion around the coordination bond is essential in the magneto-structural relationship. This logic holds also for Gd3+-nitroxide systems,23 and it has very recently been reported that the exchange coupling modulation in the Gd3+-Cu2+-nitroxide compound was associated with the rotation around the Gd3+-O(nitroxide) bond.24 The present Ni2+-bis(nitroxide) system may be advantageous to future application in the drastic dia-/paramagnetic spin-state change, like the iron(II) spin crossover. Moreover, the present system is close to the entropy-driven spin-crossover.25 The spin multiplicity is responsible for the number of microstates, and the TΔS term in ΔG = ΔH – TΔS would be substantial. The two Ni-O-N-C2py distortions were synchronized, and the ferromagnetic couplings become antiferromagnetic on both sides, because the degree of spin freedom is minimized (Stotal = 0). The results on [CuII(phpyNO)2(H2O)2](BF4)29 confirm this notion. The coupling switch occurred only on one side because of Stotal = 1/2, regardless of antiferromagnetic coupling on one side or both. No synchronized coupling switch is necessary in that case.

Figure 4. DFT Calculation results. (a) A model molecule. Substituents colored in red are introduced for reducing the calculation cost. (b) Spin density surfaces of the ground quintet state and (c) the excited singlet state at 400 K. (d) Spin density surfaces of the ground singlet state and (e) the excited quintet state at 85 K. Dark and light lobes stand for the positive and negative spin densities, respectively.

The density functional theory (DFT) MO calculation19 supports the antiferro- and ferromagnetic couplings are present in the LT and HT phases, respectively. For reducing calculation cost, the peripheral phenyl groups were replaced with hydrogen atoms, and the tert-butyl groups with methyl groups (Fig. 4a). The geometries of other atoms were available from the Xray crystallographic analysis. The SCF energies were calculated on the UB3LYP/6-311+G(d,p) level for the structure at 400 K and the quintet state is more stable by ca. 68 K.20 The spin density surfaces are drawn in Figs. 4b and 4c, and the spin structure of the ground state was confirmed to be Rad1(↑)-Ni(↑↑)-Rad2(↑). In contrast, the calculation on the structure at 85 K indicated that the singlet state was ground with the spin-structure of Rad1(↓)-Ni(↑↑)-Rad2(↓) and that the energy gap was as large as 1600 K (Figs. 4d and 4e).20 Such sensitive structural dependence has also been reported on a nickel(II) complex doubly chelated with 5-formylpyrrolyl nitronyl nitroxide.21 The theoretical treatment told us that a few degrees of the Ni-O bond rotation drastically changed the exchange coupling by an order of the magnitude as well as the sign of couplings. This work entirely supports our results. Ovcharenko and co-workers have reported several copper(II)–nitroxide complexes showing spin-transition-like behavior, and clarified that the Cu–O bond lengths change to

One may still wonder: Why is the singlet-quintet equilibrium13 insufficient for the entropy loss with a constant antiferromagnetic exchange coupling? Why does the molecule have to struggle and bend? A plausible answer is that the HT form has ferromagnetic coupling and appreciable entropy can be discarded only when the coupling is switched to be antiferromagnetic. If the HT form had antiferromagnetic coupling or practically no coupling, there would never be a driving force for the molecular deformation. In fact, the molecule has to deform on both sides to accommodate the Stotal = 0 state and corresponding structure. This finding implies that the Stotal = 2 state cannot stand in the HT form on cooling. In summary, the spin triad 1 exhibited a thermally induced spin-transition (or crossover26) on the whole molecular basis. Thanks to the small geometrical change, i.e., inner angular torsion, 1 completely maintains the single-crystalline form throughout the transition. Such spin-state equilibrium has unprecedentedly been characterized in multi-centered heterospin compounds, and is plausibly explained in terms of the entropydriven spin-crossover. Furthermore, this spin-transition drastically switches magnetic properties between dia- and paramagnetic states; actually 1 is the first non-iron(II) system to undergo an S = 0 ⇄ S = 2 spin-crossover. The present system will open a new class of spin-transition compounds and may be utilized as a potential sensor for external stimuli such as heat.

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.chemmater.xxx. Crystallographic data of 1 at 85 K (CIF) Crystallographic data of 1 at 400 K (CIF) Crystal structures at each temperature in Figure 1 (AVI) Molecular arrangement in the crystal, experimental details, and complete reference for Gaussian03 (PDF)

ACS Paragon Plus Environment

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

AUTHOR INFORMATION Corresponding Author * E-mail: [email protected] ORCID Takayuki Ishida: 0000-0001-9088-2526 Note The authors declare no conflict of interest.

ACKNOWLEDGMENT This work was partly supported by KAKENHI (Grant Number JSPS/15H03793).

REFERENCES (1) Gütlich, P.; Goodwin, H. A. Eds. Spin Crossover in Transition Metal Compounds I, II, and III. Springer-Verlag: Berlin, 2004. (2) (a) Halcrow, M.A. Spin-Crossover Materials: Properties and Applications. John Wiley & Sons, Ltd.: Oxford, UK, 2013. (b) Halcrow, M.A. Spin-crossover Compounds with Wide Thermal Hysteresis. Chem. Lett. 2014, 43, 1178–1188. (3) (a) Kepp, K. P. Theoretical Study of Spin Crossover in 30 Iron Complexes, Inorg. Chem. 2016, 55, 2717-2727. (b) Aromi, G.; Real, J. A. Special Issue “Spin Crossover (SCO) Research.” Magnetochem. 2016, 2, 28(1)-28(3). (4) Miller, J.S.; Gatteschi, D. Eds. “Molecule-Based Magnets” Themed Issue, Chem. Soc. Rev. 2011, 40, 3053-3068. (5) (a) Caneschi, A.; Gatteschi, D.; Sessoli, R.; Rey, P. Toward Molecular Magnets: the Metal-Radical Approach. Acc. Chem. Res. 1989, 22, 392-398. (b) Luneau, D.; Rey, P.; Laugir, J.; Fries, P.; Caneschi, A.; Gatteschi, D.; Sessoli, R. N-Bonded Copper(II)-Imino Nitroxide Complexes Exhibiting Large Ferromagnetic Interactions. J. Am. Chem. Soc. 1991, 113, 1245-1251. (6) Luneau, D.; Rey, P.; Laugier, J.; Belorizky, E.; Conge, A. Ferromagnetic Behavior of Nickel(II)-Imino Nitroxide Derivatives. Inorg. Chem. 1992, 31, 3578-3584. (7) (a) Ali, A.; Dhar, D.; Barman, S.K.; Lloret, F.; Mukherjee, R. Nickel(II) Complex of a Hexadentate Ligand with Two oIminosemiquinonato(1-) π-Radical Units and Its Monocation and Dication. Inorg. Chem. 2016, 55, 5759-5771. (b) Kahn, O.; Prins, R.; Reedijk, J.; Thompson, J.S. Orbital Symmetries and Magnetic Interaction between Copper(II) Ions and the o-Semiquinone Radical. Magnetic Studies of (Di-2-pyridylamine)(3,5-di-tert-butyl-osemiquinonato)copper(II) Perchlorate and Bis(bis(3,5-di-tert-butyl-osemiquinonato)copper(II)). Inorg. Chem. 1987, 26, 3557-3561. (8) (a) Osanai, K.; Okazawa, A.; Nogami, T.; Ishida, T. Strong Ferromagnetic Exchange Couplings in Copper(II) and Nickel(II) Complexes with a Paramagnetic Tridentate Chelate Ligand, 2,2‘Bipyridin-6-yl tert-Butyl Nitroxide. J. Am. Chem. Soc. 2006, 128, 14008-14009. (b) Okazawa, A.; Nogami, T.; Ishida, T. tert-Butyl 2Pyridyl Nitroxide Available as a Paramagnetic Chelate Ligand for Strongly Exchange-Coupled Metal−Radical Compounds. Chem. Mater. 2007, 19, 2733-2735. (c) Okazawa, A.; Nagaichi, Y.; Nogami, T.; Ishida, T. Magneto-Structure Relationship in Copper(II) and Nickel(II) Complexes Chelated with Stable tert-Butyl 5-Phenyl-2-pyridyl Nitroxide and Related Radicals. Inorg. Chem. 2008, 47, 8859-8868. (9) Okazawa, A.; Ishida, T. Spin-Transition-Like Behavior on One Side in a Nitroxide-Copper(II)-Nitroxide Triad System. Inorg. Chem. 2010, 49, 10144-10147. (10) Okazawa, A.; Hashizume, D.; Ishida, T. Ferro- and Antiferromagnetic Coupling Switch Accompanied by Twist Deformation around the Copper(II) and Nitroxide Coordination Bond. J. Am. Chem. Soc. 2010, 132, 11516-11524. (11) Gass, I.A.; Tewary, S.; Nafady, A.; Chilton, N.F.; Gartshore, C.J.; Asadi, M.; Lupton, D.W.; Moubaraki, B.; Bond, A. M.; Boas, J.F.; Guo, S.-X.; Rajaraman, G.; Murray, K.S. Observation of Ferromagnetic Exchange, Spin Crossover, Reductively Induced Oxidation,

Page 4 of 5

and Field-Induced Slow Magnetic Relaxation in Monomeric Cobalt Nitroxides, Inorg. Chem., 2013, 52, 7557-7572. (12) Bencini, A.; Gatteschi, D. EPR of Exchange Coupled Systems, Ch. 7; Dover: New York, 1990. (13) Gruber, S. J.; Harris, C. M.; Sinn, E. Metal Complexes as Ligands. VI. Antiferromagnetic Interactions in Trinuclear Complexes Containing Similar and Disimilar Metals. J. Chem. Phys. 1968, 9, 2183-2191. (14) Kahn, O. Molecular Magnetism. Ch. 4; VCH: New York, 1993. (15) When J1 is strongly ferromagnetic and J2 is moderately ferromagnetic, as predicted from the proposed structural correlation8 and suggested from the theoretical calculation, the entropy change (ΔSLH = R ln(4×2) = 17.3 J K–1 mol–1) became closer to the experimental data. The moderately ferromagnetic coupling might be buried by the structural transition around 170 K, but the discussion on the transition mechanism indicates that the coupling is qualitatively ferromagnetic. (16) Boca, R. Theoretical Foundations of Molecular Magnetism: Current Methods in Inorganic Chemistry, Vol. 1, Elsevier: Amsterdam, 1999. (17) Sorai, M.; Seki, S. Phonon Coupled Cooperative Low-Spin 1 A1 High-Spin 5T2 transition in [Fe(phen)2(NCS)2] and [Fe(phen)2(NCSe)2] Crystals. J. Phys. Chem. Solids 1974, 35, 555570. (18) Weber, R. T. WIN-EPR SimFonia version 1.2; Bruker Instruments, Billerica, 1995. (19) Frisch, M. J. et al. Gaussian 03 revision C.02; Gaussian Inc., Wallingford, USA, 2004. (20) The SCF energies were –3264.15278798 and –3264.15257287 a.u. for the quintet and singlet states, respectively, at 400 K. Those of 85 K were –3264.14630716 and –3264.15145334 a.u., respectively. (21) Zueva, E.M.; Tretyakov, E.V.; Fokin, S.V.; Tkacheva, A.O.; Bogomyakov, A.S.; Petrova, O.V.; Trofimov, B.A.; Sagdeev, R.Z.; Ovcharenko, V.I. Stereo Sensitivity of Exchange Interactions in NiII and CuII Heterospin Complexes with 5-Formylpyrrolyl-Substituted Nitroxides. Russ. Chem. Bull. Int. Ed. 2016, 65, 666-674. (22) (a) Ovcharenko, V.; Fokin, S.; Chubakova, E.; Romanenko, G.; Bogomyakov, A.; Dobrokhotova, Z.; Lukzen, N.; Morozov, V.; Petrova, M.; Petrova, M, Zueva, E.; Rozentsvelg, I.; Rudyakova, E.; Levkovskaya, G.; Sagdeev, R. A Copper−Nitroxide Adduct Exhibiting Separate Single Crystal-to-Single Crystal Polymerization−Depolymerization and Spin Crossover Transitions. Inorg. Chem. 2016, 55, 5853-5861. (b) Fedin, M.V., Verber, S.L., Bagryanskaya, E., Ovcharenko, V.I. Electron Paramagnetic Resonance of Switchable Copper-Nitroxide-Based Molecular Magnets: An Indispensable Tool for Intriguing Systems. Coord. Chem. Rev. 2015, 289-290, 341-356. (23) (a) Kanetomo, T.; Yoshitake, T.; Ishida, T. Strongest Ferromagnetic Coupling in Designed Gadolinium(III)–Nitroxide Coordination Compounds. Inorg. Chem. 2016, 55, 8140-8146. (b) Kanetomo, T.; Kihara, T.; Miyake, A.; Matsuo, A.; Tokunaga, M.; Kindo, K.; Nojiri, H.; Ishida, T. Giant Exchange Coupling Evidenced with a Magnetization Jump at 52 T for a Gadolinium-Nitroxide Chelate. Inorg. Chem. 2017, 56, 3310-3314. (24) Zhu, M.; Li, C.; Wang, X.; Li, L.; Sutter, J.-P. Thermal Magnetic Hysteresis in a Copper−Gadolinium−Radical Chain Compound. Inorg. Chem. 2016, 55, 2676-2678. (25) (a) Letard, J.-F.; Real, J.A.; Moliner, N.; Gaspar, A.B.; Capes, L.; Cador, O.; Kahn, O. Light Induced Excited Pair Spin State in an Iron(II) Binuclear Spin-Crossover Compound. J. Am. Chem. Soc. 1999, 121, 10630-10631. (b) Adams, D. M.; Dei, A.; Rheingold, A.L.; Hendrickson, D.N. Bistability in the [CoII(semiquinonate)2] to [CoIII(catecholate)(semiquinonate)] Valence-Tautomeric Conversion. J. Am. Chem. Soc. 1993, 115, 8221-8229. (26) The term “spin-crossover” can be expanded to multi-centered systems.22a It may be acceptable from viewing a common basis; the spin-state is regulated from the balance of ferro- and antiferromagnetic contributions which originate in Hund’s rule and the Aufbau principle, respectively, in a molecular orbital picture.

ACS Paragon Plus Environment

Page 5 of 5 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

SYNOPSIS TOC Compound [Ni(phpyNO)2Cl2] was prepared as a pseudo-symmetrical 2p-3d-2p heterospin molecule. Structural distortion was found in the variable-temperature X-ray structure analysis. The magnetic study revealed that the nickel-radical exchange couplings changed from ferro- to antiferromagnetic ones on cooling, affording an equilibrium between the Stotal = 2 and 0 states. The complex is the first non-iron(II) system to undergo an S = 0 ⇄ S = 2 spin-crossover.

ARTWORK TOC

ACS Paragon Plus Environment

5