John F. Berry

Find an error

Name: Berry, John F.
Organization: University of Wisconsin—Madison , USA
Department: Department of Chemistry
Title: Professor(PhD)

TOPICS

Co-reporter:John F. Berry and Connie C. Lu
Inorganic Chemistry July 17, 2017 Volume 56(Issue 14) pp:7577-7577
Publication Date(Web):July 17, 2017
DOI:10.1021/acs.inorgchem.7b01330
Co-reporter:Amanda R. Corcos, Michael D. Roy, Michelle M. Killian, Stephanie Dillon, Thomas C. Brunold, and John F. Berry
Inorganic Chemistry December 4, 2017 Volume 56(Issue 23) pp:14662-14662
Publication Date(Web):November 15, 2017
DOI:10.1021/acs.inorgchem.7b02557
The electronic structures of the diruthenium compounds Ru2(ap)4Cl (1, ap = 2-anilinopyridinate) and Ru2(ap)4OTf (2) were investigated with UV–vis, resonance Raman, and magnetic circular dichroism (MCD) spectroscopies; SQUID magnetometry; and density functional theory (DFT) calculations. Both compounds have quartet spin ground states with large axial zero-field splitting of ∼60 cm–1 that is characteristic of Ru25+ compounds having a (π*, δ*)3 electron configuration and a Ru–Ru bond order of ∼2.5. Two major visible absorption features are observed at ∼770 and 430 nm in the electronic spectra, the assignments of which have previously been ambiguous. Both bands have significant charge-transfer character with some contributions from d → d transitions. MCD spectra were measured to enable the identification of d → d transitions that are not easily observable by UV–vis spectroscopy. In this way, we are able to identify bands due to δ → δ* and δ → π* transitions at ∼16 100 and 11 200–12 300 cm–1, respectively, the latter band being sensitive to the π-donating character of the axial ligand. The Ru–Ru stretches are coupled with pyridine rocking motions and give rise to observed resonance Raman peaks at ∼350 and 420 cm–1, respectively.
Co-reporter:Cale Weatherly, Juliet M. Alderson, John F. Berry, Jason E. Hein, and Jennifer M. Schomaker
Organometallics April 24, 2017 Volume 36(Issue 8) pp:1649-1649
Publication Date(Web):April 3, 2017
DOI:10.1021/acs.organomet.7b00190
Catalyst-controlled, selective nitrene transfer is often challenging when both C–H and C═C bonds are present in a substrate. Interestingly, a simple change in the Ag(I):L ratio (L = bidentate N,N-donor ligand) enables tunable, chemoselective nitrene transfer that favors either C═C bond aziridination using an ∼1:1 Ag:L ratio (AgLOTf) or insertion into a C–H bond when the Ag:L ratio in the catalyst is 1:2 (AgL2OTf). In this paper, mechanistic studies, coupled with kinetic profiling of the entire reaction course, are employed to examine the reasons for this unusual behavior. Steady-state kinetics were found to be similar for both AgLOTf and AgL2OTf; both complexes yield electronically similar reactive intermediates that engage in nitrene transfer involving formation of a short-lived radical intermediate and barrierless radical recombination. Taken together, experimental and computational studies point to two effects that control tunable chemoselectivity: suppression of aziridination as the steric congestion around the silver center is increased in AgL2OTf and a decrease in the rate of C–H insertion with AgLOTf in comparison to AgL2OTf. The observation that the sterics of Ag catalysts can be varied, with minor effects on the electronic features of the putative nitrene, has important implications for the development of other silver catalysts that enable tunable, site-selective C–H bond aminations.
Co-reporter:Sungho V. Park
Dalton Transactions 2017 vol. 46(Issue 28) pp:9118-9125
Publication Date(Web):2017/07/18
DOI:10.1039/C7DT01847A
A series of RuII complexes stabilized with the pentapyridyl ligand Py5Me2 (Py5Me2 = 2,6-bis(1,1-bis(2-pyridyl)ethyl)pyridine) and with an axial X ligand (X = Cl−, H2O, N3−, MeCN) were prepared and characterized in the solid state and in non-aqueous solution. The cyclic voltammograms of these complexes in MeCN reflect a reversible substitution of the axial X ligand with MeCN. Irreversible ligand substitution of [(Py5Me2)RuN3]+ is also observed in propylene carbonate, but only at oxidizing potentials that decompose the azide ligand. The monometallic chloride and azide species are compared with analogous Ru2 metal–metal bonded complexes, which have been reported to undergo irreversible chloride dissociation upon reduction.
Co-reporter:Amanda R. Corcos
Dalton Transactions 2017 vol. 46(Issue 17) pp:5532-5539
Publication Date(Web):2017/05/02
DOI:10.1039/C6DT04328C
Five new metal–metal bonded Ru2 compounds are presented and discussed: Ru2(ap)4ONO2 (2), [Ru2(ap)4NCMe][BF4] (3), Ru2(ap)4FBF3 (4), Ru2(ap)4OTf (5), and [Ru2(ap)4OTf][Ag(OTf)2] (6) (ap = 2-anilinopyridinate). All compounds have a (4,0) arrangement of the ap ligands about the Ru–Ru bond and contain one sterically blocked axial site and one site containing a labile ligand. These compounds display some of the shortest Ru–Ru distances known for this class of compounds. We demonstrate a reversible interconversion between compounds 3 and 4 as the MeCN and BF4− ligands are readily displaced. Despite the presence of labile axial ligands, compounds 2–5 remain high spin with an S = 3/2 ground state as determined by EPR spectroscopy.
Co-reporter:John F. Berry
Accounts of Chemical Research 2016 Volume 49(Issue 1) pp:27
Publication Date(Web):January 7, 2016
DOI:10.1021/acs.accounts.5b00517
First proposed in a classic Linus Pauling paper, the two-center/three-electron (2c/3e) σ half-bond challenges the extremes of what may or may not be considered a chemical bond. Two electrons occupying a σ bonding orbital and one electron occupying the antibonding σ* orbital results in bond orders of ∼0.5 that are characteristic of metastable and exotic species, epitomized in the fleetingly stable He2+ ion.In this Account, I describe the use of coordination chemistry to stabilize such fugacious three-electron bonded species at disparate ends of the periodic table. A recent emphasis in the chemistry of metal–metal bonds has been to prepare compounds with extremely short metal–metal distances and high metal–metal bond orders. But similar chemistry can be used to explore metal–metal bond orders less than one, including 2c/3e half-bonds. Bimetallic compounds in the Ni2(II,III) and Pd2(II,III) oxidation states were originally examined in the 1980s, but the evidence collected at that time suggested that they did not contain 2c/3e σ bonds. Both classes of compounds have been re-examined using EPR spectroscopy and modern computational methods that show the unpaired electron of each compound to occupy a M–M σ* orbital, consistent with 2c/3e Ni–Ni and Pd–Pd σ half-bonds.Elsewhere on the periodic table, a seemingly unrelated compound containing a trigonal bipyramidal Cu3S2 core caused a stir, leaving prominent theorists at odds with one another as to whether the compound contains a S–S bond. Due to my previous experience with 2c/3e metal–metal bonds, I suggested that the Cu3S2 compound could contain a 2c/3e S–S σ half-bond in the previously unknown oxidation state of S23–. By use of the Cambridge Database, a number of other known compounds were identified as potentially containing S23– ligands, including a noteworthy set of cyclopentadienyl-supported compounds possessing diamond-shaped Ni2E2 units with E = S, Se, and Te. These compounds were subjected to extensive studies using X-ray absorption spectroscopy, X-ray photoelectron spectroscopy, density functional theory, and wave function-based computational methods, as well as chemical oxidation and reduction. The compounds contain E–E 2c/3e σ half-bonds and unprecedented E23– “subchalcogenide” ligands, ushering in a new oxidation state paradigm for transition metal–chalcogen chemistry.
Co-reporter:Amanda R. Corcos, József S. Pap, Tzuhsiung Yang, and John F. Berry
Journal of the American Chemical Society 2016 Volume 138(Issue 31) pp:10032-10040
Publication Date(Web):July 13, 2016
DOI:10.1021/jacs.6b05942
Three new diruthenium oxyanion complexes have been prepared, crystallographically characterized, and screened for their potential to photochemically unmask a reactive Ru—Ru═O intermediate. The most promising candidate, Ru2(chp)4ONO2 (4, chp = 6-chloro-2-hydroxypyridinate), displays a set of signals centered around m/z = 733 amu in its MALDI-TOF mass spectrum, consistent with the formation of the [Ru2(chp)4O]+ ([6]+) ion. These signals shift to 735 amu in 4*, which contains an 18O-labeled nitrate. EPR spectroscopy and headspace GC-MS analysis indicate that NO2• is released upon photolysis of 4, also consistent with the formation of 6. Photolysis of 4 in CH2Cl2 at room temperature in the presence of excess PPh3 yields OPPh3 in 173% yield; control experiments implicate 6, NO2•, and free NO3– as the active oxidants. Notably, Ru2(chp)4Cl (3) is recovered after photolysis. Since 3 is the direct precursor to 4, the results described herein constitute the first example of a synthetic cycle for oxygen atom transfer that makes use of light to generate a putative metal oxo intermediate.
Co-reporter:Nicholas S. Dolan, Ryan J. Scamp, Tzuhsiung Yang, John F. Berry, and Jennifer M. Schomaker
Journal of the American Chemical Society 2016 Volume 138(Issue 44) pp:14658-14667
Publication Date(Web):October 11, 2016
DOI:10.1021/jacs.6b07981
The development of new catalysts for selective nitrene transfer is a continuing area of interest. In particular, the ability to control the chemoselectivity of intermolecular reactions in the presence of multiple reactive sites has been a long-standing challenge in the field. In this paper, we demonstrate examples of silver-catalyzed, nondirected, intermolecular nitrene transfer reactions that are both chemoselective and flexible for aziridination or C–H insertion, depending on the choice of ligand. Experimental probes present a puzzling picture of the mechanistic details of the pathways mediated by [(tBu3tpy)AgOTf]2 and (tpa)AgOTf. Computational studies elucidate these subtleties and provide guidance for the future development of new catalysts exhibiting improved tunability in group transfer reactions.
Co-reporter:Adrián Varela-Álvarez; Tzuhsiung Yang; Heather Jennings; Katherine P. Kornecki; Samantha N. Macmillan; Kyle M. Lancaster; James B. C. Mack; J. Du Bois; John F. Berry;Djamaladdin G. Musaev
Journal of the American Chemical Society 2016 Volume 138(Issue 7) pp:2327-2341
Publication Date(Web):January 28, 2016
DOI:10.1021/jacs.5b12790
Dirhodium-catalyzed C–H amination is hypothesized to proceed via Rh2-nitrene intermediates in either the Rh2(II,II) or Rh2(II,III) redox state. Herein, we report joint theoretical and experimental studies of the ground electronic state (GES), redox potentials, and C–H amination of [Rh2II,III(O2CCH3)4(L)n]+ (1_L) (L = none, Cl–, and H2O), [Rh2(esp)2]+ (2), and Rh2(espn)2Cl (3) (esp = α,α,α′,α′-tetramethyl-1,3-benzenedipropanoate and espn = α,α,α′,α′-tetramethyl-1,3-benzenedipropanamidate). CASSCF calculations on 1_L yield a wave function with two closely weighted configurations, (δ*)2(π1*)2(π2*)1 and (δ*)2(π1*)1(π2*)2, consistent with reported EPR g values [ Chem. Phys. Lett. 1986, 130, 20−23]. In contrast, EPR spectra of 2 show g values consistent with the DFT-computed (π*)4(δ*)1 GES. EPR spectra and Cl K-edge XAS for 3 are consistent with a (π*)4(δ*)1 GES, as supported by DFT. Nitrene intermediates 2N_L and 3N_L are also examined by DFT (the nitrene is an NSO3R species). DFT calculations suggest a doublet GES for 2N_L and a quartet GES for 3N_L. CASSCF calculations describe the GES of 2N as Rh2(II,II) with a coordinated nitrene radical cation, (π*)4(δ*)2(πnitrene,1)1(πnitrene,2)0. Conversely, the GES of 3N is Rh2(II,III) with a coordinated triplet nitrene, (π*)4(δ*)1(πnitrene,1)1(πnitrene,2)1. Quartet transition states (4TSs) are found to react via a stepwise radical mechanism, whereas 2TSs are found to react via a concerted mechanism that is lower in energy compared to 4TSs for both 2N_L and 3N_L. The experimental (determined by intramolecular competition) and 2TS-calculated kinetic isotopic effect (KIE) shows a KIE ∼ 3 for both 2N and 3N, which is consistent with a concerted mechanism.
Co-reporter:Amanda R. Corcos; Omar Villanueva; Richard C. Walroth; Savita K. Sharma; John Bacsa; Kyle M. Lancaster; Cora E. MacBeth
Journal of the American Chemical Society 2016 Volume 138(Issue 6) pp:1796-1799
Publication Date(Web):January 22, 2016
DOI:10.1021/jacs.5b12643
Bimetallic (Et4N)2[Co2(L)2], (Et4N)2[1] (where (L)3– = (N(o-PhNC(O)iPr)2)3–) reacts with 2 equiv of O2 to form the monometallic species (Et4N)[Co(L)O2], (Et4N)[3]. A crystallographically characterized analog (Et4N)2[Co(L)CN], (Et4N)2[2], gives insight into the structure of [3]1–. Magnetic measurements indicate [2]2– to be an unusual high-spin CoII-cyano species (S = 3/2), while IR, EXAFS, and EPR spectroscopies indicate [3]1– to be an end-on superoxide complex with an S = 1/2 ground state. By X-ray spectroscopy and calculations, [3]1– features a high-spin CoII center; the net S = 1/2 spin state arises after the Co electrons couple to both the O2•– and the aminyl radical on redox non-innocent (L•)2–. Dianion [1]2– shows both nucleophilic and electrophilic catalytic reactivity upon activation of O2 due to the presence of both a high-energy, filled O2– π* orbital and an empty low-lying O2– π* orbital in [3]1–.
Co-reporter:Jonathan H. Christian, David W. Brogden, Jasleen K. Bindra, Jared S. Kinyon, Johan van Tol, Jingfang Wang, John F. Berry, and Naresh S. Dalal
Inorganic Chemistry 2016 Volume 55(Issue 13) pp:6376
Publication Date(Web):February 16, 2016
DOI:10.1021/acs.inorgchem.5b02545
Magnetic properties of the series of three linear, trimetallic chain compounds Cr2Cr(dpa)4Cl2, 1, Mo2Cr(dpa)4Cl2, 2, and W2Cr(dpa)4Cl2, 3 (dpa = 2,2′-dipyridylamido), have been studied using variable-temperature dc and ac magnetometry and high-frequency EPR spectroscopy. All three compounds possess an S = 2 electronic ground state arising from the terminal Cr2+ ion, which exhibits slow magnetic relaxation under an applied magnetic field, as evidenced by ac magnetic susceptibility and magnetization measurements. The slow relaxation stems from the existence of an easy-axis magnetic anisotropy, which is bolstered by the axial symmetry of the compounds and has been quantified through rigorous high-frequency EPR measurements. The magnitude of D in these compounds increases when heavier ions are substituted into the trimetallic chain; thus D = −1.640, −2.187, and −3.617 cm–1 for Cr2Cr(dpa)4Cl2, Mo2Cr(dpa)4Cl2, and W2Cr(dpa)4Cl2, respectively. Additionally, the D value measured for W2Cr(dpa)4Cl2 is the largest yet reported for a high-spin Cr2+ system. While earlier studies have demonstrated that ligands containing heavy atoms can enhance magnetic anisotropy, this is the first report of this phenomenon using heavy metal atoms as “ligands”.
Co-reporter:Travis L. Sunderland and John F. Berry  
Dalton Transactions 2016 vol. 45(Issue 1) pp:50-55
Publication Date(Web):19 Nov 2015
DOI:10.1039/C5DT03740A
Five novel homoleptic heterobimetallic bismuth(II)–rhodium(II) carboxylate complexes—BiRh(TPA)4 (1), BiRh(but)4 (2), BiRh(piv)4 (3), BiRh(esp)2 (4), and BiRh(OAc)4 (5)—were synthesized in good yields by equatorial ligand substitution starting from BiRh(TFA)4 (TPA = triphenylacetate, but = butyrate, piv = pivalate, esp = α,α,α′,α′-tetramethyl-1,3-benzenedipropionate, OAc = acetate, and TFA = trifluoroacetate). We report here 1H and 13C{1H} NMR spectra and cyclic voltammograms for complexes 1–4, and IR spectra for all complexes. Irreversible redox waves appear between −1.4 to −1.5 V for [BiRh]3+/4+ couples and 1.3 to 1.5 V vs. Fc/Fc+ for [BiRh]4+/5+ couples for complexes 1–4 indicating a wide range of stability for the compounds. The X-ray crystal structure of 1 reveals a Bi–Rh distance of 2.53 Å.
Co-reporter:Amanda R. Corcos and John F. Berry  
Dalton Transactions 2016 vol. 45(Issue 6) pp:2386-2389
Publication Date(Web):04 Jan 2016
DOI:10.1039/C5DT04875C
The complex {[Ru2(ap)4]2[AgF2]}[BF4]3 ({2}[BF4]3, ap = 2-anilinopyridine), containing the [AgF2]− anion ligated to two [Ru2]6+ cores, is prepared, characterized, and compared to dimeric dumbbell-type structures, monomeric Ru2 structures, as well as the known set of dihalo coinage-metalate anions. X-ray crystallography indicates that the Ru–Ru and Ru–F distances are rather short, 2.2835(3) Å and 2.054(1) Å, respectively, while the Ag–F distance of 2.274(1) Å is longer than that calculated for the free/un-ligated anion. Cyclic voltammetry in dichloromethane indicates that, while some of {2}3+ breaks apart into an [Ru2(ap)4F]+ ([3]+) monomer in solution, the remaining dimer has a single reversible two-electron redox feature for the Ru26/5+ couple that is at a lower potential than that of [3]+. This is one of the few examples of a ligated dihalo coinage-metalate, and it is the first example of a coinage metal difluoride anion, either free or ligated.
Co-reporter:Brian S. Dolinar, John F. Berry
Polyhedron 2016 Volume 103(Part A) pp:71-78
Publication Date(Web):8 January 2016
DOI:10.1016/j.poly.2015.09.028
The syntheses, X-ray structural characterizations, and electrochemical properties of four new paddlewheel [MMo2(SNO5)4Cl](n−1)+ (Mn+ = Na+, Ca2+, Sr2+, Y3+; HSNO5 = monothiosuccinimide) compounds are described here. By changing the Mn+ cation charge and size, we demonstrate a method for easily tuning the Lewis acidity of the MoMo quadruple bond and the [Mo2]4+/5+ redox potential. As the charge of the Mn+ cation is increased, the distal Mo atom becomes more Lewis acidic, leading to a shorter Mo2–Cl bond distance, and the [Mo2]4+/5+ redox couple becomes less accessible. As the Mn+⋯Mo2 distance is increased, the distal Mo atom becomes less Lewis acidic, leading to a longer Mo2–Cl bond distance.The syntheses, X-ray structural characterizations, and electrochemical properties of four new paddlewheel [MMo2(SNO5)4Cl]n−1+ (Mn+ = Na+, Ca2+, Sr2+, Y3+; HSNO5 = monothiosuccinimide) compounds are described here. By changing the Mn+ cation charge density, the Lewis acidity of the MoMo quadruple bond and the [Mo2]4+/5+ redox potential are tuned.
Co-reporter:Shu A. Yao; Vlad Martin-Diaconescu; Ivan Infante; Kyle M. Lancaster; Andreas W. Götz; Serena DeBeer
Journal of the American Chemical Society 2015 Volume 137(Issue 15) pp:4993-5011
Publication Date(Web):March 21, 2015
DOI:10.1021/ja511607j
The diamagnetic compounds Cp′2Ni2E2 (1: E = S, 2: E = Se, 3: E = Te; Cp′ = 1,2,3,4,-tetraisopropylcyclopentadienyl), first reported by Sitzmann and co-workers in 2001 [Sitzmann, H.; Saurenz, D.; Wolmershauser, G.; Klein, A.; Boese, R. Organometallics 2001, 20, 700], have unusual E···E distances, leading to ambiguities in how to best describe their electronic structure. Three limiting possibilities are considered: case A, in which the compounds contain singly bonded E22– units; case B, in which a three-electron E∴E half-bond exists in a formal E23– unit; case C, in which two E2– ions exist with no formal E–E bond. One-electron reduction of 1 and 2 yields the new compounds [Cp*2Co][Cp′2Ni2E2] (1red: E = S, 2red: E = Se; Cp* = 1,2,3,4,5-pentamethylcyclopentadieyl). Evidence from X-ray crystallography, X-ray absorption spectroscopy, and X-ray photoelectron spectroscopy suggest that reduction of 1 and 2 is Ni-centered. Density functional theory (DFT) and ab initio multireference methods (CASSCF) have been used to investigate the electronic structures of 1–3 and indicate covalent bonding of an E23– ligand with a mixed-valent Ni2(II,III) species. Thus, reduction of 1 and 2 yields Ni2(II,II) species 1red and 2red that bear unchanged E23– ligands. We provide strong computational and experimental evidence, including results from a large survey of data from the Cambridge Structural Database, indicating that M2E2 compounds occur in quantized E2 oxidation states of (2 × E2–), E23–, and E22–, rather than displaying a continuum of variable E–E bonding interactions.
Co-reporter:Timothy C. Berto, Linghong Zhang, Robert J. Hamers, and John F. Berry
ACS Catalysis 2015 Volume 5(Issue 2) pp:703
Publication Date(Web):December 9, 2014
DOI:10.1021/cs501641z
The ability of tetraalkylammonium ions (NR4+) to facilitate CO2 electroreduction at various electrode surfaces has been investigated. Scan rate dependence shows this process to be diffusion-controlled and largely independent of working electrode material. Variation of the R groups on NR4+ is shown to have little effect on the reduction potential, indicating that catalysis via reduced NR4• species, as previously proposed in the literature, is not a viable mechanism for CO2 electroreduction. Rather, CO2 is reduced via outer-sphere electron transfer according to the mechanism put forth by Savéant and co-workers [Lamy, E.; Nadjo, L.; Savéant, J.-M. J. Electroanal. Chem. 1977, 78, 403−407]. This view is supported by a full analysis of the reaction products, which consist exclusively of CO and CO32–, as well as solvent decomposition products formed at the counter electrode. No degradation of NR4+ ions is observed, which rules out the transient formation of NR4• radical species during electroreduction. However, Li+ ions are definitively shown to inhibit CO2 reduction, even in the presence of NR4+ ions. This inhibition likely occurs via surface adsorption of the Li+ ions.Keywords: boron-doped diamond; carbon dioxide; electrochemistry; tetraalkylammonium
Co-reporter:David W. Brogden and John F. Berry  
Chemical Communications 2015 vol. 51(Issue 44) pp:9153-9156
Publication Date(Web):30 Apr 2015
DOI:10.1039/C5CC02917A
The location of the unpaired electron in the new mixed-valent (W2)IV,V trication [W2O(dpa)4]3+ presents a challenge for DFT methods. EPR spectroscopy confirms the unpaired electron to be in the W(V)–oxo unit, in agreement with the predictions of hybrid functionals B3LYP and TPSSh, but contrary to the predictions of non-hybrid functionals.
Co-reporter:Tristan R. Brown; Brian S. Dolinar; Elizabeth A. Hillard; Rodolphe Clérac
Inorganic Chemistry 2015 Volume 54(Issue 17) pp:8571-8589
Publication Date(Web):August 10, 2015
DOI:10.1021/acs.inorgchem.5b01241
Reduction of (4,0)-Ru2(chp)4Cl (1) (chp = 6-chloro-2-oxypyridinate) with Zn or FeCl2 yields a series of axial ligand adducts of the Ru2(II,II) species Ru2(chp)4(L), with L = tetrahydrofuran (2), dimethyl sulfoxide (DMSO; 3), PPh3 (4), pyridine (5), or MeCN (6). Zn reduction in noncoordinating solvents such as toluene or CH2Cl2 leads to the dimeric species [Ru2(chp)4]2 (7) or [Ru2(chp)4]2(ZnCl2) (8), whereas addition of strongly σ-donating ligands such as CO causes cleavage of the Ru–Ru bond. Density functional theory (DFT) models of these complexes, the axially free species, and the axial adducts of several other potential ligands (H2O, NH3, CH2Cl2, S-bound DMSO, N2, and CO) indicate that these compounds can be divided into three distinct categories, based on their Ru–Ru bond length and electronic structure. Compounds 2, 3, 5, 6, 7, and 8, the hypothetical axially free species, and adducts of H2O and NH3 fit in Category 1 with a (δ*)2(π*)2 ground state, as indicated by their electronic spectra, magnetic properties, and Ru–Ru bond distances. However, compound 4 and the CH2Cl2 adduct (Category 2) show a pseudo-Jahn-Teller distortion and spectroscopic signs of δ*/π* orbital mixing suggestive of a new electronic ground state intermediate between the (δ*)2(π*)2 and (δ*)1(π*)3 configurations. Category 3 consists of the hypothetical adducts of N2, S-bound DMSO, and CO, all of which are predicted to have a (δ*)1(π*)3 configuration. Electronic spectra were recorded and assigned using time-dependent DFT, allowing assignment of a band in the 10 000–13 000 cm–1 range as the δ → π* transition. The axial ligand’s π-acid character heavily influences the δ*−π* gap, and thereby the ground-state electronic configuration, but not the axial ligand binding strength, which is dictated more by the σ-donor character of the ligands. Thus, this work greatly expands the number of axial ligand adducts known for Ru2(II,II) complexes supported by N,O-donor ligands and provides a predictive theoretical framework for their stability and electronic structures.
Co-reporter:Evan Warzecha; Timothy C. Berto
Inorganic Chemistry 2015 Volume 54(Issue 17) pp:8817-8824
Publication Date(Web):August 26, 2015
DOI:10.1021/acs.inorgchem.5b01532
The compound Rh2(esp)2 (esp = α,α,α′,α′-tetramethyl-1,3-benzenediproponoate) is the most generally effective catalyst for nitrenoid amination of C–H bonds. However, much of its fundamental coordination chemistry is unknown. In this work, we study the effects of axial ligand coordination to the catalyst Rh2(esp)2. We report here crystal structures, cyclic voltammetry, UV–vis, IR, Raman, and 1H NMR spectra for the complexes Rh2(esp)2L2 where L = pyridine, 3-picoline, 2,6-lutidine, acetonitrile, and methanol. The compounds all show well-defined π* → σ* electronic transitions in the 16500 to 20500 cm–1 range, and Rh–Rh stretching vibrations in the range from 304 to 322 cm–1. Taking these data into account we find that the strength of axial ligand binding to Rh2(esp)2 increases in the series CH3OH ∼ 2,6-lutidine < CH3CN < 3-methylpyridine ∼ pyridine. Quasi-reversible Rh24+/5+ redox waves are only obtained when either acetonitrile or no axial ligand is present. In the presence of pyridines, irreversible oxidation waves are observed, suggesting that these ligands destabilize the Rh2 complex under oxidative conditions.
Co-reporter:David W. Brogden, Jonathan H. Christian, Naresh S. Dalal, John F. Berry
Inorganica Chimica Acta 2015 Volume 424() pp:241-247
Publication Date(Web):1 January 2015
DOI:10.1016/j.ica.2014.08.020
•Preparation of Mo2Cr(dpa)4Cl2 and W2Cr(dpa)4Cl2 with short metal–metal distances.•Magnetic susceptibility indicates S = 2 states for the compounds.•Calculations reveal a 3c/3e partial bond between the M2 unit and Cr.•TDDFT indicates a δ → δ∗ transition with M2-to-Cr charge transfer character.Metalation of the quadruply bonded Mo2(dpa)4 and W2(dpa)4 (dpa = 2,2′-dipyridylamido) compounds with CrCl2 results in the formation of the new heterotrimetallic complexes Mo2Cr(dpa)4Cl2 (1) and W2Cr(dpa)4Cl2 (2). X-ray crystal structures of 1 and 2 reveal short Mo2⋯Cr and W2⋯Cr distances of 2.69 and 2.65 Å, respectively. Electronic absorption spectra and electrochemical data for 1 and 2 are presented. Variable temperature magnetic susceptibility measurements establish the electronic spin states of 1 and 2, which are S = 2. Molecular orbital analysis by density functional theory (DFT) calculations reveals a 3-center/3-electron sigma bonding configuration leading to a partial sigma type interaction between the M2 quadruply bonded unit and Cr in 1 and 2. Time-dependent DFT provides support for the assignment of the lowest energy feature in the electronic absorption spectra of 1 and 2 as a δ → δ∗ transition with M2-to-Cr charge transfer character.The new heterotrimetallic complexes Mo2Cr(dpa)4Cl2 (1) and W2Cr(dpa)4Cl2 (2) are described, having short Mo2⋯Cr and W2⋯Cr distances of 2.69 and 2.65 Å, respectively, and electronic spin states of S = 2. Molecular orbital analysis by density functional theory (DFT) calculations reveals a 3-center/3-electron sigma bonding configuration leading to a partial sigma type interaction between the M2 quadruply bonded unit and Cr in 1 and 2. Time-dependent DFT provides support for the assignment of the lowest energy feature in the electronic absorption spectra of 1 and 2 as a δ → δ∗ transition with M2-to-Cr charge transfer character.
Co-reporter:JOHN F BERRY
Journal of Chemical Sciences 2015 Volume 127( Issue 2) pp:209-214
Publication Date(Web):2015 February
DOI:10.1007/s12039-015-0773-6
Metal–metal bonded Rh2 and Ru2 complexes having a paddlewheel-type structure are exceptional catalysts for a broad range of organic transformations. I review here the recent efforts towards the observation and characterization of intermediates in these reactions that have previously eluded detection. Specifically, mechanistic investigations of carbenoid and nitrenoid reactions of Rh2(II,II)-tetracarboxylate compounds have led to the observation of a metastable Rh2(II,II) carbene complex as well as a mixed-valent Rh2(II,III)-amido intermediate. Related Ru2 nitrido compounds have been studied and found to undergo intramolecular C–H amination reactions as well as intermolecular reaction with triphenylphosphine.
Co-reporter:Shu A. Yao ; Amanda R. Corcos ; Ivan Infante ; Elizabeth A. Hillard ; Rodolphe Clérac
Journal of the American Chemical Society 2014 Volume 136(Issue 39) pp:13538-13541
Publication Date(Web):September 10, 2014
DOI:10.1021/ja507342a
The nickel hydride complex [Cp′Ni(μ-H)]2 (1, Cp′ = 1,2,3,4-tetraisopropylcyclopentadienyl) is found to have a strikingly short Ni–Ni distance of 2.28638(3) Å. Variable temperature and field magnetic measurements indicate an unexpected triplet ground state for 1 with a large zero-field splitting of +90 K (63 cm–1). Electronic structure calculations (DFT and CASSCF/CASPT2) explain this ground state as arising from half occupation of two nearly degenerate Ni–Ni π* orbitals.
Co-reporter:David W. Brogden, Yevgeniya Turov, Michael Nippe, Giovanni Li Manni, Elizabeth A. Hillard, Rodolphe Clérac, Laura Gagliardi, and John F. Berry
Inorganic Chemistry 2014 Volume 53(Issue 9) pp:4777-4790
Publication Date(Web):April 21, 2014
DOI:10.1021/ic5007204
Oxidation of quadruply bonded Cr2(dpa)4, Mo2(dpa)4, MoW(dpa)4, and W2(dpa)4 (dpa = 2,2′-dipyridylamido) with 2 equiv of silver(I) triflate or ferrocenium triflate results in the formation of the two-electron-oxidized products [Cr2(dpa)4]2+ (1), [Mo2(dpa)4]2+ (2), [MoW(dpa)4]2+ (3), and [W2(dpa)4]2+ (4). Additional two-electron oxidation and oxygen atom transfer by m-chloroperoxybenzoic acid results in the formation of the corresponding metal–oxo compounds [Mo2O(dpa)4]2+ (5), [WMoO(dpa)4]2+ (6), and [W2O(dpa)4]2+ (7), which feature an unusual linear M···M≡O structure. Crystallographic studies of the two-electron-oxidized products 2, 3, and 4, which have the appropriate number of orbitals and electrons to form metal–metal triple bonds, show bond distances much longer (by >0.5 Å) than those in established triply bonded compounds, but these compounds are nonetheless diamagnetic. In contrast, the Cr–Cr bond is completely severed in 1, and the resulting two isolated Cr3+ magnetic centers couple antiferromagnetically with J/kB= −108(3) K [−75(2) cm–1], as determined by modeling of the temperature dependence of the magnetic susceptibility. Density functional theory (DFT) and multiconfigurational methods (CASSCF/CASPT2) provide support for “stretched” and weak metal–metal triple bonds in 2, 3, and 4. The metal–metal distances in the metal–oxo compounds 5, 6, and 7 are elongated beyond the single-bond covalent radii of the metal atoms. DFT and CASSCF/CASPT2 calculations suggest that the metal atoms have minimal interaction; the electronic structure of these complexes is used to rationalize their multielectron redox reactivity.
Co-reporter:David W. Brogden and John F. Berry
Inorganic Chemistry 2014 Volume 53(Issue 21) pp:11354-11356
Publication Date(Web):October 10, 2014
DOI:10.1021/ic5024218
Co-reporter:Brian S. Dolinar and John F. Berry  
Dalton Transactions 2014 vol. 43(Issue 16) pp:6165-6176
Publication Date(Web):19 Feb 2014
DOI:10.1039/C4DT00297K
We report an exploration of the coordination chemistry of a systematic series of cyclic thioamidate ligands with the quadruply-bonded Mo24+ core. In addition to the S and N donor atoms that bind to Mo, the ligands utilized in this study have an additional O or S atom in conjugation with the thioamidate π system. The preparation of four new Mo2 complexes is described, and these compounds are characterized by X-ray crystallography, NMR and UV-vis spectroscopy, electrochemistry, and DFT calculations. These complexes provide a means to interrogate the electronics of Mo2(thioamidate)4 systems. Notably, we describe the first two examples of Mo2(thioamidate)4 complexes in their cis-2,2-regioisomer. By varying the π-system substituent and regioisomerism of these compounds, the electronics of the dimolybdenum core is shown to be altered with varying degrees of effect. Cyclic voltammetry results show that changing the π-system substituent from O to S results in an increase in the Mo24+/5+ oxidation potential by 170 mV. Changing the arrangement of ligands around the dimolybdenum core from trans-2,2 to cis-2,2 slightly weakens the metal–ligand bonds, raising the oxidation potential by a more modest 30–100 mV. MO diagrams of each compound derived from DFT calculations support these conclusions as well; the identity of the π-system substituent alters the δ–δ* (HOMO–LUMO) gap by up to 0.4 eV, whereas regioisomerism yields smaller changes in the electronic structure.
Co-reporter:Brian S. Dolinar and John F. Berry
Inorganic Chemistry 2013 Volume 52(Issue 8) pp:4658-4667
Publication Date(Web):March 29, 2013
DOI:10.1021/ic400275x
We report here the syntheses, X-ray crystal structures, electrochemistry, and density functional theory (DFT) single-point calculations of three new complexes: tetrakis(monothiosuccinimidato)dimolybdenum(II) [Mo2(SNO5)4, 1a], tetrakis(6-thioxo-2-piperidinonato)dimolybdenum(II) [Mo2(SNO6)4, 1b], and chlorotetrakis(monothiosuccinimidato)pyridinelithiumdimolybdenum(II) [pyLiMo2(SNO5)4Cl, 2-py]. X-ray crystallography shows unusually short axial Mo2–Cl bond lengths in 2-py, 2.6533(6) Å, and dimeric 2-dim, 2.644(1) Å, which we propose result from an increased Lewis acidity of the Mo2 unit in the presence of the proximal Li+ ion. When 2-py is dissolved in MeCN, the lithium reversibly dissociates, forming an equilibrium mixture of (MeCNLiMo2(SNO5)4Cl) (2-MeCN) and [Li(MeCN)4]+[Mo2(SNO5)4Cl]− (3). Cyclic voltammetry was used to determine the equilibrium lithium binding constant (room temperature, Keq = 95 ± 1). From analysis of the temperature dependence of the equilibrium constant, thermodynamic parameters for the formation of 2-MeCN from 3 (ΔH° = −6.96 ± 0.93 kJ mol–1 and ΔS° = 13.9 ± 3.5 J mol–1 K–1) were extracted. DFT calculations indicate that Li+ affects the Mo–Cl bond length through polarization of metal–metal bonding/antibonding molecular orbitals when lithium and chloride are added to the dimolybdenum core.
Co-reporter:Ama R. Corcos;Ama Kae Musch Long;Ilia A. Guzei
European Journal of Inorganic Chemistry 2013 Volume 2013( Issue 22-23) pp:3808-3811
Publication Date(Web):
DOI:10.1002/ejic.201300180

Abstract

A new Ru2 azido complex, [Ru2(chp)4N3] (4, chp = 2-chloro-6-hydroxypyridinate), was investigated under photolytic conditions to study the chemical reactivity of the corresponding Ru2 nitride species, [Ru2(chp)4N] (6), towards intermolecular N atom transfer to triphenylphosphane (PPh3). Photolysis of a dichloromethane solution of 4 at λ > 350 nm leads to a characteristic color change from purple to magenta. Upon acidic workup, triphenylphosphanamine chloride ([H2NPPh3]Cl) is produced and [Ru2(chp)4Cl] (5), the precursor to 4, is regenerated. The first stoichiometric cycle for intermolecular N atom transfer from a Ru2 nitride is thus presented.

Co-reporter:Shu A. Yao, Christopher B. Hansen, John F. Berry
Polyhedron 2013 Volume 58() pp:2-6
Publication Date(Web):13 July 2013
DOI:10.1016/j.poly.2012.05.038
Presented is a quick and efficient method for the insertion of first-row transition metal acetates into 5,10,15,20-tetraphenyl porphyrin, utilizing Soxhlet extraction for the removal of the acetic acid by-product. High yields (>90%) and purity are achieved without requiring chromatographic purification. Basic Mn(III) and Fe(III) acetate may be used, as well as hydrated Co(II), Ni(II), and Cu(II) acetates as starting materials. In the case of Fe, an initial mixture of Fe(TPP)OAc and [Fe(TPP)]2O is formed, which can be quantitatively converted to Fe(TPP)OAc by treatment with acetic acid. The crystal structure of Fe(TPP)OAc is reported, and spectral data for all compounds is presented, including a correction of the literature UV–Vis data for Fe(TPP)OAc. Attempted metalation with basic Cr(III) acetate yielded a mixture of Cr(TPP)OAc and the oxo-bridged dimer. The latter cannot be converted to the desired acetate. No reaction occurred with vanadyl or titanyl acetate under the conditions used.Graphical abstractA new high-yielding synthetic procedure for the metalation of porphyrins is described. First-row transition metal acetates may be inserted into 5,10,15,20-tetraphyenyl porphyrin utilizing Soxhlet extraction for the removal of the acetic acid by-product. Basic Mn(III) and Fe(III) acetate may be used, as well as hydrated Co(II), Ni(II), and Cu(II) acetates as starting materials. High yields (>90%) and purity are achieved without requiring chromatographic purification.Highlights► A convenient method for the metalation of porphyrins. ► The crystal structure of FeIII(TPP)OAc. ► Corrected UV–Vis data for FeIII(TPP)OAc.
Co-reporter:Shu A. Yao ; Rose E. Ruther ; Linghong Zhang ; Ryan A. Franking ; Robert J. Hamers
Journal of the American Chemical Society 2012 Volume 134(Issue 38) pp:15632-15635
Publication Date(Web):September 10, 2012
DOI:10.1021/ja304783j
We report here covalent attachment of a catalytically active cobalt complex onto boron-doped, p-type conductive diamond. Peripheral acetylene groups were appended on a cobalt porphyrin complex, and azide–alkyne cycloaddition was used for covalent linking to a diamond surface decorated with alkyl azides. The functionalized surface was characterized by X-ray photoelectron spectroscopy and Fourier transform IR spectroscopy, and the catalytic activity was characterized using cyclic voltammetry and FTIR. The catalyst-modified diamond surfaces were used as “smart” electrodes exhibiting good stability and electrocatalytic activity for electrochemical reduction of CO2 to CO in acetonitrile solution.
Co-reporter:George H. Timmer and John F. Berry  
Chemical Science 2012 vol. 3(Issue 10) pp:3038-3052
Publication Date(Web):10 Jul 2012
DOI:10.1039/C2SC20688A
Using density functional methods we have modelled the intramolecular electrophilic aryl C–H amination for 15 dimetal nitrides, both homo- and heteronuclear, along with 2 mononuclear nitrides, in the pursuit of understanding the reactivity of the dimetal nitrido Ru2(DPhF)4N (DPhF = N,N′-diphenyl formamidinate) molecule, for which this amination reaction was experimentally observed and characterized. It was found that the 3-center bonding manifold (MMN) that arises between the metal–metal bond and axial nitrido moiety has a dominant influence in the electronic structure and consequently the reactivity at each step in the reaction. It was found that transition state energetics correlate strongly with product stabilization and that these quantities depend on the number of electrons available to occupy the MMN manifold. As the reaction proceeds the number of orbitals in the manifold decreases by one and the point at which this happens determines which of two transition states is rate limiting. The dimetallic nitrides are shown to be inherently more reactive than the mononuclear complexes and so the MMN manifold that is only active in the dimetallic complexes comes through as an important factor in facilitating this amination reaction. Overall, a strong correlation between electronic structure and reactivity is established for C–H amination and new synthetic targets are proposed to develop new facets of this reactivity.
Co-reporter:Katherine P. Kornecki and John F. Berry  
Chemical Communications 2012 vol. 48(Issue 99) pp:12097-12099
Publication Date(Web):31 Oct 2012
DOI:10.1039/C2CC36614B
A new mixed-valent Rh2II,III dimer, [Rh2(espn)2Cl] (espn2− = α,α,α′,α′-tetramethyl-1,3-benzenedipropanamidate), is reported. This compound readily dissociates Cl− at low concentrations in solution to form the active [Rh2(espn)2]+ catalyst, which performs intramolecular C–H amination with TONs > 1400. This work expands the scope of Rh2II,III dimers to nitrenoid chemistry.
Co-reporter:Yevgeniya Turov and John F. Berry  
Dalton Transactions 2012 vol. 41(Issue 26) pp:8153-8161
Publication Date(Web):19 Mar 2012
DOI:10.1039/C2DT30150D
We report here two novel synthetic pathways toward the preparation of a family of trimetallic diazide compounds of the type Cr2M(dpa)4(N3)2, with M = Cr (10), Mn (4), Fe (5), and Co (11). Reaction of either Cr2M(dpa)4(OTf)2 (for M = Mn and Fe) or [Cr2M(dpa)4(MeCN)2](PF6)2 (for M = Cr and Co) with sodium azide in methanol leads to the formation of the corresponding diazide compounds, and single crystal X-ray diffraction measurements confirm the predicted structures. Compounds 4, 5, and 10 are all high-spin compounds, but 11 is a spin-crossover compound exhibiting low-spin behavior at low temperatures (∼100 K). Thermolytic characterization by DSC and TGA reveals an exothermic reaction corresponding to the loss of two dinitrogen molecules from compounds 5, 10, and 11. Further characterization by solution NMR measurements and cyclic voltammetry are also presented.
Co-reporter:Katherine P. Kornecki
European Journal of Inorganic Chemistry 2012 Volume 2012( Issue 3) pp:562-568
Publication Date(Web):
DOI:10.1002/ejic.201100814

Abstract

We report two new analogues of the well-known C–H amination catalyst [Rh2(esp)2] (1) (esp = α,α,α′,α′-tetramethyl-1,3-benzenedipropanoate) that bear redox-active supporting ligands that are structurally similar to esp. The redox-active ligands are 2-[3-(1-carboxy-1-methylethoxy)phenoxy]-2-methylpropanoic acid (H2L1) and (3-methoxycarbonyl-2,5-di-tert-butylphenoxy)ethanoic acid (H2L2), which react with Rh2(OAc)4 to form the catalysts [Rh2(L1)2] (2) and [Rh2(L2)2] (3). Both 2 and 3 have been characterized by X-ray crystallography and cyclic voltammetry, inter alia. Compounds 2 and 3 are structurally similar to 1 but show more complex electrochemical features. Whereas 1 has a single reversible redox wave that corresponds to the Rh2II,II/Rh2II,III couple, 2 and 3 show multiple oxidations that are characteristic of ligand-centered oxidation. Catalysts 1, 2, and 3 perform well in a model intramolecular C–H amination reaction, and all three catalysts perform equally well during the first four hours of a model intermolecular reaction. After this point, 2 and 3 cease to function, whereas 1 continues to be active. These results support the hypothesis that intermolecular C–H amination utilizes two distinct mechanisms: (1) a nitrene interception/insertion mechanism that is fast but ceases to be operative after four hours, and (2) a one-electron mechanism that is more robust over extended time periods, but requires the catalyst to be able to undergo Rh2-centered oxidation.

Co-reporter:Giovanni LiManni;Allison L. Dzubak;Abbas Mulla;David W. Brogden; John F. Berry; Laura Gagliardi
Chemistry - A European Journal 2012 Volume 18( Issue 6) pp:1737-1749
Publication Date(Web):
DOI:10.1002/chem.201103096

Abstract

To gain insights into the trends in metal–metal multiple bonding among the Group 6 elements, density functional theory has been employed in combination with multiconfigurational methods (CASSCF and CASPT2) to investigate a selection of bimetallic, multiply bonded compounds. For the compound [Ar-MM-Ar] (Ar=2,6-(C6H5)2-C6H3, M=Cr, Mo, W) the effect of the Ar ligand on the M2 core has been compared with the analogous [Ph-MM-Ph] (Ph=phenyl, M=Cr, Mo, W) compounds. A set of [M2(dpa)4] (dpa=2,2′-dipyridylamide, M=Cr, Mo, W, U) compounds has also been investigated. All of the compounds studied here show important multiconfigurational behavior. For the Mo2 and W2 compounds, the σ2π4δ2 configuration dominates the ground-state wavefunction, contributing at least 75 %. The Cr2 compounds show a more nuanced electronic structure, with many configurations contributing to the ground state. For the Cr, Mo, and W compounds the electronic absorption spectra have been studied, combining density functional theory and multireference methods to make absorption feature assignments. In all cases, the main features observed in the visible spectra may be assigned as charge-transfer bands. For all compounds investigated the Mayer bond order (MBO) and the effective bond order (EBO) were calculated by density functional theory and CASSCF methods, respectively. The MBO and EBO values share a similar trend toward higher values at shorter normalized metal–metal bond lengths.

Co-reporter:Dr. Shu A. Yao; Kyle M. Lancaster;Dr. Andreas W. Götz; Serena DeBeer; John F. Berry
Chemistry - A European Journal 2012 Volume 18( Issue 30) pp:9179-9183
Publication Date(Web):
DOI:10.1002/chem.201201291
Co-reporter:Amanda Kae Musch Long ; George H. Timmer ; József S. Pap ; Jamie Lynn Snyder ; Renyuan Pony Yu
Journal of the American Chemical Society 2011 Volume 133(Issue 33) pp:13138-13150
Publication Date(Web):July 11, 2011
DOI:10.1021/ja203993p
Diruthenium azido complexes Ru2(DPhF)4N3 (1a, DPhF = N,N′-diphenylformamidinate) and Ru2(D(3,5-Cl2)PhF)4N3 (1b, D(3,5-Cl2)PhF = N,N′-bis(3,5-dichlorophenyl)formamidinate) have been investigated by thermolytic and photolytic experiments to investigate the chemical reactivity of the corresponding diruthenium nitride species. Thermolysis of 1b at ∼100 °C leads to the expulsion of N2 and isolation of Ru2(D(3,5-Cl2)PhF)3NH(C13H6N2Cl4) (3b), in which a nitrogen atom has been inserted into one of the proximal aryl C–H bonds of a D(3,5-Cl2)PhF ligand. A similar C–H insertion product is obtained upon thawing a frozen CH2Cl2 solution of the nitride complex Ru2(DPhF)4N (2a), formed via photolysis at −196 °C of 1a to yield Ru2(DPhF)3NH(C13H10N2) (3a). Evidence is provided here that both reactions proceed via direct intramolecular attack of an electrophilic terminal nitrido nitrogen atom on a proximal aryl ring. Thermodynamic and kinetic data for this reaction are obtained from differential scanning calorimetric measurements and thermal gravimetric analysis of the thermolysis of Ru2(D(3,5-Cl2)PhF)4N3, and by Arrhenius/Eyring analysis of the conversion of Ru2(DPhF)4N to its C–H insertion product, respectively. These data are used to develop a detailed, experimentally validated DFT reaction pathway for N2 extrusion and C–H functionalization from Ru2(D(3,5-Cl2)PhF)4N3. The diruthenium nitrido complex is an intermediate in the calculated reaction pathway, and the C–H functionalization event shares a close resemblance to a classical electrophilic aromatic substitution mechanism.
Co-reporter:Michael Nippe ; Yevgeniya Turov
Inorganic Chemistry 2011 Volume 50(Issue 21) pp:10592-10599
Publication Date(Web):September 20, 2011
DOI:10.1021/ic2011309
The heterometallic complexes CrCrM(dpa)4Cl2 (dpa = 2,2′-dipyridylamide) featuring linear Cl–Cr≡Cr···M–Cl chains can regiospecifically be modified via axial ligand substitution to yield OTf–Cr≡Cr···M–Cl chains (OTf = triflate) with M being Fe, Mn, or Co. The effect of OTf substitution on the Cr side of the molecule has an unusual and profound structural impact on the square-pyramidal transition metal M. Specifically, elongation of the four equatorial M–Npy bonds and the axial M–Cl bonds by 0.03 and 0.09 Å for Fe and 0.07 and 0.11 Å for Mn is observed. The longer M–Cl and M–Npy bonds result from subtle interactions between the equatorial dpa ligand and the three metal ions. The equatorial dpa ligand responds to the introduction of the more labile OTf ligand at Cr by binding more strongly to this Cr ion which in turn weakens bonding to M. The ligand field experienced by M can be tuned by changing the Cr axial ligand, and this effect is observed in electrochemical measurements of the iron compounds.
Co-reporter:Michael Nippe, Eckhard Bill, and John F. Berry
Inorganic Chemistry 2011 Volume 50(Issue 16) pp:7650-7661
Publication Date(Web):July 14, 2011
DOI:10.1021/ic2011315
Binuclear quadruply bonded complexes Cr2(dpa)4 (1, dpa = 2,2′-dipyridylamide), Mo2(dpa)4 (2), and W2(dpa)4 (3) react with anhydrous FeCl2, yielding heterometallic compounds CrCrFe(dpa)4Cl2 (4), MoMoFe(dpa)4Cl2 (5), and WWFe(dpa)4Cl2 (6). These molecules are structurally similar, having a linear MM···Fe chain that is axially capped by chloride ions and is equatorially supported by the helically twisted dpa ligands. A structurally related zinc analog, CrCrZn(dpa)4Cl2 (7), can be prepared upon metalation of 1 with ZnCl2. This reaction also persistently produces a 2:1 adduct of ZnCl2 with 1, [Cr2(dpa)4](ZnCl2)2 (8), which is in equilibrium with 7 and has the two zinc ions bound externally to the Cr2 core and axial bridging chloro ligands attached to each Cr ion. The sole isolable product of the addition of ZnCl2 to 3 is a 1:1 adduct, [W2(dpa)4]ZnCl2 (9). The structurally related chain complexes 4, 5, 6, and 7 are characterized by X-ray crystallography, UV–vis spectroscopy, cyclic voltammetry, and 57Fe Mössbauer spectroscopy for the iron complexes in order to gain insights into the nature of heterometallic interactions, electronic excited states, and redox properties of these compounds, which have implications for all other MM···M′ molecules. Additionally, NMR spectroscopy has been used to gain insight into the mechanism of the metalation of 1 by Zn(II).
Co-reporter:Katherine P. Kornecki ; John F. Berry
Chemistry - A European Journal 2011 Volume 17( Issue 21) pp:5827-5832
Publication Date(Web):
DOI:10.1002/chem.201100708

Abstract

Swift and energy efficient conversion of chemical feedstocks to pharmaceuticals and agrochemicals requires the development of new methods to add nitrogen functionality to unfunctionalized organic substrates. Dirhodium-catalyzed insertion of nitrene species into CH bonds is a promising new method, the main drawback of which is the currently limited understanding of the catalytic mechanism. Herein, cyclic voltammetry and controlled potential electrolysis measurements have enabled us to solve many of the mechanistic mysteries of intermolecular CH amination catalyzed by [Rh2(esp)2] (esp=α,α,α′,α′-tetramethyl-1,3-benzenedipropanoate). The primary result is that, in addition to a simple nitrene-transfer mechanism that dominates the early stages of the reaction, another mechanism is available that relies on sequential proton-coupled electron transfer steps. Whereas the nitrene-transfer mechanism requires the use of expensive, atom-inefficient oxidants, we show that simple one-electron oxidants such as Ce4+ may be used to achieve catalytic CH amination via the one-electron mechanistic regime.

Co-reporter:Amanda Kae Musch Long ; Renyuan Pony Yu ; George H. Timmer
Journal of the American Chemical Society 2010 Volume 132(Issue 35) pp:12228-12230
Publication Date(Web):August 17, 2010
DOI:10.1021/ja1062955
Thermolysis of the terminal azido Ru2(D(3,5-Cl2)PhF)3N3 (3, D(3,5-Cl2)PhF = N,N′-bis(3,5-dichlorophenyl) formamidinate) cleanly produces Ru2[(D(3,5-Cl2)PhF)3(D(3,5-Cl2-2-NH)PhF)] (4), which is proposed to result from insertion of a nitrido N atom into a ligand aryl C−H bond. This mechanism is supported by differential scanning calorimetry and thermogravimetric analysis results, which show the two-step reaction to be exothermic by −215 kJ mol−1, in agreement with results from density functional theory calculations. This is the first example of electrophilic insertion of a terminal nitride into an aromatic C−H bond.
Co-reporter:Michael Nippe ; Jingfang Wang ; Eckhard Bill ; Håkon Hope ; Naresh S. Dalal
Journal of the American Chemical Society 2010 Volume 132(Issue 40) pp:14261-14272
Publication Date(Web):September 22, 2010
DOI:10.1021/ja106510g
Crystal structures of the heterometallic compounds CrCrFe(dpa)4Cl2 (1), CrCrMn(dpa)4Cl2 (2), and MoMoMn(dpa)4Cl2 (3) (dpa = 2,2′-dipyridylamide) show disorder in the metal atom positions such that the linear MAMA···MB array for a given molecule in the crystal is oriented in one of two opposing directions. Despite the fact that the direct coordination sphere of the metals in the two crystallographically independent orientations is identical, subtle differences in some metal−ligand bond distances are observed in 1 and 3 due to differences in the orientation of a solvent molecule of crystallization. The Fe(II) and Mn(II) ions serve as sensitive local spectroscopic probes that have been interrogated by Mössbauer spectroscopy and high-field EPR spectroscopy, respectively. The subtle differences in the two independent Fe and Mn sites in 1 and 3 unexpectedly give rise to unusually large differences in the measured Fe quadrupole splitting (ΔEQ) in 1 and Mn zero-field splitting (D) in 3. Variable-temperature/single-crystal EPR spectroscopy has allowed us to determine that the temperature-dependent D tensors in 3 are oriented along the metal−metal axis and that they show significantly different dynamic behavior with temperature. The differences in ΔEQ and D are reproduced by density functional calculations on truncated models for 1 and 3 that lack the quadruply bonded MAMA groups, though the magnitude of the calculated effect is not as large as that observed experimentally. We suggest that the large observed differences in ΔEQ and D for the individual sites could be due to the influence of the strong diamagnetic anisotropy of the quadruply bonded MM unit.
Co-reporter:JohnF. Berry
Chemistry - A European Journal 2010 Volume 16( Issue 9) pp:2719-2724
Publication Date(Web):
DOI:10.1002/chem.200902324
Co-reporter:Paula M. B. Piccoli
Journal of Cluster Science 2010 Volume 21( Issue 3) pp:351-359
Publication Date(Web):2010 September
DOI:10.1007/s10876-010-0306-x
Reaction of Os2(OAc)4Cl2 with an excess of HDPhF (HDPhF = N,N′-diphenylformamidine) gives a high yield of Os2(DPhF)4Cl2 (1), which can be converted to its azido analog, Os2(DPhF)4(N3)2 (3), by treatment with NaN3. We report a major improvement on the preparation of Os2(chp)4Cl (2; Hchp = 2-chloro-6-hydroxypyridine) by synthesizing the compound in the reducing solvent ethanol. Reaction of 2 with NaN3 affords the azido complex Os2(chp)4N3 (4). Compound 3 has been examined by X-ray crystallography, and has an Os–Os bond distance of 2.45 Å, suggesting a (π*)2 ground state for the molecule.
Co-reporter:Michael Nippe, George H. Timmer and John F. Berry  
Chemical Communications 2009 (Issue 29) pp:4357-4359
Publication Date(Web):19 May 2009
DOI:10.1039/B907402C
Addition of CrCl2 to the dinuclear synthon MoW(dpa)4 yields a regioselectively formed heterotrimetallic MoW⋯Cr chain; computational studies suggest that the polarization of the MoW quadruple bond partially accounts for this unexpected selectivity.
Co-reporter:Michael Nippe, Eric Victor and John F. Berry
Inorganic Chemistry 2009 Volume 48(Issue 24) pp:11889-11895
Publication Date(Web):November 18, 2009
DOI:10.1021/ic901965b
Reported is a facile, high-yielding one-pot synthesis of the quadruply bonded ditungsten (II,II) compound W2(dpa)4 (1) (dpa = 2,2′-dipyridylamide), which was obtained from W(CO)6 at high temperature in naphthalene. A similar reaction in 1,2-dichlorobenzene furnished a ditungsten (III, III) species as the major product that was crystallized as [W2(dpa)3Cl2][BPh4] (3). The [W2(dpa)3Cl2]+ cation is better prepared by oxidation of 1 with SO2Cl2. Compound 1 was characterized by X-ray crystallography and cyclic-voltammetry, and is compared with its earlier reported molybdenum analogue, Mo2(dpa)4 (2). One-electron oxidation products of 1 and 2, [W2(dpa)4][BPh4] (1BPh4) and [Mo2(dpa)4][BPh4] (2BPh4), respectively, have also been synthesized. The crystallographically determined metal−metal distances of 2.23 Å and 2.14 Å in 1BPh4 and 2BPh4, respectively, are in agreement with metal−metal bond orders of 3.5. Unlike most previously reported Mo25+ and W25+ compounds, the primary coordination spheres around the M2-units in 1/1BPh4 and 2/2BPh4 remain unchanged upon one-electron oxidation, because the tridentate dpa ligand hinders axial coordination of exogenous ligands.
Co-reporter:József S. Pap, Jamie L. Snyder, Paula M. B. Piccoli and John F. Berry
Inorganic Chemistry 2009 Volume 48(Issue 20) pp:9846-9852
Publication Date(Web):September 25, 2009
DOI:10.1021/ic901419w
The reaction of Ru2(OAc)4Cl with N,N′,N′′-triphenylguanidine (HTPG) produces one of two different compounds depending on the reaction conditions. In acetone in the presence of triethyl amine, the reaction produces tri-substituted Ru2(TPG)3(OAc)Cl, and in refluxing xylene, the tetra-substituted Ru2(TPG)4Cl is produced. Both of these new complexes can be cleanly converted into their corresponding azido analogues by reaction with sodium azide in methanol. The X-ray crystal structures of Ru2(TPG)3(OAc)Cl, Ru2(TPG)3(OAc)N3, and Ru2(TPG)4Cl are presented, along with magnetic, electrochemical, and spectral measurements for each compound. Studies in solution show that, in contrast to Ru2(TPG)3(OAc)Cl, Ru2(TPG)4Cl is sterically hindered at the axial positions, and readily dissociates a chloride ion at high ionic strength. Equilibrium constants for chloride association and dissociation have been estimated. Mass spectrometric data suggest that the two azido complexes are precursors to new diruthenium nitrido species.
Co-reporter:Yu-Fei Song, John F. Berry, Thomas Weyhermüller and Eckhard Bill  
Dalton Transactions 2008 (Issue 14) pp:1864-1871
Publication Date(Web):16 Jan 2008
DOI:10.1039/B717618J
The coordination chemistry of CrCl3 with three pentadentate ligands having the [1,4,7]-triazacyclononane-1,4-diacetato motif have been investigated. The new resulting six-coordinate Cr(III)-chloro species react cleanly with sodium azide to form the corresponding azido species, which undergo photolysis under irradiation at 419 nm in acetonitrile–water solution to form Cr(V)-nitrido species that are partially hydrolyzed to their corresponding Cr(III)-hydroxo counterparts. Five of these Cr complexes have been characterized by X-ray crystallography. The hydroxo and nitrido species co-crystallize complicating crystal structural refinement. What at first appeared to be the longest CrN triple bond distance yet observed (1.66 Å) was found after re-refinement to be an artifact of positional disorder of the Cr atoms of 73% nitrido and 27% hydroxo species. The re-refined CrN bond distance is estimated as 1.58 Å, which agrees well with the observed CrN stretching frequency of 971 cm−1 found by IR spectroscopy. UV-vis data for the “nitrido” complexes suggest that they are all three partially hydrolyzed. These results emphasize the care that must be taken in the characterization of compounds that may co-crystallize with structurally similar analogs.
Co-reporter:John F. Berry, Serena DeBeer George and Frank Neese  
Physical Chemistry Chemical Physics 2008 vol. 10(Issue 30) pp:4361-4374
Publication Date(Web):02 Jun 2008
DOI:10.1039/B801803K
Recent advances in synthetic chemistry have led to the discovery of “superoxidized” iron centers with valencies Fe(V) and Fe(VI) [K. Meyer et al., J. Am. Chem. Soc., 1999, 121, 4859–4876; J. F. Berry et al., Science, 2006, 312, 1937–1941; F. T. de Oliveira et al., Science, 2007, 315, 835–838.]. Furthermore, in recent years a number of high-valent Fe(IV) species have been found as reaction intermediates in metalloenzymes and have also been characterized in model systems [C. Krebs et al., Acc. Chem. Res., 2007, 40, 484–492; L. Que, Jr, Acc. Chem. Res., 2007, 40, 493–500.]. These species are almost invariably stabilized by a highly basic ligand Xn− which is either O2− or N3−. The differences in structure and bonding between oxo- and nitrido species as a function of oxidation state and their consequences on the observable spectroscopic properties have never been carefully assessed. Hence, fundamental differences between high-valent iron complexes having either FeO or FeN multiple bonds have been probed computationally in this work in a series of hypothetical trans-[FeO(NH3)4OH]+/2+/3+ (1–3) and trans-[FeN(NH3)4OH]0/+/2+ (4–6) complexes. All computational properties are permeated by the intrinsically more covalent character of the FeN multiple bond as compared to the FeO bond. This difference is likely due to differences in Z* between N and O that allow for better orbital overlap to occur in the case of the FeN multiple bond. Spin-state energetics were addressed using elaborate multireference ab initio computations that show that all species 1–6 have an intrinsic preference for the low-spin state, except in the case of 1 in which S = 1 and S = 2 states are very close in energy. In addition to Mössbauer parameters, g-tensors, zero-field splitting and iron hyperfine couplings, X-ray absorption Fe K pre-edge spectra have been simulated using time-dependent DFT methods for the first time for a series of compounds spanning the high-valent states +4, +5, and +6 for iron. A remarkably good correlation of these simulated pre-edge features with experimental data on isolated high-valent intermediates has been found, allowing us to assign the main pre-edge features to excitations into the empty Fedz2 orbital, which is able to mix with Fe 4pz, allowing an efficient mechanism for the intensification of pre-edge features.
Co-reporter:Michael Nippe;Eric Victor
European Journal of Inorganic Chemistry 2008 Volume 2008( Issue 36) pp:5569-5572
Publication Date(Web):
DOI:10.1002/ejic.200801001

Abstract

Reported here are two new compounds containing either a CrCr···Co [1, CrCrCo(dpa)4Cl2, dpa = 2,2′-dipyridylamide] or a MoMo···Co [2, MoMoCo(dpa)4Cl2] framework both having a multiply-bonded unit (CrCr in 1, MoMo in 2) in close proximity to the Co2+ ion and trans to a Co–Cl bond. Variable temperature magnetic susceptibility measurements reveal 1 to have a temperature-dependent spin equilibrium between a low-spin (S = 1/2) and high-spin (S = 3/2) state, whereas the Co2+ ion in 2 exists solely in its high-spin state. The crystal structures of 1 and 2 were determined. Variable temperature crystallographic data of 1 at 100 K and at room temperature reveal that the spin-transition affects not only the Co–ligand bond lengths but also the terminal Cr–ligand bond lengths. Whereas the Cr···Co distance becomes shorter by 0.13 Å in the low-spin form, the Co–Cldistance becomes longer by 0.2 Å. These observations,along with the crystal structure of 2, suggest that the multiply-bonded MM group has a trans influence on the Co2+ ion.(© Wiley-VCH Verlag GmbH & Co. KGaA, 69451 Weinheim, Germany, 2008)

Co-reporter:JózsefS. Pap Dr.;Serena DeBeerGeorge Dr.;JohnF. Berry Dr.
Angewandte Chemie International Edition 2008 Volume 47( Issue 52) pp:10102-10105
Publication Date(Web):
DOI:10.1002/anie.200804397
Co-reporter:JózsefS. Pap Dr.;Serena DeBeerGeorge Dr.;JohnF. Berry Dr.
Angewandte Chemie 2008 Volume 120( Issue 52) pp:10256-10259
Publication Date(Web):
DOI:10.1002/ange.200804397
Co-reporter:David W. Brogden
Inorganic Chemistry () pp:
Publication Date(Web):July 22, 2015
DOI:10.1021/acs.inorgchem.5b01370
Reaction of Mo2(dpa)4 (dpa = 2,2′-dipyridylamido) with 1/2 equiv of [Ru(CO)3Cl2]2 in molten naphthalene at 250 °C provides facile access to the first all-second-row transition metal heterometallic chain compound, MoMoRu(dpa)4Cl2 (1). The one-electron oxidized compound [MoMoRu(dpa)4Cl2](OTf) (2) is synthesized by reaction of 1 with FeCp2(OTf). X-ray crystallography reveals a contraction of the Mo–Ru bond distance from 2.38 Å in 1 to 2.30 Å in 2, and an elongation of the Mo–Mo bond distance from 2.12 Å in 1 to 2.21 Å in 2. The short Mo–Ru bond distances indicate significant electron delocalization along the Mo–Mo–Ru chain, which is quantified by density functional theory (DFT) calculations. Molecular orbital analyses of both compounds based on DFT results reveal full delocalization of the orbitals of σ and π symmetry for both compounds. Additionally, δ orbital delocalization is observed in 2.
Co-reporter:Amanda R. Corcos and John F. Berry
Dalton Transactions 2017 - vol. 46(Issue 17) pp:NaN5539-5539
Publication Date(Web):2016/12/22
DOI:10.1039/C6DT04328C
Five new metal–metal bonded Ru2 compounds are presented and discussed: Ru2(ap)4ONO2 (2), [Ru2(ap)4NCMe][BF4] (3), Ru2(ap)4FBF3 (4), Ru2(ap)4OTf (5), and [Ru2(ap)4OTf][Ag(OTf)2] (6) (ap = 2-anilinopyridinate). All compounds have a (4,0) arrangement of the ap ligands about the Ru–Ru bond and contain one sterically blocked axial site and one site containing a labile ligand. These compounds display some of the shortest Ru–Ru distances known for this class of compounds. We demonstrate a reversible interconversion between compounds 3 and 4 as the MeCN and BF4− ligands are readily displaced. Despite the presence of labile axial ligands, compounds 2–5 remain high spin with an S = 3/2 ground state as determined by EPR spectroscopy.
Co-reporter:Sungho V. Park and John F. Berry
Dalton Transactions 2017 - vol. 46(Issue 28) pp:NaN9125-9125
Publication Date(Web):2017/06/26
DOI:10.1039/C7DT01847A
A series of RuII complexes stabilized with the pentapyridyl ligand Py5Me2 (Py5Me2 = 2,6-bis(1,1-bis(2-pyridyl)ethyl)pyridine) and with an axial X ligand (X = Cl−, H2O, N3−, MeCN) were prepared and characterized in the solid state and in non-aqueous solution. The cyclic voltammograms of these complexes in MeCN reflect a reversible substitution of the axial X ligand with MeCN. Irreversible ligand substitution of [(Py5Me2)RuN3]+ is also observed in propylene carbonate, but only at oxidizing potentials that decompose the azide ligand. The monometallic chloride and azide species are compared with analogous Ru2 metal–metal bonded complexes, which have been reported to undergo irreversible chloride dissociation upon reduction.
Co-reporter:Brian S. Dolinar, Stosh A. Kozimor and John F. Berry
Dalton Transactions 2016 - vol. 45(Issue 44) pp:NaN17605-17605
Publication Date(Web):2016/10/14
DOI:10.1039/C6DT03659G
We present the synthesis, structure, and electrochemistry of K3[Mo2(SNO5)4Cl]3[Mo2(SNO5)4] (1, HSNO5 = monothiosuccinimide), the first example of a heterometallic extended metal atom node (HEMAN). The HEMAN consists of two perpendicular, intersecting lines of metal atoms formed by three [Mo2(SNO5)4Cl]− units and one [Mo2(SNO5)4] unit tethered together by K+ ions.
Co-reporter:David W. Brogden and John F. Berry
Chemical Communications 2015 - vol. 51(Issue 44) pp:NaN9156-9156
Publication Date(Web):2015/04/30
DOI:10.1039/C5CC02917A
The location of the unpaired electron in the new mixed-valent (W2)IV,V trication [W2O(dpa)4]3+ presents a challenge for DFT methods. EPR spectroscopy confirms the unpaired electron to be in the W(V)–oxo unit, in agreement with the predictions of hybrid functionals B3LYP and TPSSh, but contrary to the predictions of non-hybrid functionals.
Co-reporter:Katherine P. Kornecki and John F. Berry
Chemical Communications 2012 - vol. 48(Issue 99) pp:NaN12099-12099
Publication Date(Web):2012/10/31
DOI:10.1039/C2CC36614B
A new mixed-valent Rh2II,III dimer, [Rh2(espn)2Cl] (espn2− = α,α,α′,α′-tetramethyl-1,3-benzenedipropanamidate), is reported. This compound readily dissociates Cl− at low concentrations in solution to form the active [Rh2(espn)2]+ catalyst, which performs intramolecular C–H amination with TONs > 1400. This work expands the scope of Rh2II,III dimers to nitrenoid chemistry.
Co-reporter:Michael Nippe, George H. Timmer and John F. Berry
Chemical Communications 2009(Issue 29) pp:NaN4359-4359
Publication Date(Web):2009/05/19
DOI:10.1039/B907402C
Addition of CrCl2 to the dinuclear synthon MoW(dpa)4 yields a regioselectively formed heterotrimetallic MoW⋯Cr chain; computational studies suggest that the polarization of the MoW quadruple bond partially accounts for this unexpected selectivity.
Co-reporter:George H. Timmer and John F. Berry
Chemical Science (2010-Present) 2012 - vol. 3(Issue 10) pp:NaN3052-3052
Publication Date(Web):2012/07/10
DOI:10.1039/C2SC20688A
Using density functional methods we have modelled the intramolecular electrophilic aryl C–H amination for 15 dimetal nitrides, both homo- and heteronuclear, along with 2 mononuclear nitrides, in the pursuit of understanding the reactivity of the dimetal nitrido Ru2(DPhF)4N (DPhF = N,N′-diphenyl formamidinate) molecule, for which this amination reaction was experimentally observed and characterized. It was found that the 3-center bonding manifold (MMN) that arises between the metal–metal bond and axial nitrido moiety has a dominant influence in the electronic structure and consequently the reactivity at each step in the reaction. It was found that transition state energetics correlate strongly with product stabilization and that these quantities depend on the number of electrons available to occupy the MMN manifold. As the reaction proceeds the number of orbitals in the manifold decreases by one and the point at which this happens determines which of two transition states is rate limiting. The dimetallic nitrides are shown to be inherently more reactive than the mononuclear complexes and so the MMN manifold that is only active in the dimetallic complexes comes through as an important factor in facilitating this amination reaction. Overall, a strong correlation between electronic structure and reactivity is established for C–H amination and new synthetic targets are proposed to develop new facets of this reactivity.
Co-reporter:Amanda R. Corcos and John F. Berry
Dalton Transactions 2016 - vol. 45(Issue 6) pp:NaN2389-2389
Publication Date(Web):2016/01/04
DOI:10.1039/C5DT04875C
The complex {[Ru2(ap)4]2[AgF2]}[BF4]3 ({2}[BF4]3, ap = 2-anilinopyridine), containing the [AgF2]− anion ligated to two [Ru2]6+ cores, is prepared, characterized, and compared to dimeric dumbbell-type structures, monomeric Ru2 structures, as well as the known set of dihalo coinage-metalate anions. X-ray crystallography indicates that the Ru–Ru and Ru–F distances are rather short, 2.2835(3) Å and 2.054(1) Å, respectively, while the Ag–F distance of 2.274(1) Å is longer than that calculated for the free/un-ligated anion. Cyclic voltammetry in dichloromethane indicates that, while some of {2}3+ breaks apart into an [Ru2(ap)4F]+ ([3]+) monomer in solution, the remaining dimer has a single reversible two-electron redox feature for the Ru26/5+ couple that is at a lower potential than that of [3]+. This is one of the few examples of a ligated dihalo coinage-metalate, and it is the first example of a coinage metal difluoride anion, either free or ligated.
Co-reporter:Travis L. Sunderland and John F. Berry
Dalton Transactions 2016 - vol. 45(Issue 1) pp:NaN55-55
Publication Date(Web):2015/11/19
DOI:10.1039/C5DT03740A
Five novel homoleptic heterobimetallic bismuth(II)–rhodium(II) carboxylate complexes—BiRh(TPA)4 (1), BiRh(but)4 (2), BiRh(piv)4 (3), BiRh(esp)2 (4), and BiRh(OAc)4 (5)—were synthesized in good yields by equatorial ligand substitution starting from BiRh(TFA)4 (TPA = triphenylacetate, but = butyrate, piv = pivalate, esp = α,α,α′,α′-tetramethyl-1,3-benzenedipropionate, OAc = acetate, and TFA = trifluoroacetate). We report here 1H and 13C{1H} NMR spectra and cyclic voltammograms for complexes 1–4, and IR spectra for all complexes. Irreversible redox waves appear between −1.4 to −1.5 V for [BiRh]3+/4+ couples and 1.3 to 1.5 V vs. Fc/Fc+ for [BiRh]4+/5+ couples for complexes 1–4 indicating a wide range of stability for the compounds. The X-ray crystal structure of 1 reveals a Bi–Rh distance of 2.53 Å.
Co-reporter:Brian S. Dolinar and John F. Berry
Dalton Transactions 2014 - vol. 43(Issue 16) pp:NaN6176-6176
Publication Date(Web):2014/02/19
DOI:10.1039/C4DT00297K
We report an exploration of the coordination chemistry of a systematic series of cyclic thioamidate ligands with the quadruply-bonded Mo24+ core. In addition to the S and N donor atoms that bind to Mo, the ligands utilized in this study have an additional O or S atom in conjugation with the thioamidate π system. The preparation of four new Mo2 complexes is described, and these compounds are characterized by X-ray crystallography, NMR and UV-vis spectroscopy, electrochemistry, and DFT calculations. These complexes provide a means to interrogate the electronics of Mo2(thioamidate)4 systems. Notably, we describe the first two examples of Mo2(thioamidate)4 complexes in their cis-2,2-regioisomer. By varying the π-system substituent and regioisomerism of these compounds, the electronics of the dimolybdenum core is shown to be altered with varying degrees of effect. Cyclic voltammetry results show that changing the π-system substituent from O to S results in an increase in the Mo24+/5+ oxidation potential by 170 mV. Changing the arrangement of ligands around the dimolybdenum core from trans-2,2 to cis-2,2 slightly weakens the metal–ligand bonds, raising the oxidation potential by a more modest 30–100 mV. MO diagrams of each compound derived from DFT calculations support these conclusions as well; the identity of the π-system substituent alters the δ–δ* (HOMO–LUMO) gap by up to 0.4 eV, whereas regioisomerism yields smaller changes in the electronic structure.
Co-reporter:Yevgeniya Turov and John F. Berry
Dalton Transactions 2012 - vol. 41(Issue 26) pp:NaN8161-8161
Publication Date(Web):2012/03/19
DOI:10.1039/C2DT30150D
We report here two novel synthetic pathways toward the preparation of a family of trimetallic diazide compounds of the type Cr2M(dpa)4(N3)2, with M = Cr (10), Mn (4), Fe (5), and Co (11). Reaction of either Cr2M(dpa)4(OTf)2 (for M = Mn and Fe) or [Cr2M(dpa)4(MeCN)2](PF6)2 (for M = Cr and Co) with sodium azide in methanol leads to the formation of the corresponding diazide compounds, and single crystal X-ray diffraction measurements confirm the predicted structures. Compounds 4, 5, and 10 are all high-spin compounds, but 11 is a spin-crossover compound exhibiting low-spin behavior at low temperatures (∼100 K). Thermolytic characterization by DSC and TGA reveals an exothermic reaction corresponding to the loss of two dinitrogen molecules from compounds 5, 10, and 11. Further characterization by solution NMR measurements and cyclic voltammetry are also presented.
Co-reporter:Yu-Fei Song, John F. Berry, Thomas Weyhermüller and Eckhard Bill
Dalton Transactions 2008(Issue 14) pp:NaN1871-1871
Publication Date(Web):2008/01/16
DOI:10.1039/B717618J
The coordination chemistry of CrCl3 with three pentadentate ligands having the [1,4,7]-triazacyclononane-1,4-diacetato motif have been investigated. The new resulting six-coordinate Cr(III)-chloro species react cleanly with sodium azide to form the corresponding azido species, which undergo photolysis under irradiation at 419 nm in acetonitrile–water solution to form Cr(V)-nitrido species that are partially hydrolyzed to their corresponding Cr(III)-hydroxo counterparts. Five of these Cr complexes have been characterized by X-ray crystallography. The hydroxo and nitrido species co-crystallize complicating crystal structural refinement. What at first appeared to be the longest CrN triple bond distance yet observed (1.66 Å) was found after re-refinement to be an artifact of positional disorder of the Cr atoms of 73% nitrido and 27% hydroxo species. The re-refined CrN bond distance is estimated as 1.58 Å, which agrees well with the observed CrN stretching frequency of 971 cm−1 found by IR spectroscopy. UV-vis data for the “nitrido” complexes suggest that they are all three partially hydrolyzed. These results emphasize the care that must be taken in the characterization of compounds that may co-crystallize with structurally similar analogs.
Co-reporter:John F. Berry, Serena DeBeer George and Frank Neese
Physical Chemistry Chemical Physics 2008 - vol. 10(Issue 30) pp:NaN4374-4374
Publication Date(Web):2008/06/02
DOI:10.1039/B801803K
Recent advances in synthetic chemistry have led to the discovery of “superoxidized” iron centers with valencies Fe(V) and Fe(VI) [K. Meyer et al., J. Am. Chem. Soc., 1999, 121, 4859–4876; J. F. Berry et al., Science, 2006, 312, 1937–1941; F. T. de Oliveira et al., Science, 2007, 315, 835–838.]. Furthermore, in recent years a number of high-valent Fe(IV) species have been found as reaction intermediates in metalloenzymes and have also been characterized in model systems [C. Krebs et al., Acc. Chem. Res., 2007, 40, 484–492; L. Que, Jr, Acc. Chem. Res., 2007, 40, 493–500.]. These species are almost invariably stabilized by a highly basic ligand Xn− which is either O2− or N3−. The differences in structure and bonding between oxo- and nitrido species as a function of oxidation state and their consequences on the observable spectroscopic properties have never been carefully assessed. Hence, fundamental differences between high-valent iron complexes having either FeO or FeN multiple bonds have been probed computationally in this work in a series of hypothetical trans-[FeO(NH3)4OH]+/2+/3+ (1–3) and trans-[FeN(NH3)4OH]0/+/2+ (4–6) complexes. All computational properties are permeated by the intrinsically more covalent character of the FeN multiple bond as compared to the FeO bond. This difference is likely due to differences in Z* between N and O that allow for better orbital overlap to occur in the case of the FeN multiple bond. Spin-state energetics were addressed using elaborate multireference ab initio computations that show that all species 1–6 have an intrinsic preference for the low-spin state, except in the case of 1 in which S = 1 and S = 2 states are very close in energy. In addition to Mössbauer parameters, g-tensors, zero-field splitting and iron hyperfine couplings, X-ray absorption Fe K pre-edge spectra have been simulated using time-dependent DFT methods for the first time for a series of compounds spanning the high-valent states +4, +5, and +6 for iron. A remarkably good correlation of these simulated pre-edge features with experimental data on isolated high-valent intermediates has been found, allowing us to assign the main pre-edge features to excitations into the empty Fedz2 orbital, which is able to mix with Fe 4pz, allowing an efficient mechanism for the intensification of pre-edge features.
Sulfamic acid, N-2-cyclohexen-1-yl-, 2,2,2-trichloroethyl ester
7-Azabicyclo[4.1.0]heptane-7-sulfonic acid, 2,2,2-trichloroethyl ester
7-Azabicyclo[4.1.0]heptane,7-[S-(4-methylphenyl)-N-[(4-methylphenyl)sulfonyl]sulfonimidoyl]-
ACETIC ACID;4-(2,6,6-TRIMETHYLCYCLOHEX-2-EN-1-YL)BUT-3-EN-2-OL
Sulfamic acid, 2,2,2-trifluoro-1-(trifluoromethyl)ethyl ester
Sulfamic acid, 3-nitrophenyl ester
2,2':6',2''-Terpyridine, 4,4',4''-tris(1,1-dimethylethyl)-
Benzenesulfonamide, 4-methyl-N-[S-(4-methylphenyl)sulfonimidoyl]-
Sulfamic acid, 2,2,2-trifluoroethyl ester
CYCLOHEXENE, 1-(1-HEPTYNYL)-