Thermally persistent high spin ground states in octahedral iron clusters

reduction of the (HL)2Fe6 cluster results in an abrupt increase in spin ground state from an S = 6 for ..... crystals showed significant decay during ...
0 downloads 0 Views 2MB Size
Subscriber access provided by Kaohsiung Medical University

Article

Thermally persistent high spin ground states in octahedral iron clusters Raúl Hernández Sánchez, and Theodore A. Betley J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.8b10181 • Publication Date (Web): 07 Nov 2018 Downloaded from http://pubs.acs.org on November 8, 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 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 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.

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 19 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

Journal of the American Chemical Society

Thermally persistent high spin ground states in octahedral iron clusters Raúl Hernández Sánchez† and Theodore A. Betley* Department of Chemistry and Chemical Biology, Harvard University, 12 Oxford Street, Cambridge, MA 02138, USA.

ABSTRACT: Chemical oxidation and reduction of the all-ferrous (HL)2Fe6 in THF affords isostructural, coordinatively unsaturated clusters of the type [(HL)2Fe6]n: [(HL)2Fe6][BArF24] (1, n = +1; where [BArF24]– = tetrakis[(3,5-trifluoromethyl)phenyl]borate), [Bu4N][(HL)2Fe6] (2a, n = –1), [P][(HL)2Fe6] (2b, n = –1; where [P]+ = tributyl(1,3-dioxolan-2-ylmethyl)phosphonium), and [Bu4N]2[(HL)2Fe6] (3, n = –2). Each member of the redox transfer series was characterized by zero-field 57Fe Mössbauer spectroscopy, near-infrared spectroscopy, single-crystal X-ray crystallography, and magnetometry. Redox directed trends are observed when comparing the structural metrics within the [Fe6] core. The metal octahedron [Fe6] decreases marginally in volume as the molecular reduction state increases as gauged by the Fe–Feavg distance varying from 2.608(11) Å (n = +1) to 2.573(3) (n = –2). In contrast, the mean Fe–N distances and Fe–N–Fe angles correlate linearly with [Fe6] oxidation level; or alternatively, the changes observed within the local Fe–N4 coordination planes vary linearly with the aggregate spin ground state. In general, as the spin ground state (S) increases, the Fe–N(H)avg distances also increase. The structural metric perturbations within the [Fe6] core and measured spin ground states were rationalized extending the previously proposed molecular orbital diagram derived for (HL)2Fe6. Chemical reduction of the (HL)2Fe6 cluster results in an abrupt increase in spin ground state from an S = 6 for the all-ferrous cluster, to an S = 19/ in the monoanionic 2b and S = 11 for the dianionic 3. The observation of asymmetric intervalence charge transfer bands in 3 2 provides further evidence of the fully delocalized ground state observed by 57Fe Mössbauer spectroscopy for all species examined (1 – 3). For each of the clusters examined within the electron transfer series, the observed spin ground states thermally persist to 300 K. In particular, the S = 11 in dianionic 3 and S = 19/2 in the monoanionic 2b represent the highest spin ground states isolated up to room temperature known to date. The increase in spin ground state results from population of the antibonding orbital band comprised of the Fe–N σ* interactions. As such, the thermally persistent ground states arise from population of the resultant single spin manifolds in accordance with Hund's rules. The large spin ground states, indicative of strong ferromagnetic electronic alignment of the valence electrons, result from strong direct exchange electronic coupling mediated by Fe–Fe orbital overlap within the [Fe6] cores, equivalent to a strong double exchange magnetic coupling B for 3 that was calculated to be 309 cm–1.

1. INTRODUCTION Mixed valence compounds were first recognized as a unique class of chemical compounds in the late 1960s given their distinctive spectral, electrochemical, and magnetic properties.1-4 A consequence of having two or more metal sites within a complex at different formal oxidation states opens the possibility for electronic delocalization between the sites. The extent of electronic delocalization has been heavily studied in mixed valent dinuclear coordination complexes,5-8 but less so in larger molecular assemblies (e.g., polynuclear clusters) or bulk solids (e.g. minerals9). In particular, polynuclear clusters have attracted much attention given their potential utility as models of biologically relevant cofactors with regards to understanding their electronic structure.10-11 The biological mixed valent metal clusters represent a minor component of very large protein superstructures, complicating the cluster spectroscopic analysis and making rigorous electronic structure determination challenging. Nonetheless, significant strides have been taken towards the analysis of cofactor spin ground

states.12 In particular, super-spin proteins have been discovered wherein a one-electron redox reaction triggers the appearance of high spin states.13-16 Given the importance for the high spin electronic structures arising in mixed valent protein cofactors that mediate electron transfer and catalyze small molecule activation, gaining a detailed understanding of the electronic structure makeup for these systems and electronic structure analogues is warranted. Previously, well-defined polynuclear clusters could be obtained from dimensional reduction of 3D-lattices.17-19 Unfortunately the examples of dimensional reduction is limited in number and, to date, most current cluster syntheses rely on self-assembly reactions that lack control of the resulting cluster constitution, geometry, and local metal coordination environments.20-32 Thus reliable methods to assemble polynuclear clusters are required to rigorously undertake electronic structure evaluation of mixed valent clusters. Our lab and others33-48 have reported the synthesis of clusters with varying nuclearity (three metals and higher) templated within a polynucleating ligand architecture to predispose the metal ions to bind in a specific manner to form

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

the polynuclear aggregate. The resulting clusters have been utilized to explore the nature of the metal-metal interactions by varying auxiliary ligand coordination to the polynuclear system, analyzing the effect of distance between the metal ions to develop magneto-structural correlations, and by assessing electronic structure perturbations as a function of cluster redox level.49-61 Our own interests have been to illustrate the electronic coupling effect engendered by cluster designs featuring close M–M interactions. For tri- and hexanuclear iron based clusters, we have observed that the direct electronic exchange interaction resulting from the direct overlap of Fe–Fe valence orbitals results in thermally persistent high-spin ground states.61-62 Furthermore, the ferromagnetic alignment is significantly enhanced upon formation of mixed valent species through the introduction of double exchange (DE) interactions as a consequence of electron delocalization.63-64 Fundamentally the double exchange electronic coupling augments ferromagnetic coupling within the interacting sites.65-66 Stabilization of high spin states in strongly delocalized species has provided the evidence of this interaction.67-74 Thus, double exchange not only supports the high spin states (e.g. S ≥ 7/2) observed in biologically relevant polynuclear clusters,7582 but it provides a roadmap to achieving high spin ground states relevant for a number of applications. One area where high spin electronic architectures are desirable is for the pursuit of single-molecule magnets (SMMs). High spin states have been long targeted to increase the spin reversal barrier energy (U = |D|S2) exhibited by these materials.83-84 The majority of materials exhibiting SMM behavior rely on superexchange electronic coupling between paramagnetic ions, thus research has focused on building larger and larger polynuclear aggregates to increase S. High spin ground states (S = 7  83/2)85-124 have been obtained in this manner, yet many only exhibit the high spin ground states at low temperatures (< 30 K). In contrast we have reported mixed valent trinuclear (S = 11/2) and hexanuclear (S = 19/2) iron clusters which feature thermally persistent ground states (to 300 K) owing to both direct and double exchange mechanisms.125-126 We report herein the formation of a coordinatively unsaturated hexanuclear cluster [(HL)2Fe6]n for which four redox levels are accessible (n = +1  –2). We find that electronic delocalization of the mixed valent clusters is observable structurally and spectroscopically. The high spin ground state arising from strong direct exchange previously reported for (HL)2Fe6 is reinforced by the double exchange electronic coupling introduced by mixed valency. The high spin ground state electronic structures observed for two of the anionic redox isomers (S = 19/2 [(HL)2Fe6]– and S = 11 [(HL)2Fe6]2–) represent the highest, thermally persistent ground states ever recorded.

2. EXPERIMENTAL SECTION 2.1. General considerations. All manipulations involving metal complexes were performed under an atmosphere of dry, oxygen-free N2 by means of standard Schlenk or glovebox techniques (MBraun glovebox equipped with a –35 °C freezer).

Page 2 of 19

All glassware was oven-dried for a minimum of 10 h and cooled in an evacuated antechamber prior to use in the drybox. Diethyl ether and tetrahydrofuran (THF) were dried and deoxygenated on a Glass Contour System (SG Water USA, Nashua, NH) and stored over 4 Å molecular sieves (Strem) prior to use. Propylene carbonate (PC, Sure Seal) was purchased from Sigma Aldrich and brought into the N2-filled glovebox and stored over 4 Å molecular sieves (Strem) prior to use. Tetrabutylammonium chloride, [Bu4N]Cl; tributyl(1,3dioxolan-2-ylmethyl)phosphonium bromide (see Scheme 1), [P]Br; and ferrocenium hexafluorophosphate, Fc[PF6], were purchased from Sigma Aldrich and used without further purification. Likewise, naphthalene obtained from Alfa Aesar was used without further purification. Sodium tetrakis[3,5bis(trifluoromethyl)phenyl]borate, Na[BArF24], was purchased from TCI America and used without further purification. Salt metathesis between Fc[PF6] and Na[BArF24] to afford large crystals of Fc[BArF24] was carried out inside a dinitrogen filled glovebox according to the procedure reported by Chávez et al.127 (HL)2Fe6 was prepared according to the method previously reported by our laboratory.61 2.1.1. [(HL)2Fe6][BArF24] (1a). (HL)2Fe6 (204 mg, 0.163 mmol) was suspended in 4 mL of THF. To this was added a solution of Fc[BArF24] (172 mg, 0.164 mmol) in 2 mL of THF. The reaction mixture was stirred for 3 h at room temperature. The solvent was removed under vacuum and the resulting black powder washed in a fritted funnel with 5 x 2 mL of a 1:1 mixture of THF:Et2O. Addition of small amounts of THF to the fritted funnel produces a black filtrate solution and leaves behind a black powder. 57Fe Mössbauer on both filtrate and precipitate reveal the same species and the fractions were combined (80%). Single crystals of 1 can be obtained from a concentrated solution of THF:Et2O at –35 °C. Anal. Calcd. for 1a·2THF: C86H76BF24Fe6N12O2: C, 48.92; H, 3.63; N, 7.96%. Found: C, 48.84; H, 3.58; N, 8.09%. 2.1.2. [(HL)2Fe6(py)2][PF6] (1b). (HL)2Fe6 (105 mg, 0.084 mmol) was suspended in 2 mL of pyridine. To this was added a solution of Fc[PF6] (88 mg, 0.085 mmol) in 2 mL of pyridine. The reaction mixture was stirred for 3 h at room temperature. The solvent was removed under vacuum and the resulting black powder washed in a fritted funnel with 5 x 2 mL of a 1:1 mixture of THF:Et2O. From the frit 87 mg of black powder (1b) was isolated (74% yield). Single crystals of 1b can be readily grown by dissolving 1b in pyridine and placing this in a diethyl ether vapor diffusion cell. Anal. Calcd. for 1: C56H58F6Fe6N14P: C, 47.80; H, 4.15; N, 13.94%. Found: C, 47.66; H, 4.09; N, 13.85%. 2.1.3. [Bu4N][(HL)2Fe6] (2a). (HL)2Fe6 (100 mg, 0.080 mmol) was suspended in 2 mL of THF and cooled down to –35 °C. To this was added a freshly prepared solution of sodium naphthalenide by smearing sodium metal to the walls of a vial and adding naphthalene (10.3 mg, 0.080 mmol) in 2 mL of THF and stirred for 30 min at room temperature. The sodium naphthalenide solution was filtered through Celite in a pipette filter and added to the THF suspension of (HL)2Fe6 at –35 °C. Immediately following the sodium naphthalenide addition, [Bu4N]Cl (22.5 mg, 0.081 mmol) was added as a solid and the reaction mixture stirred for 3 h at room temperature. A dark brown precipitate was formed and filtered through a medium

ACS Paragon Plus Environment

Page 3 of 19 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

Journal of the American Chemical Society

size fritted funnel and washed with 5 x 2 mL of THF. The collected brown powder was dissolved on the frit with propylene carbonate (PC) and collected directly into a diethyl ether vapor diffusion cell. Micron-sized crystals were obtained at room temperature from which the single-crystal structure was obtained. When the PC filtrate is dried in vacuo to remove volatiles, 29 mg of 2a was obtained (27% yield). Anal. Calcd. for 2a: C62H84Fe6N13: C, 55.30; H, 6.29; N, 13.52%. Found: C, 55.21; H, 6.18; N, 13.39%. 2.1.4. [P][(HL)2Fe6] (2b). The synthesis of 2b is analogous to that of 2a. (HL)2Fe6 (100 mg, 0.080 mmol) was suspended in 2 mL of THF and cooled down to –35 °C. To this was added a freshly prepared solution of sodium naphthalenide by smearing sodium metal to the walls of a vial and adding naphthalene (10.3 mg, 0.080 mmol) in 2 mL of THF and stirred for 30 min at room temperature. The sodium naphthalenide solution was filtered through Celite in a pipette filter and added to the THF suspension of (HL)2Fe6 at –35 °C. Immediately following the sodium naphthalenide addition, tributyl(1,3-dioxolan-2ylmethyl)phosphonium bromide ([P]Br, 29.6 mg, 0.080 mmol) was added as a solid and the reaction mixture was stirred for 3 h at room temperature. In contrast to the synthesis of 2a, no precipitate formed; thus, THF and volatiles from the reaction mixture were removed in vacuo. The resulting solid was placed on a medium porosity fritted funnel and washed with 5 x 2 mL of a 1:1 THF:Et2O mixture. After washing the solid, THF was used to dissolve the powder which was filtered directly into a vapor diffusion cell with diethyl ether. High-quality, bulk crystallization results at room temperature. Yield: 56 mg (50%). 2.1.5. [Bu4N]2[(HL)2Fe6] (3). (HL)2Fe6 (100 mg, 0.080 mmol) was suspended in 2 mL of THF and cooled down to –35 °C. To this was added a freshly prepared solution of sodium naphthalenide by smearing sodium metal to the walls of a vial and adding naphthalene (20.6 mg, 0.164 mmol) in 2 mL of THF and stirred for 30 min at room temperature. The sodium naphthalenide solution was filtered through Celite in a pipette filter and added to the THF suspension of (HL)2Fe6 at –35 °C. Immediately following the sodium naphthalenide addition,[Bu4N]Cl (43 mg, 0.162 mmol) was added as a solid and the reaction mixture was stirred for 3 h at room temperature. The dark brown precipitate was filtered through a medium size fritted funnel and washed with 5 x 2 mL of THF. The remaining solids were dissolved in pyridine and collected into a diethyl ether vapor diffusion cell. High quality crystals were obtained (49% yield) in this manner at –35 °C from which the X-ray single crystal structure was obtained. Compound 3 decomposes over a period of 12 h at room temperature and under inert atmosphere. Anal. Calcd. for 3: C78H120Fe6N14: C, 58.96; H, 7.61; N, 12.34%. Found: C, 58.88; H, 7.46; N, 12.18%. 2.2. X-ray Structure Determinations. Single crystals suitable for X-ray structure analysis were coated with deoxygenated Paratone N-oil and mounted in MiTeGen Kapton loops (polyimide). Data for compounds 1 – 3 were collected at 100 K on an APEX II DUO single-crystal diffractometer. None of the crystals showed significant decay during data collection. Raw data were integrated and corrected for Lorentz and polarization effects using Bruker APEX2 v.2009.1.128

Absorption corrections were applied using SADABS.129 Space group assignments were determined by examination of systematic absences, E-statistics, and successive refinement of the structures. The program PLATON130 was employed to confirm the absence of higher symmetry for any of the crystals. The positions of the heavy atoms were determined using direct methods using the program SHELXTL.131 Successive cycles of least-square refinement followed by difference Fourier syntheses revealed the positions of the remaining nonhydrogen atoms. Non-hydrogen atoms were refined with anisotropic displacement parameters, and hydrogen atoms were added in idealized positions. Crystallographic data for 1 – 3 are given in Table S1. 2.3. Magnetic Data Measurements. Magnetic data for 1a, 1b, 2b, and 3 were collected using a Quantum Design MPMS-XL Evercool SQUID Magnetometer. The following is a general procedure for sample preparation: bulk crystals were collected and washed thoroughly with Et2O. These crystals were crushed in the presence of Et2O and the resulting fine suspension was then dried under high vacuum. The sample powder was then immobilized within a size #4 gelatin capsule by adding melted eicosane at 50 – 60 °C. The gelatin capsule was inserted into a plastic straw. Samples were prepared under a dinitrogen atmosphere. Magnetization data at 100 K from 0 to 7 T were used to detect ferromagnetic impurities (Figure S10 – S13). Variable-temperature direct current (dc) magnetic susceptibility measurements were collected in the temperature range 1.8 – 300 K under applied fields of 0.1, 0.5, and 1 T. The magnetic susceptibility (MT) were collected repeatedly until they were reproduced at least three times where samples feature linear 100 K magnetization scans. Variable-temperature, variable-field magnetization data were acquired on heating from 1.8 to 10 K at increasing fields of 1, 2, 3, 4, 5, 6, and 7 T. Magnetic susceptibility data were corrected for diamagnetism of the sample, estimated using Pascal’s constants, in addition to contributions from the sample holder and eicosane. AC magnetic susceptibility data were collected at zero applied dc field and with a 4 Oe oscillating ac field. Additionally, magnetic hysteresis data were collected at 1.8 K; in addition, zero-field- (ZFC) and field-cooled (FC) data were obtained between 1.8 and 10 K by measuring under a dc applied field of 0.1 T. 2.4. Electrochemical Measurements. Cyclic voltammetry measurements were acquired on a CHI660d potentiostat. A three-electrode cell setup was used with a 3mm in diameter glassy carbon working electrode, Pt wire as counter electrode, and Ag/AgNO3 as reference electrode. Saturated AgNO3 solutions in MeCN for the reference electrode were prepared fresh before each experiment. All measurements were done under a dinitrogen atmosphere and at room temperature. A 0.1 M solution of [Bu4N][PF6] in THF was used as the supporting electrolyte. 2.5. Zero-Field 57Fe Mössbauer Spectroscopy. Zero-field 57Fe Mössbauer Spectroscopy was collected at 4.2 and 90 K for all compounds (1a, 1b, 2a, and 3). Solid samples (ca. 20 mg) were restrained with Paratone-N oil. The data were measured with a constant acceleration spectrometer (SEE Co., Minneapolis, MN). Isomer shifts are given relative to -Fe metal at 298 K.

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

The data of 1a and 3 were fit by employing WMOSS4.132 Additionally the 90 K data of 1a, 1b, and 3 were fit using an inhouse software written in IGOR Pro WaveMetrics by Evan King.

Page 4 of 19

[(HL)2Fe6]+ readily binds two trans-disposed equivalents of pyridine to generate [(HL)2Fe6(py)2]+ (cation of 1b, see Scheme S1).

2.6. Other Physical Measurements. Elemental analysis was performed by Complete Analysis Laboratories, Inc., Parsippany, New Jersey. UV-Vis-NIR spectra were collected in 1 cm path length cuvettes on a Varian 5000 spectrophotometer at room temperature. All solutions were prepared under N2 atmosphere in a glovebox and the cuvettes sealed with a J-Young screw Teflon cap. Absorbance values were kept under 1 for all concentrations measured.

3. RESULTS 3.1 Synthesis and Electrochemical Characterization. We have previously reported on the redox isomers of (HL)2Fe6 and the nature of redox delocalization of the mixed valent clusters.61-62 Upon oxidation, the [Fe6] cluster binds two equivalents of solvent (MeCN or DMF) to trans-disposed iron sites within the cluster for every one-electron oxidation incurred until the cluster becomes coordinatively saturated at the tricationic state [(HL)2Fe6(L’)6]3+ (L’ = MeCN or DMF). We were thus interested to examine the nature of redox delocalization within the core in the absence of solvent or anion binding to the [Fe6] core. Scheme 1. Synthesis of unsolvated [(HL)2Fe6]n clusters

To prevent solvent ligation, oxidation of (HL)2Fe6 in THF was effected using Fc[BArF24] ([BArF24]– = tetrakis[(3,5trifluoromethyl)phenyl]borate) to render the oxidized product THF-soluble. Thus, one equivalent of Fc[BArF24] in THF was added to a suspension of (HL)2Fe6 in THF at room temperature and stirred for 3 h. Removal of the volatiles in vacuo followed by washing with 1:1 THF:Et2O affords [(HL)2Fe6][BArF24] (1a) in 80% yield (Scheme 1). Micron size single crystals can be grown from a standing THF:Et2O concentrated solution of 1a at –35 °C. Akin to the solvent ligation pattern observed during oxidation of (HL)2Fe6 in MeCN,61 the unsolvated cluster core

Figure 1. (a) Molecular crystal structure of the dianion [(HL)2Fe6]2– in 3 obtained at 100 K (the ammonium ion [Bu4N]+ and the hydrogen atoms are omitted for clarity). Thermal ellipsoids are set at 50% probability level. The Fe, C and N atoms are colored orange, grey and blue, respectively. (b) Cyclic voltammogram of [P][(HL)2Fe6] (2b) in THF at room temperature. Scan rate: 10 mV/s. A 0.1 M [Bu4N][PF6] solution was used as supporting electrolyte. The labels on italics correspond to the oxidation level n in [(HL)2Fe6]n.

ACS Paragon Plus Environment

Page 5 of 19 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

Journal of the American Chemical Society

Table 1. Structural metrics and spin ground state for unsolvated complexes 1a, 2a, 2b, and 3 n

+1 1a

(HL)

0

–1

–1

–2

a 2Fe6

2a

2b

3

Fe–Fecis (Å)

2.608(11)

2.597(1)

2.599(11)

2.597(9)

2.573(6)

Fe–N(H)avg (Å)

1.988(5)

2.01(2)

2.081(13)

2.078(10)

2.127(9)

Fe–N(H)–Feavg (deg)

82.0(6)

80.1(9)

77.3(8)

77.4(6)

74.4(1)

3.6044(4)

3.610(27) 3.651(13)

3.630(1)

3.7111(4) 3.7135(5)

3.757(9)

3.646(1)

---

19/

11

3.63(2) Fe–Fetrans (Å)b

3.66(1)

3.673(2)

3.78(5) S aPreviously

11/

2

6

2

3.643(1)

reported.61 bIndividual values reported to highlight the [Fe6] core asymmetry.

We previously reported that chemical reduction of a suspension of (HL)2Fe6 in THF with sodium naphthalenide will afford the monoanionic, coordinatively unsaturated cluster [Na(OEt2)2(NCMe)2][(HL)2Fe6].61 Alternatively, the addition of [Bu4N]Cl to the reaction mixture immediately following chemical reduction (3 h, room temperature) yields the cation metathesis product [Bu4N][(HL)2Fe6] (2a, Scheme 1). Initial electrochemical examination of 2a was challenging owing to the poor-to-negligible solubility of 2a in THF (or any other suitable solvent for electrochemical studies). Thus, other cations were examined to solubilize [(HL)2Fe6]–. We found that by performing the salt metathesis with tributyl(1,3dioxolan-2-ylmethyl)phosphonium bromide ([P]Br) after one-electron reduction of (HL)2Fe6 in the same manner 2a was prepared, results in the THF-soluble compound [P][(HL)2Fe6] (2b) in 50% yield. Bulk crystallization of 2b can be carried out by diffusing Et2O into a concentrated THF solution at room temperature. The electrochemical properties of the unsolvated [Fe6] core were investigated by employing the THF-soluble species 2b. Cyclic voltammetry in THF reveals one reduction event, as expected to result from reduction to the dianionic cluster core [(HL)2Fe6]2– (E1/2 = –2.41 V vs Fc0/+, EP = 196 mV); and two oxidation events corresponding to the formation of the neutral all-ferrous and monocationic [(HL)2Fe6]+ core (E1/2 = – 1.68 and –1.18 V vs Fc0/+; EP = 158 and 192 mV, respectively). The calculated comproportionation constant of the mixed valent unsolvated [Fe6] core at the oxidation level of 2b is thus 2.9 x 1012; which is slightly larger than that found for the pyridine species [(HL)2Fe6(py)2]– (1.3 x 1010).125 The large comproportionation constants obtained for both species indicates strong electron delocalization within the hexanuclear iron core.62 The cyclic voltammogram of 2b showed a reversible reduction indicating a doubly reduced species was accessible. The doubly reduced cluster [Bu4N]2[(HL)2Fe6] (3) could be obtained in 49% yield following the same synthetic procedure described above for 2a using two equivalents of

sodium naphthalenide. Compound 3 is completely insoluble in THF, but readily soluble in pyridine. Crystals of 3 suitable for X-ray analysis can be grown by diffusing Et2O into a concentrated pyridine solution of 3 at –35 °C. The doubly reduced species 3 could be maintained as a crystalline powder at –35 °C, though it readily decomposes upon standing at room temperature in contrast to the thermally stable 2a and 2b. The X-ray molecular crystal structure of the clusters in compounds 1a, 2a, 2b, and 3 resemble that of the parent allferrous (HL)2Fe6, where all the iron sites remain locally 4coordinate square planar (neglecting Fe–Fe interactions, see Figure 1a). Given their isostructural nature a direct comparison of the bond metrics of the all-ferrous parent cluster and 1a, 2a, 2b, and 3 reveals significant perturbations on the bond lengths involving Fe–N(H)avg bonds and Fe– N(H)–Feavg angles; conversely, the Fe–Feavg distances change minimally. The volume of the M6 octahedron decreases marginally with cis-disposed Fe–Feavg distances of 2.597(1) (n = 0 (HL)2Fe6), 2.608(11) (n = +1, 1a), 2.599(11) (n = –1, 2a), and 2.573(6) Å (n = –2, 3) (data summarized in Table 1). Given the almost identical parameters within the [(HL)2Fe6]– core fragment in 2a and 2b, only the metrics of 2a will be discussed. Moreover, the Fe–N and Fe–NH distances increase linearly with the increasing overall cluster reduction level: 2.006(4) and 1.969(4) Å (n = +1, 1); 2.037(8) and 1.998(17) Å (n = 0, (HL)2Fe6); 2.099(17) and 2.063(21) Å (n = –1, 2a); and 2.144(8) and 2.110(15) Å (n = –2, 3). In a similar fashion, the Fe–N–Fe and Fe–NH–Fe angles become more acute with increasing core reduction level [Fe–N–Fe, Fe–NH–Fe (°)]: 80.2(5), 83.9(4) (n = +1, 1); 79.2(2), 81.1(2) (n = 0, (HL)2Fe6); 76.8(11), 77.9(12) (n = –1, 2a); and 74.2(1), 74.6(1) (n = –2, 3). 3.2. Zero-Field 57Fe Mössbauer and Near-IR Spectroscopy. The Mössbauer spectrum of the hexanuclear all-ferrous cluster (HL)2Fe6, in addition to the electron transfer families

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

of compounds: [(HL)2Fe6(NCMe)m]n+ (n = 1  4), [(HL)2Fe6(DMF)m]n+ (n = 2  4), and [(HL)2Fe6(CN)6]n–6 (n = 3  6) displayed a single quadrupole doublet for geometrically equivalent coordination sites,61-62 indicative of a delocalized ground state. One exception to this general observation was the singly reduced [Na(Et2O)2(NCMe)2][(HL)2Fe6], whose spectrum featured an asymmetric doublet at 90 K.61, 125 The spectral asymmetry was attributed to a proposed mixture of neutral (HL)2Fe6 and reduced [Na(Et2O)2(NCMe)2][(HL)2Fe6] clusters, owing to the presumed instability of the reduced [Fe6]–. We later ascertained the spectral asymmetry was due to slow magnetic relaxation in related monoanionic clusters [(HL)2Fe6(L’)2]– that is not fully resolved at 90 K,125 which has also been observed on other iron-based SMMs.133-134 Analysis of the Mössbauer spectrum of 1a and 2a gave insight into concluding that its asymmetry is caused by the slow relaxation of the magnetization imparted by the ground state, |𝑆,𝑚𝑠⟩.51

Page 6 of 19

electronically delocalized species.137 The intensity (max), position (max), full-width at half-maximum (1/2), and shape of this band is related to the extent of delocalization.1-4 The corresponding data for [Bu4N][(HL)2Fe6(py)2] display no absorption bands in the near-IR region of the spectrum where it is customary to find the IVCT bands.125 In the same way, electronic absorption data were collected for 1a, 2b, and 3. In the visible

Zero-field, 57Fe Mössbauer spectra were collected for 1a, 1b, 2a, and 3 at 4.2 and 90 K; an additional spectrum at 240 K was also collected for 2b. The spectrum of the cationic cluster [(HL)2Fe6][BArF24] (1a) at 4.2 K displays a hyperfine split spectrum (Figure 2a). Based on the almost identical heights of all absorption lines we presumed that this spectrum is composed by several individual sextets. Similar to 1a, compound 2a displays a hyperfine split spectrum at 4.2 K; although the spectral features are significantly broader than that of 1a (Figure 2b). The spectrum of 3 at 4.2 K (and 90 K) displays a single broad quadrupole doublet (Figure 2c). This is remarkable since the analogous unsolvated salts at n = +1 and –1 (clusters in 1a and 2a) display well-defined-toincipient hyperfine split spectra, respectively. The fit parameters that best model the spectrum of 3 at 4.2 K are [, |EQ| (mm/s)]: 0.62, 1.44. The zero–field spectrum of 1a was modeled by employing a nuclear Hamiltonian including, apart from the electric quadrupole interaction, the nuclear Zeeman interaction, 𝐻 = I ∙ 𝑄 ∙ I + 𝑔𝑛𝛽𝑛H ∙ I;135 where 𝑄 is proportional to the electric field gradient. Note that the only contribution to H originates from the internal magnetic field (Hint) since Hext = 0. The data were fit by employing the software WMOSS4 model 5.132 Three distinct crystallographic sites are observed per [Fe6] thus three sextets were considered when fitting the spectrum of 1a. The fit components are shown above the experimental data of 1a in Figure 2a and have the following parameters [, EQ (mm/s), Hint (T)]: 0.83, –2.45, 9.6 (33%, blue trace); 0.83, –2.45, 2.8 (33%, green trace); and 0.37, – 2.45, 6.8 (34%, brown trace). The composite fit is depicted by the red trace. On warming the relaxation rate of Hint increases and at 90 K averages to zero (Figure S6).136 At this temperature a single quadrupole doublet is observed that is best fit by [, |EQ| (mm/s)]: 0.40, 2.45 (Figure S6). The breadth of the spectrum of 2a precluded further analysis (Figure 2b). However, as temperature is increased to 90 and 240 K the spectrum coalesces into a single quadrupole doublet (Figure S8). A common feature for mixed valence compounds is the appearance of intervalence charge transfer bands (IVCTs) for

Figure 2. Zero-field 57Fe Mössbauer spectra for: (a) [(HL)2Fe6][BArF24] (1a); (b) [Bu4N][(HL)2Fe6] (2a); and (c) [Bu4N]2[(HL)2Fe6] (3) at 4.2 K. The blue, green and brown tarces correspond to the individual components of the fit for 1. The red traces indicate the overall fit to the data as described in the text. (d) Near-IR spectrum of 3 in pyridine at room temperature. Inset: UV–Vis part of the spectrum. The red trace corresponds to the fit of the spectrum considering three gaussian absorption bands (dashed and dotted grey, and green trace) between 7400 and 17500 cm–1.

ACS Paragon Plus Environment

Page 7 of 19 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

Journal of the American Chemical Society

region 1a displays two broad absorption bands at max = 24907 and 19608 cm–1 (Figure S14), while no other discernible peak is observed down to 6000 cm–1. The absorption spectrum of 2b displays one peak maximum in the UV region at 33900 cm–1 and a second one in the visible

at 23900 cm–1 (Figure S16). Similar to 1a, no absorption band is observed in the near-infrared for 2b. For 3 only one broad absorption is seen in the visible region at 23200 cm–1 (Figure 2d inset). In contrast, in the near-IR two

Figure 3. Magnetic data collected for unsolvated complexes [(HL)2Fe6][BArF24] (1a), [P][(HL)2Fe6] (2b), and [Bu4N]2[(HL)2Fe6] (3). (a) VT magnetic susceptibility of 1a (), 2b (), and 3 () collected under a 0.1 T field. (b) Reduced magnetization data of 1a (), 2b (), and 3 () at temperatures from 1.8 to 10 K and fields 1 to 7 T. The continuous black traces in (a) and (b) correspond to the fit of the data as described in the text. (c) Plot of 𝜒′MT versus T of 1a (), 2b (), and 3 () at an oscillating field frequency of 1 Hz.

bands are observed between 13000 and 6000 cm–1. The low energy absorption band is asymmetric and skewed as predicted for moderately-to-strongly delocalized complexes.138 Thus the absorption profile fit is limited to max of the lowest absorption band; while the high energy side of the IVCT retains a Gaussian-shape, the low energy side does not.139 To gain further insight, the data between 7400 and 17500 cm–1 were fit to three Gaussian-shaped curves, where two (high and intermediate energy, dotted and dashed traces) are used only to provide an adequate absorption profile for the lowest energy band (Figure 2d). The fit parameters of the lowest absorption band (Figure 2d green trace) are /max = 0.1253 mol–1L ( = 890 mol–1cm–1L), max = 7107 cm–1, and 1/2 = 3911 cm–1. As can be observed the Gaussian fit overestimates the real width of the absorption band. To estimate the experimental 1/2, the difference between the high and low energy values of the fit and experimental data, respectively, was taken at half the maximum intensity, giving an estimate of 2696 cm–1. The extent of electron delocalization can be quantified by calculating the electronic coupling (Hab) between the donor and acceptor ground state energy surfaces using the spectral deconvolution results. Two methods were employed to assess the extent of electron delocalization: 1) electronic coupling calculated from the Hush formula (Hab, considering the donor-acceptor distance rab as the average Fe–Fe distance),1-4 and 2) a treatment introducing the parameter  that classifies the mixed valent species into: weakly coupled class II (0 <  < 0.1), moderately coupled class II (0.1 <  < 0.5), borderline class II-III (  0.5), and class III ( > 0.5).138 Hab and  can be calculated from Eq. S1 and S2 (See Supplementary Information). Thus, we obtain from the Hush formula Hab = 1045.4 cm–1; and alternatively, from the second method  = 0.33.

3.3. Spin Ground State Analysis via Magnetometry. To investigate the nature of the spin ground state in compounds 1a, 1b, 2b, and 3, magnetic susceptibility data (𝜒MT, 1.8 – 300 K) and variable-temperature, variable-field (VTVH) magnetization data (1.8 – 10 K) were collected. The ground state of the all-ferrous species (HL)2Fe6 has been determined to be a well-isolated S = 6 that persists to room temperature.61, 125 Its magnetic anisotropy is dictated by an axial zero-field splitting parameter (D) of +1 cm–1 and intermediate rhombicity of |E/D| = 0.16 (giso = 2.004) extracted from VTVH magnetization data. To date, the clusters prepared in this laboratory featured the highest, thermally persistent ground state electronic structures yet reported (S = 19/2, [Bu4N][(HL)2Fe6(L’)2] (L’ = pyridine or DMF;125 S = 6, (HL)2Fe661, (PhL)Fe3(thf)353 and (tbsL)Fe3(thf)60, 126). Additionally, ac magnetic susceptibility data for[Bu4N][(HL)2Fe6(L’)2] (L’ = py or DMF) reveal singlemolecule magnet (SMM) behavior at low temperatures for the anions with magnetic blocking temperatures (TB) of 2.6 (L’ = py) and 2.1 K (L’ = DMF) and effective spin reversal barriers (Ueff) of 42.5(8) and 33.5(1) cm–1, respectively.125 Magnetic data collected on [(HL)2Fe6][BArF24] (1a) display a plateau above 20 K (𝜒MT = 16.0 cm3K/mol) that persists to room temperature (𝜒MT300 K = 18.1 cm3K/mol). The room temperature 𝜒MT is slightly above the predicted spin-only value for an S = 11/2 (17.87 cm3K/mol, g = 2.0). In contrast, on decreasing temperature to 1.8 K a sudden decrease to 𝜒MT = 5.2 cm3K/mol is observed and is likely the result of zero-field splitting (Figure 3a). While oxidation of the all-ferrous cluster (HL)2Fe6 (S = 6) decreases the spin ground state of the [Fe6] core, reduction of (HL)2Fe6 increases the spin ground state significantly as observed in 2b (Figure 3a). VT 𝜒MT data of 2b display a plateau from 40 (𝜒MT = 49.5 cm3K/mol) to 300 K (𝜒MT = 48.9

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

cm3K/mol); in addition to a precipitous drop to 2.2 cm3K/mol at 1.8 K. This magnetic behavior is rather similar to the previously reported S = 19/2 clusters [Bu4N][(HL)2Fe6(L’)2] (L’ = py, dmf).125 The doubly reduced cluster [(HL)2Fe6]2– displayed

Page 8 of 19

even higher 𝜒MT values. VT 𝜒MT data of [Bu4N]2[(HL)2Fe6] (3) display a room temperature susceptibility of 𝜒MT300 K = 65.2 cm3K/mol (Figure 3a), remarkably close to the spin only value

Figure 4. Alternating-current magnetic susceptibility data for [(HL)2Fe6][BArF24] (1a, a-c) and [P][(HL)2Fe6] (2b, d-f) under zero-applied external dc field. Out-of-phase (𝜒′′M, a and d) and in-phase (𝜒′M, c and f) components of the ac magnetic susceptibility versus frequency (𝜈). (b and e) Cole-Cole plots. The solid continuous lines (a-f) represent a fit to the data as described in the text. (g) Relaxation times (ln ) versus inverse temperature (1/T) obtained by fitting the data in (a-f) to an Arrhenius temperature law as described in the text. Temperature labels in (a) are applicable to (a-c); similarly, those in (d) are applicable to (d-f). (h) Magnetic hysteresis of 2b at 1.8 K cycling from +7  –7 T. Zoomed-in region (+1  –1 T) rate is 0.18 mT/s. Inset: zero-field-cooled (ZFC) and field-cooled (FC) data for 2b at a temperature sweeping rate of 0.057 K/min.

predicted for an S = 11 (66 cm3K/mol, g = 2.0). The susceptibility increases to 68.2 cm3K/mol on decreasing the temperature to 80 K. Below 3.5 K the effect of zero-field splitting is manifest and a sudden decrease in 𝜒MT is observed to 55.6 cm3K/mol at 1.8 K. The 𝜒MT data of 1a, 2b and 3 were fit using PHI140 according 2 to the spin Hamiltonian 𝐻 = 𝐷𝑆𝑧 + 𝑔𝑖𝑠𝑜𝜇𝐵S ∙ H, for an S = 11/2 (1a), 19/2 (2b), and 11 (3). The fit parameters that best reproduced the data of 1a are: g = 1.91, D = –0.01 cm–1, and TIP = 5 x 10–3 cm3/mol. Similarly, for 2b and 3 the fits are: g = 2.04 and 2.03; and D = +10.9 and ~0 cm–1, respectively. Further insight into the ground states of these clusters was obtained by collecting VTVH magnetization data. In all cases the reduced magnetization data display non-superimposable isofield curves indicative of zero-field splitting (ZFS); although for 3 this appears to be negligible (Figure 3b). At 1.8 K and under a 7 T magnetic field saturation of magnetization in 1a, 2b, and 3 occurs at 8.1, 14.9, and 22.0 B, respectively (Figure 3b). Saturation in 3 occurs at the expected value for an S = 11

when ZFS is absent; in contrast the magnetic saturation for 1a and 2b fall short of the expected values of 11 and 19 B, respectively. To quantify this effect more reliably the data 2 2 2 were fit to the spin Hamiltonian 𝐻 = 𝐷𝑆𝑧 +𝐸(𝑆𝑥 ― 𝑆𝑦) + 𝑔𝑖𝑠𝑜𝜇𝐵S ∙ H, considering the ground states established. The data of 1a is best modeled by an S = 11/2 with g = 1.87, D = – 1.13 cm–1, and |E/D| ≈ 0. Comparably high-quality fits were obtained for 2b and 3 resulting in the following parameters: g = 2.04 and 2.003; D = –0.76 and +0.07 cm–1; and |E/D| = 0.04 and 0.07, respectively. As a final note the ground state assignment of 1a, 2b, and 3 was complemented by analyzing the 𝜒′MT vs. T dependence. When dealing with large spin ground states, low lying excited states have the potential to be populated under the traditional dc magnetic susceptibility measurement. To circumvent this complication, the susceptibility is recorded using a small ac field in place of the dc measurement.109, 121 The ac measurement resembles the measurement acquired under a static dc field as long as the oscillation frequency of

ACS Paragon Plus Environment

Page 9 of 19 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

Journal of the American Chemical Society

the ac field is slow enough to allow the magnetization vector to reorient. At 1.8 K 1a displays a 𝜒′MT value of 14.3 cm3K/mol (Figure 3c). On slight heating it reaches a plateau at T > 2.4 K (16.8 cm3K/mol). Further heating increases 𝜒′MT subtly to approach the spin-only value of an S = 11/2 (17.87 cm3K/mol; 𝜒′MT10 K = 17.77 cm3K/mol). For 2b a marked transition is observed where at low temperatures 𝜒′MT increase mildly from 1.68 to 3.59 cm3K/mol, from 1.8 to 3 K, respectively; followed by a sudden increase in 𝜒′MT which occurs between 3 and 4.4 K. For T > 4.4 K a monotonic increase brings the moment up and by T = 10 K the recorded 𝜒′MT is 37.6 cm3K/mol. The closely related S = 19/2 solvated species [Bu4N][(HL)2Fe6(L’)2] displays similar behavior.125 Of greater significance, the downturn in the susceptibility product corresponds to a decrease in the in-phase susceptibility (𝜒′M), which marks the occurrence of slow relaxation of the magnetization as has been described elsewhere.141 In contrast, species 3 displays at the lowest temperature a moment that corresponds already with population of an S = 11 (𝜒′MT1.8 K = 62.7 cm3K/mol, Figure 3c). 3.4. Slow Relaxation of the Magnetization. The relaxation dynamics of 1a, 2b, and 3 were probed by ac magnetic susceptibility at zero dc-field. The frequencies of the 4 Oe oscillating field were in the range of 1 to 1488 Hz. To our surprise 3 does not display an out-of-phase (𝜒′′M) component indicative of SMM behavior. Therefore only [(HL)2Fe6][BArF24] (1a) and [P][(HL)2Fe6] (2b) were investigated in detail. At 1.9 K a peak in the 𝜒′′M vs.  plot is evident for 1a which shifts to higher frequencies as the temperature is increased (Figure 4a). An equivalent behavior for 2b is observed at T ≥ 3.8 K (Figure 4d). Accordingly, a decrease in 𝜒′M is observed for both species (Figure 4c and f).142 The SMM behavior can also be visualized by the semicircular profiles displayed in the Cole-Cole plots (𝜒′′M vs. 𝜒′M) illustrated in Figure 4b and e; where a single relaxation pathway is invoked judging from the appearance of a single arc at any given temperature.143 The relaxation dynamics were fit according to a generalized Debye model.144 IGOR Pro Wavemetrics was used to fit 𝜒′′M and 𝜒′M vs. 𝜈 to a distribution of single relaxation processes.145-146 CC-FIT was employed to fit the Cole-Cole plots.147 Panels (a) through (f) in Figure 4 were fit independently and the extracted temperature-dependent relaxation times () fit to an Arrhenius temperature law 𝜏 = 𝜏0exp(𝑈eff/kBT). The average values from all the three methods fit to an effective spin relaxation barrier of Ueff = 21.7(1) and 43.8(5) cm–1 and an attempt time of 𝜏0 = 8.7(6) x 10–9 and 6(1) x 10–9 s for 1a and 2b, respectively (Figure 4g). Based on the definition of Gatteschi et al. of the magnetic blocking temperature (TB) as that at which the relaxation rate (𝜏) becomes 100 s,142 we calculate a TB of 1.3 and 2.7 K for 1a and 2b, respectively. Further evidence of magnetization blocking was manifest from magnetic hysteresis experiments. Saturation of the magnetization at 1.8 K was observed at ±8.16 and ±14.9 µB when cycling between ±7 T for 1a and 2b, respectively. As expected from the calculated blocking temperature,

compound 1a did not exhibit any magnetic hysteresis (Figure S18); in contrast 2b displayed a remnant magnetization of 4.9 µB and a coercive field of ~0.7 T (Figure 4h). In addition to the magnetic hysteresis experiments, zero-field-cooled (ZFC) and field-cooled (FC) data were collected at 0.1 T to explore the magnetization dynamics. Divergence between the ZFC and FC data were observed only for 2b, while the corresponding data for 1a superimposes at all temperatures (Figure S19). In 2b overlapping magnetization is observed at T > 2.9. The FC data display a plateau between 1.8 and 3 K at ~1.25 µB; in comparison the ZFC data diverge dramatically displaying a maximum of 1.22 µB at 2.9 K to subsequently drop to 0.21 µB at 1.8 K (Figure 5h inset).

4. DISCUSSION 4.1. Synthesis and structure. Oxidation of the all-ferrous cluster (HL)2Fe6 affords the monocationic, unsolvated cluster [(HL)2Fe6][BArF24] (1a), while isostructural species are also obtained under reduction conditions to yield the monoanionic [Bu4N][(HL)2Fe6] (2a), [P][(HL)2Fe6] (2b) and dianionic [Bu4N]2[(HL)2Fe6] (3) (Scheme 1). Direct comparison between the overall oxidation level (n) and the core ironanilide structural metrics yield a remarkable correlation. In general, while the average metal-metal distance within the hexanuclear octahedron decreases slightly throughout the unsolvated clusters (n = +1  –2, Figure 5a) the iron-anilide mean distance expands with increasing number of valence electrons (VE) at [Fe6] (Figure 5b). The opposite trend is seen for the Fe–N(H)–Feavg angle which becomes more acute upon increasing the number of VE (Figure 5c). The comparison of the Fe–Feavg in the pyridine-bound species [Bu4N][(HL)2Fe6(py)2] (n = –1; VE = 37)125 and [(HL)2Fe6(py)2][PF6] (1b, n = +1; VE = 35) is less straight forward due to the significant distortion enforced by solvent ligation, the Fe–N(H)avg distance and Fe–N(H)–Feavg angle preserves the overall trend observed in the unsolvated clusters (Figure S21). 4.2. Electronic structure. Remarkably the spin ground states of the reduced clusters undergo a significant electronic ground state perturbation that drastically increases from an S = 6 in the all-ferrous (HL)2Fe6 to an S = 19/2 for 2b and 11 for 3. In contrast, the removal of an electron from the [Fe6] core reduces the spin to an S = 11/2 for unsolvated 1a where upon ligation of pyridine decreases further to 9/2 in [(HL)2Fe6(py)2][PF6] (1b) (Figure S17 and S22). A qualitative molecular orbital bonding approach developed previously for hexanuclear edge-bridged Nb and Ta clusters148 has been adapted to the all-ferrous cluster (HL)2Fe6 to account for its S = 6 ground state.61 The axis system at each iron site was chosen so that the dz2 orbital is oriented normal to the local N4 plane comprised of the ligand anilides to which the dx2–y2 orbital is directed. The frontier iron 3d orbitals were combined according to symmetry considerations and filled with the available 36 iron valence electrons. To account for the observed S = 6, the symmetry adapted linear combinations (SALCs) of the iron-anilido (Fe– N) antibonding interactions (1a2g, 2eg and 2t2u) are highest in energy and must exceed the mean

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

Page 10 of 19

Figure 5. Structural comparison of (a) Fe–Feavg and (b) Fe–N(H)avg distances, and (c) Fe–N(H)-Feavg angle in the series 1a, 2a, 2b, and 3 (n = +1  –2) including the all-ferrous cluster (HL)2Fe6 (n = 0). The black open symbol at n = –1 correspond to the metrics of 2b. Qualitative molecular orbital diagram derived for neutral (HL)2Fe6 (e);61 oxidized [(HL)2Fe6]+ (1a, d); and reduced [(HL)2Fe6]– (2a or 2b, f), [(HL)2Fe6]2– (3, g) to explain the observed spin ground state.

spin-pairing energy149 to remain unpopulated (ΔE(1a2g  1eg) > spin pairing energy; Figure 5e). As described in our previous report for [Bu4N][(HL)2Fe6(L’)2] the addition of an electron to (HL)2Fe6 likely proceeds by population of the σ*(FeN) band (1a2g, 2eg, 2t2u), weakening the Fe–N bonds as a result.125 The latter effect is structurally manifest in longer Fe–N bonds (Table 1) and follows a correlation with the overall oxidation level n (Figure 5b), tracking with the spin ground state (i.e., longer Fe–N with increasing S). Following our proposed MO diagram, we predict that removal of an electron from the HOMO (2 t1u) in the absence of ligand perturbations would result in an S = 11/2 species (Figure 5d). This is confirmed experimentally by the S = 11/2 cluster 1a. Pyridine ligation to trans-disposed iron sites in 1b affords [(HL)2Fe6(py)2][PF6] that results in a spin ground state of an S = 9/2 (Figure S17 and S22). The decrease in spin ground state going from 1a to 1b, can be explained where two nominally dz2 based SALCs rise in energy due to pyridine coordination, thus effectively decreasing the number of energetically accessible molecular orbitals and forcing the spin to decrease to 9/2 (Figure S22ab). 4.2.1. Spin ground state of monoanionic 2. Previously125 we speculated that perturbation of two dz2-based SALCs arising from pyridine ligation raised two orbitals high enough in energy so as to remain unpopulated akin to the effects of

solvent ligation in the cationic clusters. Furthermore, we posited that removal of the pyridine might result in a maximally high-spin configuration of S = 23/2, but that is not observed. Indeed, the ground spin-state observed for [P][(HL)2Fe6] (2b) is identical (S = 19/2) to the solvento adducts [(HL)2Fe6(L)2] (L = py, dmf). Thus, the molecular orbital configuration proposed consistent with the true orbital ordering, the observed spin ground state (S = 19/2) must solely arise from partial population of the σ*(FeN) band (1a2g, 2eg, 2t2u), leaving two nominally antibonding orbitals of x2y2 parentage empty (Figure 5f and S22d-e). A closer inspection of the FeN bond metrics of 2b reveal a clear contraction in the xy plane for two, trans-disposed iron centers of the anion. Each of the anions in the asymmetric unit (constituting one half of each [(HL)2Fe6] of crystal structure of 2b feature one iron site with short FeN bonds (2.010(4), 2.033(4) Å) and two iron sites with longer FeN bonds (2.109(4), 2.104(4) Å). Thus, two iron sites per anionic cluster each feature an orbital of x2-y2 parentage that remains empty in the partially filled σ*(FeN) band (1a2g, 2eg, 2t2u) giving rise to the distortion observed. Furthermore, the two highest lying orbitals that remain unpopulated simply exceed the spin-pairing energy for iron in this configuration, rather than for the reasons previously proposed (Figure 5f).125

ACS Paragon Plus Environment

Page 11 of 19 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

Journal of the American Chemical Society

4.2.2. Spin ground state of dianionic 3. Addition of an electron to [(HL)2Fe6]– affords the dianionic species [Bu4N]2[(HL)2Fe6] (3) that has been determined to have an S = 11 ground state. From our molecular orbital considerations, 2b has not populated all the available antibonding SALCs (1a2g, 2eg, or 2t2u) yielding an S = 19/2, not 23/2 spin ground state (Figure 5f). Upon reduction of [(HL)2Fe6]– to 3, however, the additional electron promotes a maximally high spin species via population of the entire FeN antibonding set, thus achieving an S = 11 configuration (Figure 5g). Consistent with that assessment, 3 features the longest Fe–Navg bonds (2.127(9) Å) measured in the electron-transfer series (Figure 5b), consistent with the bond metrics observed for trinuclear clusters that feature maximally high spin ground states with iron centers in similar geometries (i.e., (PhL)Fe3(thf)3, dFeN 2.167(6) Å).53 Each successive reduction ([Fe6]  [Fe6]  [Fe6]2) adds an electron to nominally antibonding orbitals of σ*FeFe or σ*Fe-N origin (Figure 5E). Population of the FeN antibonding orbital set (σ*Fe-N) lowers the overall band energy, promoting a higher population of the σ*FeN orbitals, resulting in a higher overall spin for the anion and dianion. Thus, assuming the high spin electronic configurations maximizes the electron exchange interaction stabilization effect, offsetting the destabilizing impact of populating a greater number of antibonding interactions. Conversely, oxidation of the [Fe6] core stabilizes the cluster by removing electrons from antibonding orbitals, thus the electronic configurations arrange into lower spin states. For 1  3 the spin ground states remain isolated up to room temperature. We attribute the available intracore direct orbital overlap to provide a direct exchange pathway that results in the observed S = 6 ground state for (HL)2Fe6. Both theoretical64 and experimental69, 78, 150 work has shown that double exchange enhances the ferromagnetic interaction in mixed valence complexes. For the clusters studied herein, the double exchange pathway is the same as the direct exchange pathways mediated by direct FeFe orbital overlap. This electron exchange pathway enhancement is manifest in the modest shortening of the Fe–Feavg distance upon reduction (Figure 5a), similar to other strongly coupled clusters, e.g. [(tbsL)Fe3]–.126 4.3. Mixed valency and redox delocalization. Fully delocalized compounds have spin-dependent IVCTs given by 𝐸op(𝑆) = 2 |𝐵|(𝑆 + 1 2),66 where B corresponds to the double exchange interaction energy. For 3 the calculated value of B is 309 cm–1. Fully delocalized complexes have no thermal barrier for electron transfer and thus the ground state adiabatic energy surface has a single minimum.139 This occurs when the following condition is met 2Hab/max ≥ 1. Two methods were presented earlier by which the extent of delocalization is quantified. The first (Hab) is well-known to work best for partially delocalized complexes (Class II) and significantly underestimates Hab as delocalization increases towards borderline Class II-III or fully delocalized Class III (the underestimation reaches 50% when 2Hab = max).151 Thus, via the Hush formulae Hab is calculated to be 1045.4 cm–1; while by the second method we obtain  = 0.33. Both methods point towards a moderately coupled Class II species; in

contrast the experimental shape of the IVCT band in 3 (Gaussian on the high energy and skewed on the low energy side) suggest strong delocalization. While the transitions observed could also be of d-to-d origin, the low energy band asymmetry displays a common characteristic of IVCTs. Other experimental evidence of delocalization in 3 is seen in the Mössbauer spectrum at 4.2 K where the six iron centers display a single iron environment indicative of complete shared oxidation, therefore the electron transfer rate within the [Fe6] core occurs faster than 10–7 s. 4.4. Single molecule magnetism behavior. The monoanionic 2b features a spin ground state (S = 19/2) equal to the solvated clusters [(HL)2Fe6(L)2] (L = py, dmf).125 Cluster 2b also displays slow relaxation of the magnetization. The effective spin reversal barrier has an almost identical magnitude, 42.5(8) and 43.8(5) cm–1 for [Bu4N][(HL)2Fe6(py)2] and 2b, respectively. In contrast they differ quite significantly below the blocking temperature, for instance the pyridine-bound cluster displays a well-resolved hyperfine split Mössbauer spectrum at 4.2 K and a step-like behavior in its magnetic hysteresis at 1.8 K, while 2b features a broad and unresolved spectrum with narrower magnetic hysteresis measured under identical conditions. Magnetic analysis of 1a confirm its SMM behavior with Ueff = 21.7(1) cm–1. The theoretical values calculated from 𝑈 = |𝐷|(𝑆2 ― 1 4) for 1a (33.9 cm–1) and 2b (68.4 cm–1) yield values higher than the observed Ueff which is typical of systems where the spin reversal does not follow strictly the thermal barrier, but instead tunnels through the barrier at some higher energy ms levels.83, 152 We previously observed that structural distortions of the [Fe6] octahedron accompany the emergence of SMM behavior. In fact, elongation on the solvated clusters [(HL)2Fe6(L’)2]– (L’ = pyridine or DMF) was observed along the Fe–Fetrans sites coordinated by L’.125 However, we now note that the solvation sites do not occur at the iron sites where an xy-contraction is observed. Similarly, the [Fe6] clusters in 1a, 2a and 2b distort, each featuring two short (almost identical) and one long Fe–Fetrans distance (Table 1). In comparison 3 has a highly symmetric [Fe6] octahedron with Fe–Fetrans distances of 3.630(1), 3.643(1) and 3.646(1) Å; thus, 3 would be expected to be absent of slow relaxation of the magnetization effects. In agreement, Mössbauer data at 4.2 K and ac magnetic susceptibility studies on 3 corroborated the absence of SMM behavior. To date the most successful methods to achieve high spin ground states in polynuclear systems rely on the use of superexchange electronic coupling interactions. Alternatively, the current account demonstrates that significantly strong magnetic interaction in mixed valence complexes can be achieved by the introduction of double exchange. This has been previously articulated150 and shown experimentally in a few dinuclear coordination complexes.6774 Nonetheless, the spin ground states of [P][(HL) Fe ] (2b) 2 6 and [Bu4N]2[(HL)2Fe6] (3) represent the highest spin ground states recorded to date that persist to room temperature. Manipulation of the auxiliary ligands around the iron centers may provide a pathway for enhanced SMM behavior as already shown in [(HL)2Fe6(L’)2]– (L’ = pyridine or DMF).

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

5. CONCLUSIONS

2.

A series of isostructural octahedral hexairon clusters are shown to display spin ground states varying from S = 9/2 to S = 11, marking thus the record of the highest spin states that remain isolated up to 300 K. Such behavior is attributed to double exchange manifested in the strongly delocalized clusters. Consequences of the latter are the hyperfine splitting of the Mössbauer spectrum at low temperature and the emergence of SMM behavior. As seen in the structural metrics, the existence of magnetic anisotropy can be predicted by the asymmetry of the [Fe6] octahedron and further manipulated by simple solvation of the cluster. The latter provides a pathway by which the magnetic behavior can be tuned and enhanced by simple exploration of other auxiliary ligands.

Hush, N. S., Intervalence-Transfer Absorption. Part 2. Theoretical Considerations and Spectroscopic Data. In Prog.

Inorg. Chem., John Wiley & Sons, Inc.: 1967; pp 391-444. 3.

Robin, M. B.; Day, P., Mixed Valence Chemistry-A Survey and Classification. In Advances in Inorganic Chemistry and

Radiochemistry, Emeléus, H. J.; Sharpe, A. G., Eds. Academic Press: 1968; Vol. Volume 10, pp 247-422. 4.

Hush, N. S., "Homogeneous and heterogeneous optical and thermal electron transfer." Electrochim. Acta. 1968, 13, 1005.

5.

Creutz, C., "Mixed-valence complexes of d5-d6 metal centers."

Prog. Inorg. Chem. 1983, 30, 1. 6.

Ward, M. D., "Metal-metal interactions in binuclear complexes exhibiting mixed-valency: molecular wires and switches."

Chem. Soc. Rev. 1995, 24, 121.

ASSOCIATED CONTENT

7.

Supporting Information

Kaim, W.; Klein, A.; Glockle, M., "Exploration of mixed valence chemistry: Inventing new analogues of the Creutz-

The Supporting Information is available free of charge on the ACS Publications website. X-ray crystal structure images, crystallographic data (cif), Mössbauer and near-infrared spectra, and magnetometry data.

8.

AUTHOR INFORMATION

9.

Taube ion." Acc. Chem. Res. 2000, 33, 755. Demadis, K. D.; Hartshorn, C. M.; Meyer, T. J., "The localizedto-delocalized transition in mixed-valence chemistry." Chem.

Rev. 2001, 101, 2655. Burns, R. G., "Intervalence transitions in mixed-valence minerals of iron and titanium." Annu. Rev. Earth Pl. Sc. 1981,

Corresponding Author

9, 345.

*[email protected]

10. Beinert, H.; Holm, R. H.; Munck, E., "Iron-sulfur clusters:

Present Addresses †Department of Chemistry, University of Pittsburgh, 219 Parkman Avenue, Pittsburgh, PA 15260, United States

Nature's modular, multipurpose structures." Science 1997,

277, 653. 11. Rao, P. V.; Holm, R. H., "Synthetic analogues of the active sites of iron-sulfur proteins." Chem. Rev. 2004, 104, 527.

Notes The authors declare no competing financial interest.

12. Hagen, W. R., "EPR spectroscopy of iron-sulfur proteins."

Adv. Inorg. Chem. 1992, 38, 165.

ACKNOWLEDGMENT This work was supported by a grant from the NIH (GM 098395), DOE (DE-SC0008313), and Harvard University. R.H.S. gratefully acknowledges Consejo Nacional de Ciencia y Tecnología (CONACYT) and Fundación México for a doctoral fellowship. T.A.B. gratefully acknowledges support from the Dreyfus Foundation (Teacher-Scholar Award). ChemMatCARS Sector 15 is supported by the National Science Foundation under grant number NSF/CHE-1346572. This research used resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract No. DEAC02-06CH11357.

REFERENCES 1.

Page 12 of 19

13. Hagen, W. R.; Wassink, H.; Eady, R. R.; Smith, B. E.; Haaker, H., "Quantitative EPR of an S = 7/2 system in thionine-oxidized Mofe proteins of nitrogenase - a redefinition of the P-cluster concept." Eur. J. Biochem. 1987, 169, 457. 14. Pierik, A. J.; Hagen, W. R., "S = 9/2 EPR signals are evidence against coupling between the siroheme and the Fe/S cluster prosthetic groups in Desulfovibrio vulgaris (Hildenborough) dissimilatory sulfite reductase." Eur. J. Biochem. 1991, 195, 505. 15. Pierik, A. J.; Hagen, W. R.; Dunham, W. R.; Sands, R. H., "Multifrequency spectroscopy

of

EPR

and

a

putative

[Fe6S6]

prismane-cluster-

Desulfovibrio

vulgaris

containing

Part 1. Qualitative Evidence for Intervalence-Transfer

(Hildenborough) - characterization of a supercluster and

Absorption in Inorganic Systems in Solution and in the Solid

superspin model protein." Eur. J. Biochem. 1992, 206, 705.

pp 357-389.

from

Mössbauer-

Allen, G. C.; Hush, N. S., Intervalence-Transfer Absorption.

State. In Prog. Inorg. Chem., John Wiley & Sons, Inc.: 1967;

protein

high-resolution

16. Pierik, A. J.; Wassink, H.; Haaker, H.; Hagen, W. R., "Redox properties and EPR spectroscopy of the P-clusters of

Azotobacter vinelandii MoFe protein." Eur. J. Biochem. 1993, 212, 51.

ACS Paragon Plus Environment

Page 13 of 19 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

Journal of the American Chemical Society

17. Long, J. R.; Williamson, A. S.; Holm, R. H., "Dimensional

29. Mullins, C. S.; Pecoraro, V. L., "Reflections on small molecule

Reduction of Re6Se8Cl2 - sheets, chains, and discrete clusters

manganese models that seek to mimic photosynthetic water

composed of chloride-terminated [Re6Q8]2+ (Q = S, Se)

oxidation chemistry." Coord. Chem. Rev. 2008, 252, 416.

cores." Angew. Chem. Int. Ed. 1995, 34, 226.

30. Krinsky, J. L.; Anderson, L. L.; Arnold, J.; Bergman, R. G.,

18. Tulsky, E. G.; Long, J. R., "Dimensional reduction: a practical formalism for manipulating solid structures." Chem. Mat. 2001, 13, 1149.

"Oxygen-centered

hexatantalum

tetradecaimido

cluster

complexes." Inorg. Chem. 2008, 47, 1053. 31. Corrigan, J. F.; Fuhr, O.; Fenske, D., "Metal chalcogenide

19. Long, J. R.; McCarty, L. S.; Holm, R. H., "A solid-state route to molecular clusters: access to the solution chemistry of [Re6Q8]2+ (Q = S, Se) core-containing clusters via dimensional reduction." J. Am. Chem. Soc. 1996, 118, 4603.

clusters on the border between molecules and materials."

Adv. Mat. 2009, 21, 1867. 32. Keen, A. L.; Doster, M.; Han, H.; Johnson, S. A., "Facile assembly of a Cu9 amido complex: a new tripodal ligand

20. Averill, B. A.; Herskovitz, T.; Holm, R. H.; Ibers, J. A., "Synthetic analogs of active-sites of iron-sulfur proteins. Synthesis and structure of tetra[mercapto-3-sulfido-iron] clusters, [Fe4S4(Sr)4]2-." J. Am. Chem. Soc. 1973, 95, 3523.

design that promotes transition metal cluster formation."

Chem. Commun. 2006, 1221. 33. Zhou, J.; Raebiger, J. W.; Crawford, C. A.; Holm, R. H., "Metal ion incorporation reactions of the cluster [Fe3S4(LS3)]3-,

21. Saito, T.; Yamamoto, N.; Yamagata, T.; Imoto, H., "Synthesis of [Mo6S8(PEt3)6] by reductive dimerization of a trinuclear

containing the cuboidal [Fe3S4]0 core." J. Am. Chem. Soc. 1997, 119, 6242.

molybdenum chloro sulfido cluster complex coordinated with

34. Kanady, J. S.; Tsui, E. Y.; Day, M. W.; Agapie, T., "A synthetic

triethylphosphine and methanol - a molecular-model for

model of the Mn3Ca subsite of the oxygen-evolving complex

superconducting chevrel phases." J. Am. Chem. Soc. 1988,

in photosystem II." Science 2011, 333, 733.

110, 1646.

35. Tsui, E. Y.; Kanady, J. S.; Day, M. W.; Agapie, T., "Trinuclear

22. Saito, T.; Yoshikawa, A.; Yamagata, T.; Imoto, H.; Unoura, K.,

first row transition metal complexes of a hexapyridyl, trialkoxy

3

"Synthesis, structure, and electronic-properties of octakis( -

1,3,5-triarylbenzene ligand." Chem. Commun. 2011, 47,

sulfido)hexakis(triethylphosphine)hexatungsten

a

4189.

as

tungsten analog of the molecular-model for superconducting chevrel phases." Inorg. Chem. 1989, 28, 3588.

36. Liu, R. L.; von Malotki, C.; Arnold, L.; Koshino, N.; Higashimura, H.; Baumgarten, M.; Mullen, K., "Triangular

23. Fenske, D.; Grissinger, A.; Loos, M.; Magull, J., "New clusters

trinuclear metal-N4 complexes with high electrocatalytic

of Zr and V - the crystal-structures of [Cp6Zr6S9] and

activity for oxygen reduction." J. Am. Chem. Soc. 2011, 133,

[V6Se8O(PMe3)6]." Z. Anorg. Allg. Chem. 1991, 598, 121.

10372.

24. Goddard, C. A.; Long, J. R.; Holm, R. H., "Synthesis and

37. Tsui, E. Y.; Day, M. W.; Agapie, T., "Trinucleating copper:

characterization of four consecutive members of the five-

synthesis

member [Fe6S8(PEt3)6]n+ (n = 0-4) cluster electron transfer

complexes supported by a hexapyridyl 1,3,5-triarylbenzene

series." Inorg. Chem. 1996, 35, 4347.

and

magnetostructural

characterization

of

ligand." Angew. Chem. Int. Ed. 2011, 50, 1668.

25. Tran, N. T.; Kawano, M.; Dahl, L. F., "High-nuclearity

38. Guillet, G. L.; Sloane, F. T.; Ermert, D. M.; Calkins, M. W.;

palladium carbonyl trimethylphosphine clusters containing

Peprah, M. K.; Knowles, E. S.; Cizmar, E.; Abboud, K. A.;

unprecedented

icosahedral-based

Meisel, M. W.; Murray, L. J., "Preorganized assembly of three

transition-metal core geometries: proposed growth patterns

iron(II) or manganese(II) beta-diketiminate complexes using

face-condensed

from a centered Pd13 icosahedron." J. Chem. Soc. Dalton 2001, 2731. 26. Crawford, N. R. M.; Long, J. R., "Edge-bridged octahedral tungsten-oxygen-chlorine

a cyclophane ligand." Chem. Commun. 2013, 49, 6635. 39. Tsui, E. Y.; Agapie, T., "Reduction potentials of heterometallic

clusters:

synthesis

and

characterization of 2D 3d-symmetric [W6O6Cl12] isomers and 2-

[W6O7Cl11]3-." Inorg. Chem. 2001, 40, 3456.

manganese-oxido cubane complexes modulated by redoxinactive metals." Proc. Natl. Acad. Sci. USA 2013, 110, 10084. 40. Lionetti, D.; Day, M. W.; Agapie, T., "Metal-templated ligand

27. Gray, T. G., "Hexanuclear and higher nuclearity clusters of the Groups 4-7 metals with stabilizing -donor ligands."

Coord. Chem. Rev. 2003, 243, 213.

architectures for trinuclear chemistry: tricopper complexes and their O2 reactivity." Chem. Sci. 2013, 4, 785. 41. Suseno, S.; Horak, K. T.; Day, M. W.; Agapie, T., "Trinuclear

28. Lee, S. C.; Holm, R. H., "The clusters of nitrogenase:

nickel complexes with metal-arene interactions supported by

synthetic methodology in the construction of weak-field

tris-

clusters." Chem. Rev. 2004, 104, 1135.

Organometallics 2013, 32, 6883.

ACS Paragon Plus Environment

and

bis(phosphinoaryl)benzene

frameworks."

Journal of the American Chemical Society 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

42. Herbert,

D.

E.;

Lionetti,

D.;

Rittle,

J.;

Agapie,

T.,

Page 14 of 19

54. Eames, E. V.; Hernández Sánchez, R.; Betley, T. A., "Metal

"Heterometallic triiron-oxo/hydroxo clusters: effect of redox-

atom lability in polynuclear complexes." Inorg. Chem. 2013,

inactive metals." J. Am. Chem. Soc. 2013, 135, 19075.

52, 5006.

43. Di Francesco, G. N.; Gaillard, A.; Ghiviriga, I.; Abboud, K. A.;

55. Kuppuswamy, S.; Powers, T. M.; Johnson, B. M.; Bezpalko,

Murray, L. J., "Modeling biological copper clusters: synthesis

M. W.; Brozek, C. K.; Foxman, B. M.; Berben, L. A.; Thomas,

of a tricopper complex, and its chloride- and sulfide-bridged

C. M., "Metal-metal interactions in C3-symmetric diiron imido

congeners." Inorg. Chem. 2014, 53, 4647.

complexes linked by phosphinoamide ligands." Inorg. Chem.

44. Murray, L. J.; Weare, W. W.; Shearer, J.; Mitchell, A. D.; Abboud, K. A., "Isolation of a (dinitrogen)tricopper(I) complex." J. Am. Chem. Soc. 2014, 136, 13502.

2013, 52, 4802. 56. Krogman, J. P.; Thomas, C. M., "Metal-metal multiple bonding in C-3-symmetric bimetallic complexes of the first row

45. Guillet, G. L.; Gordon, J. B.; Di Francesco, G. N.; Calkins, M.

transition metals." Chem. Commun. 2014, 50, 5115.

W.; Cizmar, E.; Abboud, K. A.; Meisel, M. W.; Garcia-Serres,

57. Kuppuswamy, S.; Bezpalko, M. W.; Powers, T. M.; Wilding,

R.; Murray, L. J., "A family of tri- and dimetallic pyridine

M. J. T.; Brozek, C. K.; Foxman, B. M.; Thomas, C. M., "A

dicarboxamide cryptates: unusual O,N,O-coordination and

series of C3-symmetric heterobimetallic Cr-M (M = Fe, Co and

facile access to secondary coordination sphere hydrogen bonding interactions." Inorg. Chem. 2015, 54, 2691.

Cu) complexes." Chem. Sci. 2014, 5, 1617. 58. Fout, A. R.; Xiao, D. J.; Zhao, Q. L.; Harris, T. D.; King, E. R.;

46. Hatnean, J. A.; Raturi, R.; Lefebvre, J.; Leznoff, D. B.; Lawes,

Eames, E. V.; Zheng, S. L.; Betley, T. A., "Trigonal Mn3 and

G.; Johnson, S. A., "Assembly of triangular trimetallic

Co3 clusters supported by weak-field ligands: a structural,

complexes by triamidophosphine ligands: spin-frustrated

spectroscopic, magnetic, and computational investigation into

Mn2+ plaquettes and diamagnetic Mg2+ analogues with a

the correlation of molecular and electronic structure." Inorg.

combined through-space, through-bond pathway for

31

P- P 31

spin-spin coupling." J. Am. Chem. Soc. 2006, 128, 14992. 47. Mateus, P.; Delgado, R.; Lloret, F.; Cano, J.; Brandão, P.; Félix, V., "A trinuclear copper(II) cryptate and its μ 3-CO3 cascade complex: thermodynamics, structural and magnetic properties." Chem. Eur. J. 2011, 17, 11193.

Chem. 2012, 51, 10290. 59. Powers, T. M.; Betley, T. A., "Testing the polynuclear hypothesis: multielectron reduction of small molecules by triiron reaction sites." J. Am. Chem. Soc. 2013, 135, 12289. 60. Powers, T. M.; Fout, A. R.; Zheng, S. L.; Betley, T. A., "Oxidative group transfer to a triiron complex to form a

48. Parmelee, S. R.; Mazzacano, T. J.; Zhu, Y.; Mankad, N. P., "A heterobimetallic mechanism for C-H borylation elucidated from experimental and computational data." ACS Catal. 2015,

5, 3689.

nucleophilic μ3-nitride, [Fe3(μ3-N)]-." J. Am. Chem. Soc. 2011,

133, 3336. 61. Zhao, Q. L.; Harris, T. D.; Betley, T. A., "[(HL)2Fe6(NCMe)m]n+ (m = 0, 2, 4, 6; n = -1, 0, 1, 2, 3, 4, 6): An electron-transfer

49. Harris, T. D.; Betley, T. A., "Multi-site reactivity: reduction of

series featuring octahedral Fe6 clusters supported by a

six equivalents of nitrite to give an Fe6(NO)6 cluster with a

hexaamide ligand platform." J. Am. Chem. Soc. 2011, 133,

dramatically expanded octahedral core." J. Am. Chem. Soc. 2011, 133, 13852.

8293. 62. Hernández Sánchez, R.; Zheng, S. L.; Betley, T. A., "Ligand

50. Harris, T. D.; Zhao, Q. L.; Hernández Sánchez, R.; Betley, T.

field strength mediates electron delocalization in octahedral

A., "Expanded redox accessibility via ligand substitution in an

[(HL)2Fe6(L ')m]n+ clusters." J. Am. Chem. Soc. 2015, 137,

octahedral Fe6Br6 cluster." Chem. Commun. 2011, 47, 6344.

11126.

51. Hazra, S.; Sasmal, S.; Fleck, M.; Grandjean, F.; Sougrati, M.

63. Zener, C., "Interaction between the d-shells in the transition

T.; Ghosh, M.; Harris, T. D.; Bonville, P.; Long, G. J.;

metals. 2. Ferromagnetic compounds of manganese with

Mohanta, S., "Slow magnetic relaxation and electron delocalization in an S = 9/2 iron(II/III) complex with two crystallographically inequivalent iron sites." J. Chem. Phys. 2011, 134, 174507.

perovskite structure." Phys. Rev. 1951, 82, 403. 64. Anderson, P. W.; Hasegawa, H., "Considerations on double exchange." Phys. Rev. 1955, 100, 675. 65. Girerd, J. J., "Electron-transfer between magnetic ions in

52. Eames, E. V.; Betley, T. A., "Site-isolated redox reactivity in a trinuclear iron complex." Inorg. Chem. 2012, 51, 10274.

mixed-valence binuclear systems." J. Chem. Phys. 1983, 79, 1766.

53. Eames, E. V.; Harris, T. D.; Betley, T. A., "Modulation of

66. Blondin, G.; Girerd, J. J., "Interplay of electron exchange and

magnetic behavior via ligand-field effects in the trigonal

electron transfer in metal polynuclear complexes in proteins

clusters ( L)Fe3L*3 (L* = thf, py, PMe2Ph)." Chem. Sci. 2012,

or chemical models." Chem. Rev. 1990, 90, 1359.

Ph

3, 407.

ACS Paragon Plus Environment

Page 15 of 19 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

Journal of the American Chemical Society

67. Drueke, S.; Chaudhuri, P.; Pohl, K.; Wieghardt, K.; Ding, X.

77. Gaillard, J.; Moulis, J. M.; Auric, P.; Meyer, J., "High-

Q.; Bill, E.; Sawaryn, A.; Trautwein, A. X.; Winkler, H.;

multiplicity spin states of [Fe4Se4]+ clostridial ferredoxins."

Gurman, S. J., "The novel mixed-valence, exchange-coupled,

Biochem. 1986, 25, 464.

class-III dimer [L2Fe2(mu-OH)3] (L = N,N',N''-trimethyl-1,4,7-

78. Papaefthymiou, V.; Girerd, J. J.; Moura, I.; Moura, J. J. G.;

triazacyclononane)." J. Chem. Soc. Chem. Comm. 1989, 59.

Munck, E., "Mössbauer study of D. gigas ferredoxin II and

2+

68. Ding, X. Q.; Bominaar, E. L.; Bill, E.; Winkler, H.; Trautwein, A. X.; Drueke, S.; Chaudhuri, P.; Wieghardt, K., "Mossbauer

spin-coupling model for the Fe3S4 cluster with valence delocalization." J. Am. Chem. Soc. 1987, 109, 4703.

and electron-paramagnetic resonance study of the double-

79. Fox, B. G.; Surerus, K. K.; Munck, E.; Lipscomb, J. D.,

exchange and Heisenberg-exchange interactions in a novel

"Evidence for a -oxo-bridged binuclear iron cluster in the

binuclear Fe(II/III) delocalized-valence compound." J. Chem.

hydroxylase component of methane monooxygenase -

Phys. 1990, 92, 178.

Mössbauer and EPR studies." J. Biol. Chem. 1988, 263,

69. Gamelin, D. R.; Bominaar, E. L.; Kirk, M. L.; Wieghardt, K.;

10553.

Solomon, E. I., "Excited-state contributions to ground-state

80. Carney, M. J.; Papaefthymiou, G. C.; Spartalian, K.; Frankel,

properties of mixed-valence dimers: spectral and electronic-

R. B.; Holm, R. H., "Ground spin state variability in [Fe4S4Sr4]3-

structural studies of [Fe2(OH)3(tmtacn)2]

related to the

- synthetic analogs of the reduced clusters in ferredoxins and

[Fe2S2]+ active sites of plant-type ferredoxins." J. Am. Chem.

other iron sulfur proteins - cases of extreme sensitivity of

2+

electronic state and structure to extrinsic factors." J. Am.

Soc. 1996, 118, 8085. 70. Cotton, F. A.; Daniels, L. M.; Falvello, L. R.; Murillo, C. A., "A new class of dinuclear compounds: the synthesis and X-ray structural

characterization

of

tris(-

Chem. Soc. 1988, 110, 6084. 81. Hendrich, M. P.; Elgren, T. E.; Que, L., "A mixed-valence form of the iron cluster in the B2 protein of ribonucleotide reductase

diphenylformamidinato)diiron." Inorg. Chim. Acta 1994, 219,

from Escherichia coli." Biochem. Biophys. Res. Commun.

7.

1991, 176, 705.

71. Cotton, F. A.; Daniels, L. M.; Maloney, D. J.; Murillo, C. A., "Tri-bridged

amidinato

compounds

of

dicobalt

1:

82. Achim, C.; Golinelli, M. P.; Bominaar, E. L.; Meyer, J.; Munck, E., "Mössbauer study of Cys56Ser mutant 2Fe ferredoxin

Co2[PhNC(R)NPh]3 with R = H and C6H5." Inorg. Chim. Acta

from

1996, 249, 9.

exchange in an [Fe2S2]+ cluster." J. Am. Chem. Soc. 1996,

72. Cotton, F. A.; Daniels, L. M.; Falvello, L. R.; Matonic, J. H.; Murillo, C. A., "Trigonal-lantern dinuclear compounds of diiron(I,II): the synthesis and characterization of two highly paramagnetic Fe2(amidinato)3 species with short metal-metal bonds." Inorg. Chim. Acta 1997, 256, 269.

Clostridium

pasteurianum:

evidence

for

double

118, 8168. 83. Christou, G.; Gatteschi, D.; Hendrickson, D. N.; Sessoli, R., "Single-molecule magnets." MRS Bull. 2000, 25, 66. 84. Pedersen, K. S.; Bendix, J.; Clerac, R., "Single-molecule magnet engineering: building-block approaches." Chem.

73. Zall, C. M.; Zherebetskyy, D.; Dzubak, A. L.; Bill, E.; Gagliardi,

Commun. 2014, 50, 4396.

L.; Lu, C. C., "A combined spectroscopic and computational

85. Pilawa, B.; Kelemen, M. T.; Wanka, S.; Geisselmann, A.;

study of a high-spin S = 7/2 diiron complex with a short iron-

Barra, A. L., "Magnetic properties of a new spin cluster

iron bond." Inorg. Chem. 2012, 51, 728.

topology with high-spin ground state: the spin cluster

74. Zall, C. M.; Clouston, L. J.; Young, V. G.; Ding, K. Y.; Kim, H. J.; Zherebetskyy, D.; Chen, Y. S.; Bill, E.; Gagliardi, L.; Lu, C.

[Mn4IIMn3III(teaH)3(tea)3](ClO4)2 . 3MeOH." Europhys. Lett. 1998, 43, 7.

C., "Mixed-valent dicobalt and iron-cobalt complexes with

86. Miyasaka, H.; Nakata, K.; Sugiura, K.; Yamashita, M.; Clerac,

high-spin configurations and short metal-metal bonds." Inorg.

R., "A three-dimensional ferrimagnet composed of mixed-

Chem. 2013, 52, 9216.

valence Mn4 clusters linked by an {Mn[N(CN)2]6}4- unit."

75. Lindahl, P. A.; Day, E. P.; Kent, T. A.; Ormejohnson, W. H.;

Angew. Chem. Int. Ed. 2004, 43, 707.

Munck, E., "Mössbauer, EPR, and magnetization studies of

87. Koizumi, S.; Nihei, M.; Shiga, T.; Nakano, M.; Nojiri, H.;

the Azotobacter-vinelandii Fe protein - evidence for a

Bircherrl, R.; Waldmann, O.; Ochsenbein, S. T.; Guedel, H.

cluster with spin S = 3/2." J. Biol. Chem. 1985, 260,

U.; Fernandez-Alonso, F.; Oshio, H., "A wheel-whaped

[Fe4S4]

1+

1160.

wingle-molecule magnet of [Mn3IIMn4III]: quantum tunneling of

76. Munck, E.; Kent, T. A., "Structure and magnetism of ironsulfur clusters in proteins." Hyperfine Interact. 1986, 27, 161.

magnetization under static and pulse magnetic fields." Chem.

Eur. J. 2007, 13, 8445. 88. Caneschi, A.; Gatteschi, D.; Sessoli, R.; Barra, A. L.; Brunel, L. C.; Guillot, M., "Alternating current susceptibility, high-field

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

magnetization, and millimeter band EPR evidence for a

S

ground

=

10

state

in

J.

[Mn12O12(CH3COO)16(H2O)4].2CH3COOH.4H2O."

Am.

Chem. Soc. 1991, 113, 5873. Hanke, D., "Magnetic properties of an octanuclear Iron(III) cation." Inorg. Chem. 1993, 32, 3099.

G.; Dalal, N. S., "Mn7 species with an S = high-frequency

EPR

studies

of

a

29

/2 ground state:

species

at

the

133, 17586. 101. Stamatatos, T. C.; Foguet-Albiol, D.; Poole, K. M.; Wernsdorfer, W.; Abboud, K. A.; O'Brien, T. A.; Christou, G.,

90. Barra, A. L.; Debrunner, P.; Gatteschi, D.; Schulz, C. E.; R.,

100. Wang, Z. X.; van Tol, J.; Taguchi, T.; Daniels, M. R.; Christou,

classical/quantum spin interface." J. Am. Chem. Soc. 2011,

89. Delfs, C.; Gatteschi, D.; Pardi, L.; Sessoli, R.; Wieghardt, K.;

Sessoli,

Page 16 of 19

"Superparamagnetic-like

behavior

in

an

octanuclear iron cluster." Europhys. Lett. 1996, 35, 133.

"Spin Maximization from S = 11 to S = 16 in Mn7 Disk-Like Clusters: Spin Frustration Effects and Their Computational Rationalization." Inorg. Chem. 2009, 48, 9831.

91. Caneschi, A.; Gatteschi, D.; Laugier, J.; Rey, P.; Sessoli, R.;

102. Langley, S. K.; Chilton, N. F.; Moubaraki, B.; Murray, K. S.,

Zanchini, C., "Preparation, crystal structure, and magnetic

"Self-assembled decanuclear NaI2MnII4MnIII4 complexes: from

properties of an oligonuclear complex with 12 coupled spins

discrete clusters to 1-D and 2-D structures, with the MnII4MnIII4

and an S = 12 ground state." J. Am. Chem. Soc. 1988, 110,

unit displaying a large spin ground state and probable SMM behaviour." Dalton Trans. 2011, 40, 12201.

2795. 92. Boyd, P. D. W.; Li, Q. Y.; Vincent, J. B.; Folting, K.; Chang, H.

103. Powell, A. K.; Heath, S. L.; Gatteschi, D.; Pardi, L.; Sessoli,

R.; Streib, W. E.; Huffman, J. C.; Christou, G.; Hendrickson,

R.; Spina, G.; Delgiallo, F.; Pieralli, F., "Synthesis, structures,

D. N., "Potential building blocks for molecular ferromagnets -

and magnetic properties of Fe2, Fe17, and Fe19 oxo-bridged

[Mn12O12(O2CPh)16(H2O)4] with a S = 14 ground state." J. Am.

iron clusters: the stabilization of high ground state spins by

Chem. Soc. 1988, 110, 8537.

cluster aggregates." J. Am. Chem. Soc. 1995, 117, 2491.

93. Sessoli, R.; Tsai, H. L.; Schake, A. R.; Wang, S. Y.; Vincent,

104. Goodwin, J. C.; Sessoli, R.; Gatteschi, D.; Wernsdorfer, W.;

J. B.; Folting, K.; Gatteschi, D.; Christou, G.; Hendrickson, D.

Powell, A. K.; Heath, S. L., "Towards nanostructured arrays

N., "High spin molecules: [Mn12O12(O2CR)16(H2O)4]." J. Am.

of single molecule magnets: new Fe19 oxyhydroxide clusters

Chem. Soc. 1993, 115, 1804.

displaying high ground state spins and hysteresis." J. Chem.

94. Sessoli, R.; Gatteschi, D.; Caneschi, A.; Novak, M. A., "Magnetic bistability in a metal-ion cluster." Nature 1993, 365, 141.

Soc. Dalton 2000, 1835. 105. Zhong, Z. J.; Seino, H.; Mizobe, Y.; Hidai, M.; Fujishima, A.; Ohkoshi, S.; Hashimoto, K., "A high-spin cyanide-bridged

95. Friedman, J. R.; Sarachik, M. P.; Tejada, J.; Ziolo, R., "Macroscopic measurement of resonant magnetization tunneling in high-spin molecules." Phys. Rev. Lett. 1996, 76, 3830.

Mn9W6 cluster S = 39/2 with a full-capped cubane structure." J.

Am. Chem. Soc. 2000, 122, 2952. 106. Stamatatos, T. C.; Abboud, K. A.; Wernsdorfer, W.; Christou, G., "High-nuclearity, high-symmetry, high-spin molecules: a

96. Aromi, G.; Claude, J. P.; Knapp, M. J.; Huffman, J. C.;

mixed-valence Mn10 cage possessing rare T symmetry and

Hendrickson, D. N.; Christou, G., "High-spin molecules:

an S = 22 ground state." Angew. Chem. Int. Ed. 2006, 45,

hexanuclear [Mn6O4Cl4(Me2dbm)6] (Me2dbmH = 4,4 '-

4134.

dimethyldibenzoylmethane)

tetrahedral

107. Manoli, M.; Johnstone, R. D. L.; Parsons, S.; Murrie, M.;

[Mn6O4Cl4]6+ core and a S = 12 ground state." J. Am. Chem.

with

a

near

Affronte, M.; Evangelisti, M.; Brechin, E. K., "A serromagnetic

Soc. 1998, 120.

mixed-valent Mn supertetrahedron: towards low-temperature

97. Milios, C. J.; Vinslava, A.; Wernsdorfer, W.; Moggach, S.; Parsons, S.; Perlepes, S. P.; Christou, G.; Brechin, E. K., "A record anisotropy barrier for a single-molecule magnet." J.

Am. Chem. Soc. 2007, 129, 2754.

magnetic refrigeration with molecular clusters." Angew.

Chem. Int. Ed. 2007, 46, 4456. 108. Stamatatos, T. C.; Abboud, K. A.; Wernsdorfer, W.; Christou, G., "Ferromagnetically-coupled decanuclear, mixed-valence

98. Moushi, E. E.; Stamatatos, T. C.; Wernsdorfer, W.;

[Mn10O4(N3)4(hmp)12]2+ [hmpH = 2-(hydroxymethyl)pyridine]

Nastopoulos, V.; Christou, G.; Tasiopoulos, A. J., "A family of

clusters with rare T symmetry and an S = 22 ground state."

3D coordination polymers composed of Mn19 magnetic units."

Angew. Chem. Int. Ed. 2006, 45, 7722.

Polyhedron 2007, 26, 2042. 109. Stamatatos, T. C.; Poole, K. M.; Abboud, K. A.; Wernsdorfer,

99. Goldberg, D. P.; Caneschi, A.; Lippard, S. J., "A decanuclear

W.; O'Brien, T. A.; Christou, G., "High-spin Mn4 and Mn10

mixed-valent manganese complex with a high-spin multiplicity

molecules: large spin changes with structure in mixed-

in the ground-state." J. Am. Chem. Soc. 1993, 115, 9299.

valence Mn4IIMn6III clusters with azide and alkoxide-based ligands." Inorg. Chem. 2008, 47, 5006.

ACS Paragon Plus Environment

Page 17 of 19 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

Journal of the American Chemical Society

110. Wu, G.; Huang, J.; Sun, L. L.; Bai, J. Q.; Li, G. H.; Cremades,

120. Ge, C. H.; Ni, Z. H.; Liu, C. M.; Cui, A. L.; Zhang, D. Q.; Kou,

E.; Ruiz, E.; Clerac, R.; Qiu, S. L., "ST = 22 [Mn10]

H. Z., "Nonadecanuclear Mn7IIMn12III cluster with a spin

supertetrahedral building-block to design extended magnetic

ground state of S =

networks." Inorg. Chem. 2011, 50, 8580.

73

/2." Inorg. Chem. Commun. 2008, 11,

675.

111. Low, D. M.; Jones, L. F.; Bell, A.; Brechin, E. K.; Mallah, T.;

121. Moushi, E. E.; Stamatatos, T. C.; Wernsdorfer, W.;

Riviere, E.; Teat, S. J.; McInnes, E. J. L., "Solvothermal

Nastopoulos, V.; Christou, G.; Tasiopoulos, A. J., "A Mn17

synthesis of a tetradecametallic FeIII cluster." Angew. Chem.

octahedron with a giant ground-state spin: occurrence in

Int. Ed. 2003, 42, 3781.

discrete

112. Shaw, R.; Laye, R. H.; Jones, L. F.; Low, D. M.; Talbot-

form

and

as

multidimensional

coordination

polymers." Inorg. Chem. 2009, 48, 5049.

Eeckelaers, C.; Wei, Q.; Milios, C. J.; Teat, S.; Helliwell, M.;

122. Ako, A. M.; Hewitt, I. J.; Mereacre, V.; Clerac, R.;

Raftery, J.; Evangelisti, M.; Affronte, M.; Collison, D.; Brechin,

Wernsdorfer, W.; Anson, C. E.; Powell, A. K., "A

E.

K.;

McInnes,

E.

tetradecametallic

J.

L.,

"1,2,3-Triazolate-bridged

transition

metal

clusters

ferromagnetically coupled Mn19 aggregate with a record S = 83

/2 ground spin state." Angew. Chem. Int. Ed. 2006, 45, 4926.

[M14(L)6O6(OMe)18X6] (M = FeIII, CrIII and VIII/IV) and related

123. Cremades, E.; Ruiz, E., "Magnetic properties of largest-spin

compounds: ground-state spins ranging from S = 0 to S = 25

single molecule magnets: Mn17 complexes - a density

and spin-enhanced magnetocaloric effect." Inorg. Chem.

functional theory approach." Inorg. Chem. 2010, 49, 9641.

2007, 46, 4968.

124. Mameri, S.; Ako, A. M.; Yesil, F.; Hibert, M.; Lan, Y. H.; Anson,

113. Larionova, J.; Gross, M.; Pilkington, M.; Andres, H.; Stoeckli-

C. E.; Powell, A. K., "Coordination cluster analogues of the

Evans, H.; Gudel, H. U.; Decurtins, S., "High-spin molecules:

high-spin

a novel cyano-bridged Mn9IIMo6V molecular cluster with a S =

bis(hydroxymethyl)phenol ligands." Eur. J. Inorg. Chem.

51

/2 ground state and ferromagnetic intercluster ordering at

low temperatures." Angew. Chem. Int. Ed. 2000, 39, 1605.

[Mn19]

system

with

functionalized

2,6-

2014, 4326. 125. Hernández Sánchez, R.; Betley, T. A., "Meta-atom behavior

114. Murugesu, M.; Habrych, M.; Wernsdorfer, W.; Abboud, K. A.; Christou, G., "Single-molecule magnets: a Mn25 complex with a record S = 51/2 spin for a molecular species." J. Am. Chem.

Soc. 2004, 126, 4766.

in clusters revealing large spin ground states." J. Am. Chem.

Soc. 2015, 137, 13949. 126. Hernández Sánchez, R.; Bartholomew, A. K.; Powers, T. M.; Ménard, G.; Betley, T. A., "Maximizing electron exchange in

115. Murugesu, M.; Takahashi, S.; Wilson, A.; Abboud, K. A.; Wernsdorfer, W.; Hill, S.; Christou, G., "Large Mn25 singlemolecule magnet with spin S =

a [Fe3] cluster." J. Am. Chem. Soc. 2016, 138, 2235. 127. Chavez, I.; Alvarez-Carena, A.; Molins, E.; Roig, A.;

/2: magnetic and high-

Maniukiewicz, W.; Arancibia, A.; Arancibia, V.; Brand, H.;

frequency electron paramagnetic resonance spectroscopic

Manriquez, J. M., "Selective oxidants for organometallic

characterization of a giant spin state." Inorg. Chem. 2008, 47,

compounds containing a stabilising anion of highly reactive

9459.

cations: (3,5(CF3)2C6H3)4B-Cp2Fe+ and (3,5(CF3)2C6H3)4B-

51

116. Charalambous, M.; Moushi, E. E.; Papatriantafyllopoulou, C.; Wernsdorfer, W.; Nastopoulos, V.; Christou, G.; Tasiopoulos, A.

J.,

"A

Mn36Ni4

'loop-of-loops-and-supertetrahedra'

aggregate possessing a high ST = 26 +/- 1 spin ground state."

Chem. Commun. 2012, 48, 5410. N. G. R.; Wernsdorfer, W.; Ansona, C. E.; Powell, A. K., "Two edge-sharing Mn4 Mn6 supertetrahedra give an anisotropic III

S = 28 +/- 1 Mn6IIMn11III complex." Dalton Trans. 2009, 1901. 118. Stamatatos, T. C.; Abboud, K. A.; Wernsdorfer, W.; Christou, G., ""Spin Tweaking" of a High-Spin Molecule: An Mn25 Single-Molecule Magnet with an S =

128. APEX2, v.2009; Bruker Analytical X-Ray Systems, Inc.: Madison, WI, 2009. 129. Scheldrick, G. M. SADABS, Version 2.03; Bruker Analytical X-Ray Systems, Inc.: Madison, WI, 2000.

117. Nayak, S.; Beltran, L. M. C.; Lan, Y. H.; Clerac, R.; Hearns, II

Cp2*Fe+." J Organomet Chem 2000, 601, 126.

61

/2 Ground State."

Angew. Chem. Int. Ed. 2007, 46, 884.

130. Spek, A. L. PLATON, A Multipurpose Crystallographic Tool; Utrecht University: Utrecht, The Netherlands, 2010. 131. Sheldrick, G. M. SHELXTL, Version 6.12; Bruker Analytical, X-Ray Systems, Inc.: Madison, WI, 2000. 132. Prisecaru, I., WMOSS4 Mössbauer Spectral Analysis Software. 2009-2013. 133. Zadrozny, J. M.; Xiao, D. J.; Long, J. R.; Atanasov, M.; Neese, F.; Grandjean, F.; Long, G. J., "Mössbauer spectroscopy as a

119. Stamatatos, T. C.; Abboud, K. A.; Wernsdorfer, W.; Christou,

probe of magnetization dynamics in the linear iron(I) and

G., "A new Mn25 single-molecule magnet with an S =

iron(II) complexes [Fe(C(SiMe3)3)2]1-/0." Inorg. Chem. 2013,

61

/2

ground state arising from ligand-induced 'spin-tweaking' in a

52, 13123.

high-spin molecule." Polyhedron 2007, 26, 2095.

ACS Paragon Plus Environment

Journal of the American Chemical Society 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

134. Zadrozny, J. M.; Xiao, D. J.; Atanasov, M.; Long, G. J.;

Page 18 of 19

143. Guo, Y. N.; Xu, G. F.; Guo, Y.; Tang, J. K., "Relaxation

Grandjean, F.; Neese, F.; Long, J. R., "Magnetic blocking in a

dynamics of dysprosium(III) single molecule magnets." Dalton

linear iron(I) complex." Nat. Chem. 2013, 5, 577.

Trans. 2011, 40, 9953.

135. Schulz, C. E.; Hu, C. J.; Scheidt, W. R., "On spin Hamiltonian

144. Cole, K. S.; Cole, R. H., "Dispersion and absorption in

fits to Mössbauer spectra of high-spin Fe(II) porphyrinate

dielectrics I. alternating current characteristics." J. Chem.

systems." Hyperfine Interact. 2006, 170, 55.

Phys. 1941, 9, 341.

136. Dutta, S. K.; Ensling, J.; Werner, R.; Florke, U.; Haase, W.;

145. Aubin, S. M. J.; Sun, Z. M.; Pardi, L.; Krzystek, J.; Folting, K.;

Gutlich, P.; Nag, K., "Valence-delocalized and valence-

Brunel, L. C.; Rheingold, A. L.; Christou, G.; Hendrickson, D.

trapped Fe Fe complexes: drastic influence of the ligands."

N., "Reduced anionic Mn12 molecules with half-integer ground

Angew. Chem. Int. Ed. 1997, 36, 152.

states as single-molecule magnets." Inorg. Chem. 1999, 38,

II

III

137. Day, P.; Hush, N. S.; Clark, R. J. H., "Mixed valence: origins and developments." Philos. Trans. Royal Soc. A 2008, 366, 5.

5329. 146. Yang, P., The Chemistry of Nanostructured Materials. World Scientific: 2003.

138. Brunschwig, B. S.; Creutz, C.; Sutin, N., "Optical transitions of symmetrical mixed-valence systems in the Class II-III transition regime." Chem. Soc. Rev. 2002, 31, 168. 139. D'alessandro, D. M.; Keene, F. R., "Current trends and future

147. Chilton, N. F. CC-FIT, 2014. 148. Cotton, F. A.; Haas, T. E., "Molecular orbital treatment of bonding in certain metal atom clusters." Inorg. Chem. 1964,

3, 10.

challenges in the experimental, theoretical and computational

149. Vanquickenborne, L. G.; Haspeslagh, L., "On the meaning of

analysis of intervalence charge transfer (IVCT) transitions."

spin-pairing energy in transition-metal ions." Inorg. Chem.

Chem. Soc. Rev. 2006, 35, 424.

1982, 21, 2448.

140. Chilton, N. F.; Anderson, R. P.; Turner, L. D.; Soncini, A.;

150. Bechlars, B.; D'Alessandro, D. M.; Jenkins, D. M.; Iavarone,

Murray, K. S., "PHI: A powerful new program for the analysis

A. T.; Glover, S. D.; Kubiak, C. P.; Long, J. R., "High-spin

of anisotropic monomeric and exchange-coupled polynuclear

ground states via electron delocalization in mixed-valence

d- and f-block complexes." J. Comput. Chem. 2013, 34, 1164.

imidazolate-bridged divanadium complexes." Nat. Chem.

141. Aromi, G.; Knapp, M. J.; Claude, J. P.; Huffman, J. C.; Hendrickson, D. N.; Christou, G., "High-spin molecules: hexanuclear Mn clusters with [Mn6O4X4] (X = Cl-, Br-) faceIII

6+

capped octahedral cores and S =12 ground states." J. Am.

Chem. Soc. 1999, 121, 5489.

S.

F.,

"“Almost

delocalized”

intervalence

compounds." Chem. Eur. J. 2000, 6, 581. 152. Soler, M.; Wernsdorfer, W.; Folting, K.; Pink, M.; Christou, G., "Single-molecule

142. Gatteschi, D.; Sessoli, R.; Villain, J., Molecular Nanomagnets. OUP Oxford: 2006.

2010, 2, 362. 151. Nelsen,

magnets:

a

Large

Mn30

molecular

nanomagnet exhibiting quantum tunneling of magnetization."

J. Am. Chem. Soc. 2004, 126, 2156.

ACS Paragon Plus Environment

Page 19 of 19 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

Journal of the American Chemical Society

19 ACS Paragon Plus Environment