@article{daubert_afroz_borodin_seo_boyle_henderson_2022, title={Solvate Structures and Computational/Spectroscopic Characterization of LiClO4 Electrolytes}, ISSN={["1932-7455"]}, DOI={10.1021/acs.jpcc.2c03805}, abstractNote={A Raman spectral evaluation of numerous crystalline solvates with lithium perchlorate (LiClO4) has been conducted over a wide temperature range. Two new solvate crystal structures─(PMDETA)1:LiClO4 and (THF)1:LiClO4 with N,N,N′,N″,N″-pentamethyldiethylenetriamine and tetrahydrofuran─have been determined to aid in this study. With a help of density functional theory (DFT) and molecular dynamics (MD) simulations, the spectroscopic data have been correlated with varying modes of ClO4–···Li+ cation coordination within the solvate structures to create a characterization tool to facilitate the Raman band assignments for the determination of ionic association interactions within solid and liquid electrolytes containing LiClO4. This study demonstrates that many of the spectroscopic evaluation conclusions reported in the scientific literature for LiClO4-based electrolytes are inaccurate.}, journal={JOURNAL OF PHYSICAL CHEMISTRY C}, author={Daubert, James S. and Afroz, Taliman and Borodin, Oleg and Seo, Daniel M. and Boyle, Paul D. and Henderson, Wesley A.}, year={2022}, month={Aug} } @article{feese_gracz_boyle_ghiladi_2019, title={Towards microbe-targeted photosensitizers: Synthesis, characterization and in vitro photodynamic inactivation of the tuberculosis model pathogen M. smegmatis by porphyrin-peptide conjugates}, volume={23}, ISSN={1088-4246 1099-1409}, url={http://dx.doi.org/10.1142/S1088424619501505}, DOI={10.1142/S1088424619501505}, abstractNote={ Porphyrin-peptide conjugates have a breadth of potential applications, including use in photodynamic therapy, boron neutron capture therapy, as fluorescence imaging tags for tracking subcellular localization, as magnetic resonance imaging (MRI) positive-contrast reagents and as biomimetic catalysts. Here, we have explored three general routes to porphyrin-peptide conjugates using the Cu(I)-catalyzed Huisgen-Medal-Sharpless 1,3-dipolar cycloaddition of peptide-containing azides with a terminal alkyne-containing porphyrin, thereby generating porphyrin-peptide conjugates (PPCs) comprised of a cationic porphyrin coupled to short antimicrobial peptides. In addition to characterizing the PPCs using a variety of spectroscopic (UV-vis, [Formula: see text]H- and [Formula: see text]C-NMR) and mass spectrometric methods, we evaluated their efficacy as photosensitizers for the in vitro photodynamic inactivation of Mycobacterium smegmatis as a model for the pathogen Mycobacterium tuberculosis. Difficulties that needed to be overcome for the efficient synthesis of PPCs were the limited solubility of the quaternized pyridyl porphyrin in common solvents, undesired (de)metallation and transmetallation, and chromatographic purification. Photodynamic inactivation studies of a small library of PPCs against Mycobacterium smegmatis confirmed our hypothesis that the porphyrin-based photosensitizer maintains its ability to efficiently inactivate bacteria when conjugated to a small peptide by upwards of 5–6 log units (99.999[Formula: see text]%) using white light illumination (400–700 nm, 60 mW/cm[Formula: see text], 30 min). Further, hemolysis assays revealed the lack of toxicity of the PPCs against sheep blood at concentrations employed for in vitro photodynamic inactivation. Taken together, the results demonstrated the ability of PPCs to maintain their antimicrobial photodynamic inactivation efficacy when possessing a short cationic peptides for enabling the potential targeting of pathogens in vivo. }, number={11n12}, journal={Journal of Porphyrins and Phthalocyanines}, publisher={World Scientific Pub Co Pte Lt}, author={Feese, Elke and Gracz, Hanna S. and Boyle, Paul D. and Ghiladi, Reza A.}, year={2019}, month={Dec}, pages={1414–1439} } @article{perez_smeltz_sommer_boyle_ison_2017, title={Cationic rhenium ((III)) complexes: synthesis, characterization, and reactivity for hydrosilylation of aldehydes}, volume={46}, DOI={10.1039/c7dt00271h}, abstractNote={Cationic Re(iii) complexes are shown to be more active for the catalytic hydrosilylation of benzaldehydes than their neutral acetate precursors.}, number={14}, journal={Dalton Transactions (Cambridge, England : 2003)}, author={Perez, D. E. and Smeltz, J. L. and Sommer, Roger D. and Boyle, P. D. and Ison, E. A.}, year={2017}, pages={4609–4616} } @article{afroz_seo_han_boyle_henderson_2015, title={Structural Interactions within Lithium Salt Solvates: Acyclic Carbonates and Esters}, volume={119}, ISSN={["1932-7447"]}, DOI={10.1021/acs.jpcc.5b00309}, abstractNote={Solvate crystal structures serve as useful models for the molecular-level interactions within the diverse solvates present in liquid electrolytes. Although acyclic carbonate solvents are widely used for Li-ion battery electrolytes, only three solvate crystal structures with lithium salts are known for these and related solvents. The present work, therefore, reports six lithium salt solvate structures with dimethyl and diethyl carbonate, (DMC)2:LiPF6, (DMC)1:LiCF3SO3, (DMC)1/4:LiBF4, (DEC)2:LiClO4, (DEC)1:LiClO4, and (DEC)1:LiCF3SO3 and four with the structurally related methyl and ethyl acetate, (MA)2:LiClO4, (MA)1:LiBF4, (EA)1:LiClO4, and (EA)1:LiBF4.}, number={13}, journal={JOURNAL OF PHYSICAL CHEMISTRY C}, author={Afroz, Taliman and Seo, Daniel M. and Han, Sang-Don and Boyle, Paul D. and Henderson, Wesley A.}, year={2015}, month={Apr}, pages={7022–7027} } @article{robbins_lilly_smeltz_boyle_ison_2015, title={Synthesis and Reactivity of Oxorhenium(V) Methyl, Benzyl, and Phenyl Complexes with CO: Implications for a Unique Mechanism for Migratory Insertion}, volume={34}, ISSN={["1520-6041"]}, DOI={10.1021/acs.organomet.5b00177}, abstractNote={The complexes [(DAAm)Re(O)(R)] [DAAm = N,N-bis(2-arylaminoethyl)methylamine; aryl = C6F5], 1, R = Me; 3a–d (R = benzyl, a; 4-methylbenzyl, b; 4-fluorobenzyl, c; 4-methoxybenzyl, d); and 4, R = Ph, were synthesized. CO insertion into the Re–R bond in 1 and 3a–d resulted in the formation of the acetyl complex, 2, and the (aryl)acetyl complexes, 5a–d respectively. The formation of 5a–d proceeded at a faster rate (7 h) than the formation of 2 (72 h) under the same conditions. No reaction was observed however for the phenyl complex 4 with CO. Kinetics for CO insertion into the various Re–R bonds were examined, and the experimental rate law was determined to be Rate = kobs[Re][CO]. The activation parameters for CO insertion into 1 and 3a were determined to be ΔG⧧(298 K) = 24(1). The enthalpy of activation ΔH⧧ was determined to be 9(1) and 10(3) kcal/mol for 1 and 3a, respectively, and the entropy of activation, ΔS⧧, was −49(2) and −36(4) cal/mol·K. Computational studies (M06) are consistent with the hypothesis ...}, number={13}, journal={ORGANOMETALLICS}, author={Robbins, Leanna K. and Lilly, Cassandra P. and Smeltz, Jessica L. and Boyle, Paul D. and Ison, Elon A.}, year={2015}, month={Jul}, pages={3152–3158} } @article{he_freeman_nakpathom_boyle_2015, title={Synthesis and X-ray analysis of a perfluoroalkyl-substituted azobenzene dye}, volume={120}, ISSN={["1873-3743"]}, DOI={10.1016/j.dyepig.2015.04.014}, abstractNote={A search for colorants capable of reducing the surface energy of synthetic fibers as well as adding color led to the synthesis and characterization of the repellency behavior of dyes such as 4-N,N-diethylamino-4′-tridecafluorohexylazobenzene. Unexpectedly high surface energy values and a low fluorine content on fiber surfaces then led to an interest in determining the crystal structure of this new dye. Accordingly, a single crystal was grown from acetone solution, and its structure was established using X-ray diffraction analysis. Interestingly, it was found that the azobenzene skeleton is appreciably nonplanar, having an N2–N1–C1–C2 torsion angle of 30.6°, despite the absence of substituents ortho to the azo bond. Further, the structure is characterized by head-to-tail molecular stacking and the N-ethyl groups in the molecule are positioned above the aminobenzene plane on the same side. It is likely that this combination of factors contribute to the observed surface properties of the target dye.}, journal={DYES AND PIGMENTS}, author={He, Liang and Freeman, Harold S. and Nakpathom, Monthon and Boyle, Paul D.}, year={2015}, month={Sep}, pages={245–250} } @article{tichnell_shultz_popescu_sokirniy_boyle_2015, title={Synthesis, Characterization, and Photophysical Studies of an Iron(III) Catecholate-Nitronylnitroxide Spin-Crossover Complex}, volume={54}, ISSN={["1520-510X"]}, DOI={10.1021/acs.inorgchem.5b00298}, abstractNote={The synthesis and characterization of an Fe(III) catecholate-nitronylnitroxide (CAT-NN) complex (1-NN) that undergoes Fe(III) spin-crossover is described. Our aim is to determine whether the intraligand exchange coupling of the semiquinone-nitronylnitroxide Fe(II)(SQ-NN) excited state resulting from irradiation of the CAT → Fe(III) LMCT band would affect either the intrinsic photophysics or the iron spin-crossover event when compared to the complex lacking the nitronylnitroxide radical (1). X-ray crystallographic analysis provides bond lengths consistent with a ferric catecholate charge distribution. Mössbauer spectroscopy clearly demonstrates Fe(III) spin-crossover, hyperfine couplings, and a weak ferromagnetic Fe(III)-CAT-NN exchange, and spin-crossover is corroborated by variable-temperature magnetic susceptibility and electronic absorption studies. To explore the effect of the NN radical on photophysical processes, we conducted room-temperature transient absorption experiments. Upon excitation of the ligand-to-metal charge transfer band, an Fe(II)SQ state is populated and most likely undergoes fast intersystem crossing to the ligand field manifold, where it rapidly decays into a metastable low-spin Fe(III)CAT state, followed by repopulation of the high-spin Fe(III)CAT ground state. The decay components of 1-NN are slightly faster than those obtained for 1, perhaps due to the higher number of microstates present within the LMCT and LF manifolds for 1-NN. Although the effects of the NN radical are manifest in neither the spin-crossover nor the photophysics, our results lay the groundwork for future studies.}, number={9}, journal={INORGANIC CHEMISTRY}, author={Tichnell, Christopher R. and Shultz, David A. and Popescu, Codrina V. and Sokirniy, Ivan and Boyle, Paul D.}, year={2015}, month={May}, pages={4466–4474} } @article{mcowen_delp_paillard_herriot_han_boyle_sommer_henderson_2014, title={Anion Coordination Interactions in Solvates with the Lithium Salts LiDCTA and LiTDI}, volume={118}, ISSN={["1932-7447"]}, DOI={10.1021/jp412601x}, abstractNote={Lithium 4,5-dicyano-1,2,3-triazolate (LiDCTA) and lithium 2-trifluoromethyl-4,5-dicyanoimidazole (LiTDI) are two salts proposed for lithium battery electrolyte applications, but little is known about the manner in which the DCTA– and TDI– anions coordinate Li+ cations. To explore this in depth, crystal structures are reported here for two solvates with LiDCTA—(G2)1:LiDCTA and (G1)1:LiDCTA—with diglyme and monoglyme, respectively; and seven solvates with LiTDI—(G1)2:LiTDI, (G2)2:LiTDI, (G3)1:LiTDI, (THF)1:LiTDI, (EC)1:LiTDI, (PC)1:LiTDI, and (DMC)1/2:LiTDI—with monoglyme, diglyme, triglyme, tetrahydrofuran, ethylene carbonate, propylene carbonate, and dimethyl carbonate, respectively. These latter solvate structures are compared with the previously reported acetonitrile (AN)2:LiTDI structure. The solvates indicate that the LiTDI salt is much less associated than the LiDCTA salt and that the ions in LiTDI, when aggregated in solvates, have a very similar TDI–···Li+ cation mode of coordination through both t...}, number={15}, journal={JOURNAL OF PHYSICAL CHEMISTRY C}, publisher={American Chemical Society (ACS)}, author={McOwen, Dennis W. and Delp, Samuel A. and Paillard, Elie and Herriot, Cristelle and Han, Sang-Don and Boyle, Paul D. and Sommer, Roger D. and Henderson, Wesley A.}, year={2014}, month={Apr}, pages={7781–7787} } @article{mcowen_seo_borodin_vatamanu_boyle_henderson_2014, title={Concentrated electrolytes: decrypting electrolyte properties and reassessing Al corrosion mechanisms}, volume={7}, ISSN={["1754-5706"]}, DOI={10.1039/c3ee42351d}, abstractNote={Highly concentrated electrolytes containing carbonate solvents with lithium bis(trifluoromethanesulfonyl)imide (LiTFSI) have been investigated to determine the influence of eliminating bulk solvent (i.e., uncoordinated to a Li+ cation) on electrolyte properties. The phase behavior of ethylene carbonate (EC)–LiTFSI mixtures indicates that two crystalline solvates form—(EC)3:LiTFSI and (EC)1:LiTFSI. Crystal structures for these were determined to obtain insight into the ion and solvent coordination. Between these compositions, however, a crystallinity gap exists. A Raman spectroscopic analysis of the EC solvent bands for the 3–1 and 2–1 EC–LiTFSI liquid electrolytes indicates that ∼86 and 95%, respectively, of the solvent is coordinated to the Li+ cations. This extensive coordination results in significantly improved anodic oxidation and thermal stabilities as compared with more dilute (i.e., 1 M) electrolytes. Further, while dilute EC–LiTFSI electrolytes extensively corrode the Al current collector at high potential, the concentrated electrolytes do not. A new mechanism for electrolyte corrosion of Al in Li-ion batteries is proposed to explain this. Although the ionic conductivity of concentrated EC–LiTFSI electrolytes is somewhat low relative to the current state-of-the-art electrolyte formulations used in commercial Li-ion batteries, using an EC–diethyl carbonate (DEC) mixed solvent instead of pure EC markedly improves the conductivity.}, number={1}, journal={ENERGY & ENVIRONMENTAL SCIENCE}, author={McOwen, Dennis W. and Seo, Daniel M. and Borodin, Oleg and Vatamanu, Jenet and Boyle, Paul D. and Henderson, Wesley A.}, year={2014}, month={Jan}, pages={416–426} } @article{bis_boyle_carino_igo_katrincic_2014, title={Crystallization and Solid-State Characterization of the Hemihydrate of Albuterol Hemisulfate}, volume={14}, ISSN={["1528-7505"]}, DOI={10.1021/cg4016405}, abstractNote={Comprehensive solid-state characterization and development of a controlled crystallization process of the recently discovered hemihydrate of albuterol hemisulfate (ASH) are presented. The ASH is physically stable when stored at ambient conditions for at least 3 months and at 0% relative humidity for at least 2 months. The thermal dehydration of the hydrate occurs at ∼100 °C and proceeds without a loss of crystallinity. The dehydrated crystal structure persists up to decomposition temperature (∼160–185 °C). An inspection of the single-crystal X-ray structure revealed an isolated-site (class A) hydrate with a complex three-dimensional supramolecular network sustained via multiple strong hydrogen bonds. Thermodynamic stability studies established that the hydrate is more-stable relative to the nonsolvated form at aw > 0.63 at 5 °C, aw > 0.65 at 20 °C, and aw > 0.80 at 40 °C. The constructed phase diagram for the anhydrate–hydrate system guided the development of a controlled crystallization process at a 0.5 ...}, number={2}, journal={CRYSTAL GROWTH & DESIGN}, author={Bis, Joanna A. and Boyle, Paul D. and Carino, Stephen A. R. and Igo, David H. and Katrincic, Lee M.}, year={2014}, month={Feb}, pages={775–782} } @article{lehman_boyle_sommer_ison_2014, title={Oxyfunctionalization with Cp*Ir-III(NHC)(Me)L Complexes}, volume={33}, ISSN={["1520-6041"]}, DOI={10.1021/om5007352}, abstractNote={A series of monomethyl Cp*IrIII complexes were synthesized and studied for the formation of methanol in water. Methanol yields of 75(4)% in the presence of O2 were obtained. From isotope labeling studies, it was determined that O2 is the source of the oxygen atom in the product. From kinetic studies, oxyfunctionalization appears to proceed by dissociation of an L-type ligand followed by O2 binding and insertion.}, number={19}, journal={ORGANOMETALLICS}, publisher={American Chemical Society (ACS)}, author={Lehman, Matthew C. and Boyle, Paul D. and Sommer, Roger D. and Ison, Elon A.}, year={2014}, month={Oct}, pages={5081–5084} } @article{seo_boyle_allen_han_jonsson_johansson_henderson_2014, title={Solvate Structures and Computational/Spectroscopic Characterization of LiBF4 Electrolytes}, volume={118}, ISSN={["1932-7455"]}, DOI={10.1021/jp5046782}, abstractNote={Crystal structures have been determined for both LiBF4 and HBF4 solvates: (acetonitrile)2:LiBF4, (ethylene glycol diethyl ether)1:LiBF4, (diethylene glycol diethyl ether)1:LiBF4, (tetrahydrofuran)1:LiBF4, (methyl methoxyacetate)1:LiBF4, (succinonitrile)1:LiBF4, (N,N,N′,N″,N″-pentamethyldiethylenetriamine)1:HBF4, (N,N,N′,N′-tetramethylethylenediamine)3/2:HBF4, and (phenanthroline)2:HBF4. These, as well as other known LiBF4 solvate structures, have been characterized by Raman vibrational spectroscopy to unambiguously assign the anion Raman band positions to specific forms of BF4–···Li+ cation coordination. In addition, complementary DFT calculations of BF4–···Li+ cation complexes have provided additional insight into the challenges associated with accurately interpreting the anion interactions from experimental Raman spectra. This information provides a crucial tool for the characterization of the ionic association interactions within electrolytes.}, number={32}, journal={JOURNAL OF PHYSICAL CHEMISTRY C}, author={Seo, Daniel M. and Boyle, Paul D. and Allen, Joshua L. and Han, Sang-Don and Jonsson, Erlendur and Johansson, Patrik and Henderson, Wesley A.}, year={2014}, month={Aug}, pages={18377–18386} } @article{seo_boyle_sommer_daubert_borodin_henderson_2014, title={Solvate Structures and Spectroscopic Characterization of LiTFSI Electrolytes}, volume={118}, ISSN={["1520-6106"]}, DOI={10.1021/jp505006x}, abstractNote={A Raman spectroscopic evaluation of numerous crystalline solvates with lithium bis(trifluoromethanesulfonyl)imide (LiTFSI or LiN(SO2CF3)2) has been conducted over a wide temperature range. Four new crystalline solvate structures-(PHEN)3:LiTFSI, (2,9-DMPHEN)2:LiTFSI, (G3)1:LiTFSI and (2,6-DMPy)1/2:LiTFSI with phenanthroline, 2,9-dimethyl[1,10]phenanthroline, triglyme, and 2,6-dimethylpyridine, respectively-have been determined to aid in this study. The spectroscopic data have been correlated with varying modes of TFSI(-)···Li(+) cation coordination within the solvate structures to create an electrolyte characterization tool to facilitate the Raman band deconvolution assignments for the determination of ionic association interactions within electrolytes containing LiTFSI. It is found, however, that significant difficulties may be encountered when identifying the distributions of specific forms of TFSI(-) anion coordination present in liquid electrolyte mixtures due to the wide range of TFSI(-)···Li(+) cation interactions possible and the overlap of the corresponding spectroscopic data signatures.}, number={47}, journal={JOURNAL OF PHYSICAL CHEMISTRY B}, publisher={American Chemical Society (ACS)}, author={Seo, Daniel M. and Boyle, Paul D. and Sommer, Roger D. and Daubert, James S. and Borodin, Oleg and Henderson, Wesley A.}, year={2014}, month={Nov}, pages={13601–13608} } @article{seo_afroz_allen_boyle_trulove_de long_henderson_2014, title={Structural Interactions within Lithium Salt Solvates: Cyclic Carbonates and Esters}, volume={118}, ISSN={["1932-7447"]}, DOI={10.1021/jp5079168}, abstractNote={Only limited information is available regarding the manner in which cyclic carbonate and ester solvents coordinate Li+ cations in electrolyte solutions for lithium batteries. One approach to gleaning significant insight into these interactions is to examine crystalline solvate structures. To this end, eight new solvate structures are reported with ethylene carbonate, γ-butyrolactone, and γ-valerolactone: (EC)3:LiClO4, (EC)2:LiClO4, (EC)2:LiBF4, (GBL)4:LiPF6, (GBL)1:LiClO4, (GVL)1:LiClO4, (GBL)1:LiBF4, and (GBL)1:LiCF3SO3. The crystal structure of (EC)1:LiCF3SO3 is also re-reported for comparison. These structures enable the factors that govern the manner in which the ions are coordinated and the ion/solvent packing—in the solid-state—to be scrutinized in detail.}, number={45}, journal={JOURNAL OF PHYSICAL CHEMISTRY C}, author={Seo, Daniel M. and Afroz, Taliman and Allen, Joshua L. and Boyle, Paul D. and Trulove, Paul C. and De Long, Hugh C. and Henderson, Wesley A.}, year={2014}, month={Nov}, pages={25884–25889} } @article{frasco_lilly_boyle_ison_2013, title={Cp*Ir-III-Catalyzed Oxidative Coupling of Benzoic Acids with Alkynes}, volume={3}, ISSN={["2155-5435"]}, DOI={10.1021/cs400656q}, abstractNote={Cp*Ir(III) complexes have been shown to catalyze the oxidative coupling of benzoic acids with alkynes in methanol at 60 °C to form a variety of isocoumarins. The use of AgOAc as an oxidant was required to facilitate significant product formation. Alkyl alkynes were shown to be more reactive substrates than aryl alkynes, and a number of functional groups were tolerated on benzoic acid. Combined mechanistic and computational studies (BP86) revealed that (1) C–H activation occurs via an acetate-assisted mechanism; (2) C–H activation is not turnover limiting; and (3) the oxidant oxidizes the reduced form of the catalyst via an Ir(I)–Ir(II)–Ir(III) sequence.}, number={10}, journal={ACS CATALYSIS}, author={Frasco, Daniel A. and Lilly, Cassandra P. and Boyle, Paul D. and Ison, Elon A.}, year={2013}, month={Oct}, pages={2421–2429} } @article{lehman_gary_boyle_sanford_ison_2013, title={Effect of Solvent and Ancillary Ligands on the Catalytic H/D Exchange Reactivity of Cp*Ir-III(L) Complexes}, volume={3}, ISSN={["2155-5435"]}, DOI={10.1021/cs400420n}, abstractNote={The reactivity of a series of Cp*IrIII(L) complexes that contain a diverse set of ancillary ligands, L, (L = PMe3, N-heterocyclic carbene, NHC = 1,3-dimethylimidazol-2-ylidene, aqua, 4-t-butylpyridine, and 4-(2,4,6-tris-(4-t-butylphenyl)pyridinium)pyridine tetrafluoroborate) has been examined in catalytic H/D exchange reactions between C6H6 and a series of deuterated solvents (methanol-d4, acetic acid-d4, and trifluoroacetic acid-d1). These studies demonstrate that (1) the mechanism of catalytic H/D exchange is significantly influenced by the nature of the solvent; (2) electron-donating ligands (PMe3, NHC) promote the formation of Ir hydrides in methanol-d4, and these are critical intermediates in catalytic H/D exchange processes; and (3) weak/poorly donating ligands (4-t-butylpyridine, 4-(2,4,6-tris-(4-t-butylphenyl)pyridinium)pyridine tetrafluoroborate and aqua) can support efficient H/D exchange catalysis in acetic acid-d4.}, number={10}, journal={ACS CATALYSIS}, author={Lehman, Matthew C. and Gary, J. Brannon and Boyle, Paul D. and Sanford, Melanie S. and Ison, Elon A.}, year={2013}, month={Oct}, pages={2304–2310} } @article{kirk_shultz_stasiw_habel-rodriguez_stein_boyle_2013, title={Electronic and Exchange Coupling in a Cross-Conjugated D–B–A Biradical: Mechanistic Implications for Quantum Interference Effects}, volume={135}, ISSN={0002-7863 1520-5126}, url={http://dx.doi.org/10.1021/JA405354X}, DOI={10.1021/ja405354x}, abstractNote={A combination of variable-temperature EPR spectroscopy, electronic absorption spectroscopy, and magnetic susceptibility measurements have been performed on Tp(Cum,Me)Zn(SQ-m-Ph-NN) (1-meta) a donor-bridge-acceptor (D-B-A) biradical that possesses a cross-conjugated meta-phenylene (m-Ph) bridge and a spin singlet ground state. The experimental results have been interpreted in the context of detailed bonding and excited-state computations in order to understand the excited-state electronic structure of 1-meta. The results reveal important excited-state contributions to the ground-state singlet-triplet splitting in this cross-conjugated D-B-A biradical that contribute to our understanding of electronic coupling in cross-conjugated molecules and specifically to quantum interference effects. In contrast to the conjugated isomer, which is a D-B-A biradical possessing a para-phenylene bridge, admixture of a single low-lying singly excited D → A type configuration into the cross-conjugated D-B-A biradical ground state makes a negligible contribution to the ground-state magnetic exchange interaction. Instead, an excited state formed by a Ph-NN (HOMO) → Ph-NN (LUMO) one-electron promotion configurationally mixes into the ground state of the m-Ph bridged D-A biradical. This results in a double (dynamic) spin polarization mechanism as the dominant contributor to ground-state antiferromagnetic exchange coupling between the SQ and NN spins. Thus, the dominant exchange mechanism is one that activates the bridge moiety via the spin polarization of a doubly occupied orbital with phenylene bridge character. This mechanism is important, as it enhances the electronic and magnetic communication in cross-conjugated D-B-A molecules where, in the case of 1-meta, the magnetic exchange in the active electron approximation is expected to be J ~ 0 cm(-1). We hypothesize that similar superexchange mechanisms are common to all cross-conjugated D-B-A triads. Our results are compared to quantum interference effects on electron transfer/transport when cross-conjugated molecules are employed as the bridge or molecular wire component and suggest a mechanism by which electronic coupling (and therefore electron transfer/transport) can be modulated.}, number={39}, journal={Journal of the American Chemical Society}, publisher={American Chemical Society (ACS)}, author={Kirk, Martin L. and Shultz, David A. and Stasiw, Daniel E. and Habel-Rodriguez, Diana and Stein, Benjamin and Boyle, Paul D.}, year={2013}, month={Sep}, pages={14713–14725} } @article{han_allen_jonsson_johansson_mcowen_boyle_henderson_2013, title={Solvate Structures and Computational/Spectroscopic Characterization of Lithium Difluoro(oxalato)borate (LiDFOB) Electrolytes}, volume={117}, ISSN={["1932-7447"]}, DOI={10.1021/jp309102c}, abstractNote={Lithium difluoro(oxalato)borate (LiDFOB) is a relatively new salt designed for battery electrolyte usage. Limited information is currently available, however, regarding the ionic interactions of this salt (i.e., solvate formation) when it is dissolved in aprotic solvents. Vibrational spectroscopy is a particularly useful tool for identifying these interactions, but only if the vibrational bands can be correctly linked to specific forms of anion coordination. Single crystal structures of LiDFOB solvates have therefore been used to both explore the DFOB-center dot center dot center dot Li+ cation coordination interactions and serve as unambiguous models for the assignment of the Raman vibrational bands. The solvate crystal structures determined indude (monoglyme)(2):LiDFOB, (1,2-diethoxyethane)(3/2):LiDFOB, (acetonitrile)(3):LiDFOB, (acetonitrile)(1):LiDFOB, (dimethyl carbonate)(3/2):LiDFOB, (succinonitrile)(1):LiDFOB, (adiponitrile)(1):LiDFOB, (PMDETA)(1):LiDFOB, (CRYPT-222)(2/3):LiDFOB, and (propylene carbonate)(1):LiDFOB. DFT calculations have been incorporated to provide additional insight into the origin (i.e., vibrational modes) of the Raman vibrational bands to aid in the interpretation of the experimental analysis.}, number={11}, journal={JOURNAL OF PHYSICAL CHEMISTRY C}, author={Han, Sang-Don and Allen, Joshua L. and Jonsson, Erlendur and Johansson, Patrik and McOwen, Dennis W. and Boyle, Paul D. and Henderson, Wesley A.}, year={2013}, month={Mar}, pages={5521–5531} } @article{kirk_shultz_stasiw_lewis_wang_brannen_sommer_boyle_2013, title={Superexchange Contributions to Distance Dependence of Electron Transfer/Transport: Exchange and Electronic Coupling in Oligo(para-Phenylene)- and Oligo(2,5-Thiophene)-Bridged Donor–Bridge–Acceptor Biradical Complexes}, volume={135}, ISSN={0002-7863 1520-5126}, url={http://dx.doi.org/10.1021/JA4081887}, DOI={10.1021/ja4081887}, abstractNote={The preparation and characterization of three new donor-bridge-acceptor biradical complexes are described. Using variable-temperature magnetic susceptibility, EPR hyperfine coupling constants, and the results of X-ray crystal structures, we evaluate both exchange and electronic couplings as a function of bridge length for two quintessential molecular bridges: oligo(para-phenylene), β = 0.39 Å(-1) and oligo(2,5-thiophene), β = 0.22 Å(-1). This report represents the first direct comparison of exchange/electronic couplings and distance attenuation parameters (β) for these bridges. The work provides a direct measurement of superexchange contributions to β, with no contribution from incoherent hopping. The different β values determined for oligo(para-phenylene) and oligo(2,5-thiophene) are due primarily to the D-B energy gap, Δ, rather than bridge-bridge electronic couplings, H(BB). This is supported by the fact that the H(BB) values extracted from the experimental data for oligo(para-phenylene) (H(BB) = 11,400 cm(-1)) and oligo(2,5-thiophene) (12,300 cm(-1)) differ by <10%. The results presented here offer unique insight into the intrinsic molecular factors that govern H(DA) and β, which are important for understanding the electronic origin of electron transfer and electron transport mediated by molecular bridges.}, number={45}, journal={Journal of the American Chemical Society}, publisher={American Chemical Society (ACS)}, author={Kirk, Martin L. and Shultz, David A. and Stasiw, Daniel E. and Lewis, Geoffrey F. and Wang, Guangbin and Brannen, Candice L. and Sommer, Roger D. and Boyle, Paul D.}, year={2013}, month={Nov}, pages={17144–17154} } @article{harris_vezzu_bartolotti_boyle_huo_2013, title={Synthesis, Structure, Photophysics, and a DFT Study of Phosphorescent C*(NN)-N-boolean AND- and (CNN)-N-boolean AND-N-boolean AND-Coordinated Platinum Complexes}, volume={52}, ISSN={["1520-510X"]}, DOI={10.1021/ic400732g}, abstractNote={The reaction of N,N-diphenyl-2,2'-bipyridin-6-amine (L1) and N,N-diphenyl-6-(1H-pyrazol-1-yl)pyridin-2-amine (L2) with K2PtCl4 produced C*N(∧)N-coordinated cycloplatinated compounds with a five-six fused metallacycle 1a and 2a, respectively, which were then converted into their phenylacetylide derivatives 1b and 2b, respectively. Similar reactions starting from 2-phenyl-6-(1H-pyrazol-1-yl)pyridine (L3) produced C(∧)N(∧)N-coordinated platinum complexes 3a and 3b with a five-five-fused metallacycle. The structures of 1a, 1b, 2b, 3a, and 3b were determined by X-ray crystallography. The C*N(∧)N-coordinated platinum complexes are closer to a square geometry, whereas the C(∧)N(∧)N-coordinated complexes display a nearly perfect planar geometry. The π···π interactions were revealed in the crystal packing for 1a, 2b, and 3a with a π···π contact of 3.450, 3.422, and 3.414 Å, respectively. Two conformers were revealed in the crystal structure of 2b, one with the phenyl ring of the phenylacetylide being approximately parallel with the coordination plane and the other with the phenyl ring being approximately perpendicular to the coordination plane. Both 1a and 1b are weakly emissive in the red region. Complexes 2a and 3a are also weakly emissive, but their acetylide derivatives 2b and 3b emitted strongly green light at room temperature with quantum yields of 43 and 62%, respectively. DFT/TDDFT calculations were performed to elucidate the nature of their electronic transitions. The calculations suggested that lowest singlet and triplet excited states are characteristic of a mixed state involving one or more charge-transfer transitions such as ILCT, MLCT, and LLCT.}, number={20}, journal={INORGANIC CHEMISTRY}, author={Harris, Caleb F. and Vezzu, Dileep A. K. and Bartolotti, Libero and Boyle, Paul D. and Huo, Shouquan}, year={2013}, month={Oct}, pages={11711–11722} } @article{smeltz_lilly_boyle_ison_2013, title={The Electronic Nature of Terminal Oxo Ligands in Transition-Metal Complexes: Ambiphilic Reactivity of Oxorhenium Species}, volume={135}, ISSN={["1520-5126"]}, DOI={10.1021/ja401390v}, abstractNote={The synthesis of the Lewis acid-base adducts of B(C6F5)3 and BF3 with [DAAmRe(O)(X)] DAAm = N,N-bis(2-arylaminoethyl)methylamine; aryl = C6F5 (X = Me, 1, COCH3, 2, Cl, 3) as well as their diamidopyridine (DAP) (DAP=(2,6-bis((mesitylamino)methyl)pyridine) analogues, [DAPRe(O)(X)] (X = Me, 4, Cl, 5, I, 6, and COCH3,7), are described. In these complexes the terminal oxo ligands act as nucleophiles. In addition we also show that stoichiometric reactions between 3 and triarylphosphine (PAr3) result in the formation of triarylphosphine oxide (OPAr3). The electronic dependence of this reaction was studied by comparing the rates of oxygen atom transfer for various para-substituted triaryl phosphines in the presence of CO. From these experiments a reaction constant ρ = -0.29 was obtained from the Hammett plot. This suggests that the oxygen atom transfer reaction is consistent with nucleophilic attack of phosphorus on an electrophilic metal oxo. To the best of our knowledge, these are the first examples of mono-oxo d(2) metal complexes in which the oxo ligand exhibits ambiphilic reactivity.}, number={25}, journal={JOURNAL OF THE AMERICAN CHEMICAL SOCIETY}, author={Smeltz, Jessica L. and Lilly, Cassandra P. and Boyle, Paul D. and Ison, Elon A.}, year={2013}, month={Jun}, pages={9433–9441} } @article{kielar_wang_boyle_franz_2012, title={A boronate prochelator built on a triazole framework for peroxide-triggered tridentate metal binding}, volume={393}, ISSN={["0020-1693"]}, DOI={10.1016/j.ica.2012.06.011}, abstractNote={Iron chelating agents have the potential to minimize damage associated with oxidative stress in a range of diseases; however, this potential is countered by risks of indiscriminant metal binding or iron depletion in conditions not associated with systemic iron overload. Deferasirox is a chelator used clinically for iron overload, but also is cytotoxic to cells in culture. In order to test whether a prodrug version of deferasirox could minimize its cytotoxicity but retain its protective properties against iron-induced oxidative damage, we synthesized a prochelator that contains a self-immolative boronic ester masking group that is removed upon exposure to hydrogen peroxide to release the bis-hydroxyphenyltriazole ligand deferasirox. We present here the synthesis and characterization of this triazole-based, self-immolative prochelator: TIP (4-(5-(2-((4-boronobenzyl)oxy)phenyl)-3-(2-hydroxyphenyl)-1H-1,2,4-triazol-1-yl)benzoic acid). TIP does not coordinate to Fe(3+) and shows only weak affinity for Cu(2+) or Zn(2+), in stark contrast to deferasirox, which avidly binds all three metal ions. TIP converts efficiently in vitro upon reaction with hydrogen peroxide to deferasirox. In cell culture, TIP protects retinal pigment epithelial cells from death induced by hydrogen peroxide; however, TIP itself is more cytotoxic than deferasirox in unstressed cells. These results imply that the cytotoxicity of deferasirox may not derive exclusively from its iron withholding properties.}, journal={INORGANICA CHIMICA ACTA}, author={Kielar, Filip and Wang, Qin and Boyle, Paul D. and Franz, Katherine J.}, year={2012}, month={Dec}, pages={294–303} } @article{henderson_seo_zhou_boyle_shin_de long_trulove_passerini_2012, title={An Alternative Ionic Conductivity Mechanism for Plastic Crystalline Salt-Lithium Salt Electrolyte Mixtures}, volume={2}, ISSN={["1614-6840"]}, DOI={10.1002/aenm.201200130}, abstractNote={Abstract}, number={11}, journal={ADVANCED ENERGY MATERIALS}, author={Henderson, Wesley A. and Seo, Daniel M. and Zhou, Qian and Boyle, Paul D. and Shin, Joon-Ho and De Long, Hugh C. and Trulove, Paul C. and Passerini, Stefano}, year={2012}, month={Nov}, pages={1343–1350} } @inproceedings{han_allen_boyle_henderson_2012, title={Delving into the properties and solution structure of nitrile-lithium difluoro(oxalato)borate (LiDFOB) electrolytes for Li-ion batteries}, volume={41}, number={41}, booktitle={Rechargeable lithium and lithium ion batteries}, author={Han, S. D. and Allen, J. L. and Boyle, P. D. and Henderson, W. A.}, year={2012}, pages={47–51} } @article{seo_borodin_han_boyle_henderson_2012, title={Electrolyte Solvation and Ionic Association II. Acetonitrile-Lithium Salt Mixtures: Highly Dissociated Salts}, volume={159}, ISSN={["1945-7111"]}, DOI={10.1149/2.035209jes}, abstractNote={The electrolyte solution structure for acetonitrile (AN)-lithium salt mixtures has been examined for highly dissociated salts. Phase diagrams are reported for (AN)n-LiN(SO2CF3)2 (LiTFSI) and -LiPF6 electrolytes. Single crystal structures and Raman spectroscopy have been utilized to provide information regarding the solvate species present in the solid-state and liquid phases, as well as the average solvation number variation with salt concentration. Molecular dynamics (MD) simulations of the mixtures have been correlated with the experimental data to provide additional insight into the molecular-level interactions. Quantum chemistry (QC) calculations were performed on (AN)n-Li-(anion)m clusters to validate the ability of the developed many-body polarizable force field (used for the simulations) to accurately describe cluster stability (ionic association). The combination of these techniques provides tremendous insight into the solution structure within these electrolyte mixtures.}, number={9}, journal={JOURNAL OF THE ELECTROCHEMICAL SOCIETY}, author={Seo, Daniel M. and Borodin, Oleg and Han, Sang-Don and Boyle, Paul D. and Henderson, Wesley A.}, year={2012}, pages={A1489–A1500} } @article{seo_borodin_han_ly_boyle_henderson_2012, title={Electrolyte solvation and ionic association I. acetonitrile-lithium salt mixtures: Intermediate and highly associated salts}, volume={159}, number={5}, journal={Journal of the Electrochemical Society}, author={Seo, D. M. and Borodin, O. and Han, S. D. and Ly, Q. and Boyle, P. D. and Henderson, W. A.}, year={2012}, pages={A553–565} } @article{seo_boyle_borodin_henderson_2012, title={Li+ cation coordination by acetonitrile-insights from crystallography}, volume={2}, ISSN={["2046-2069"]}, DOI={10.1039/c2ra21290k}, abstractNote={Solvation is a critical factor for determining the properties of electrolytes and lithium reagents, but only limited information is available about the coordination number for Li+ cations in solution with different solvents. The present manuscript examines the manner in which acetonitrile (AN) fully solvates Li+ cations. The results are also likely pertinent to other nitrile and dinitrile solvents. In particular, the crystal structure for a (AN)6:LiPF6 solvate is reported—this is the first 6/1 AN/Li solvate structure to be determined. The structure consists of Li+ cations fully solvated by four AN molecules (i.e., [(AN)4Li]+ species), uncoordinated PF6− anions and uncoordinated AN molecules (two per Li+ cation). This structure validates, in part, density functional theory (DFT) calculations which predict that there is little to no energetic benefit to coordinating Li+ cations with more than four AN solvent molecules.}, number={21}, journal={RSC ADVANCES}, author={Seo, Daniel M. and Boyle, Paul D. and Borodin, Oleg and Henderson, Wesley A.}, year={2012}, pages={8014–8019} } @article{smeltz_boyle_ison_2012, title={Role of Low-Valent Rhenium Species in Catalytic Hydrosilylation Reactions with Oxorhenium Catalysts}, volume={31}, ISSN={["0276-7333"]}, DOI={10.1021/om300654q}, abstractNote={The catalytic competency of a Re(III) complex has been demonstrated. In the presence of silane, oxorhenium(V) catalysts are deoxygenated to produce species that are significantly more active than t...}, number={17}, journal={ORGANOMETALLICS}, author={Smeltz, Jessica L. and Boyle, Paul D. and Ison, Elon A.}, year={2012}, month={Sep}, pages={5994–5997} } @inproceedings{allen_seo_ly_boyle_henderson_2012, title={Solvent-LiBF4 phase diagrams, ionic association and solubility - cyclic carbonates and lactones}, volume={41}, number={41}, booktitle={Rechargeable lithium and lithium ion batteries}, author={Allen, J. L. and Seo, D. M. and Ly, Q. D. and Boyle, P. D. and Henderson, W. A.}, year={2012}, pages={41–45} } @article{lilly_boyle_ison_2012, title={Synthesis of Oxorhenium Acetyl and Benzoyl Complexes Incorporating Diamidopyridine Ligands: Implications for the Mechanism of CO Insertion}, volume={31}, ISSN={["0276-7333"]}, DOI={10.1021/om3002872}, abstractNote={A series of oxorhenium alkyl, phenyl, and vinyl complexes of the form [(DAP)Re(O)(R)] (R = aryl, vinyl, alkyl) was reported, and their reactivity with CO was examined. The methyl complex 5a reacts with CO at a significantly faster rate (2.5 h) than the phenyl complex 7a (24 h). Computational (B3PW91) studies reveal that although the acyl complex is the least stable (ΔG353 = −11.2 kcal/mol) with respect to CO insertion compared to the benzoyl complex (ΔG353 = −14.5 kcal/mol), the activation energy for CO insertion is lower for the methyl complex (ΔG⧧353 = 14.6 kcal/mol) than for the phenyl complex (ΔG⧧353 = 17.4 kcal/mol). This is consistent with the previously proposed mechanism, where CO inserts directly into the Re–R bond without prior formation of a CO adduct. The X-ray crystal structures of complexes 6, 7a, 8a, and 9a are reported.}, number={11}, journal={ORGANOMETALLICS}, author={Lilly, Cassandra P. and Boyle, Paul D. and Ison, Elon A.}, year={2012}, month={Jun}, pages={4295–4301} } @article{allen_han_boyle_henderson_2011, title={Crystal structure and physical properties of lithium difluoro(oxalato)borate (LiDFOB or LiBF(2)Ox)}, volume={196}, ISSN={["1873-2755"]}, DOI={10.1016/j.jpowsour.2011.07.065}, abstractNote={The structural characterization and properties of lithium difluoro(oxalato)borate (LiDFOB) are reported. LiDFOB was synthesized as previously described in the literature via direct reaction of boron trifluoride diethyl etherate with lithium oxalate. The crystal structure of the salt was determined from single crystal X-ray diffraction yielding a highly symmetric orthorhombic structure (Cmcm, a = 6.2623(8) Å, b = 11.4366(14) Å, c = 6.3002(7) Å, V = 451.22(9) Å3, Z = 4 at 110 K). Single crystal X-ray diffraction of a dihydrate of LiDFOB yielded a monoclinic structure (P21/c, a = 9.5580(3) Å, b = 12.7162(4) Å, c = 5.4387(2) Å, V = 634.63(4) Å3, Z = 4 at 110 K). Along with the crystal structures, additional structural information and the properties of LiDFOB (via 11B and 19F NMR, DSC, TGA and Raman spectroscopy) have been compared with those of LiBF4 and LiBOB to better understand the differences between these lithium battery electrolyte salts.}, number={22}, journal={JOURNAL OF POWER SOURCES}, author={Allen, Joshua L. and Han, Sang-Don and Boyle, Paul D. and Henderson, Wesley A.}, year={2011}, month={Nov}, pages={9737–9742} } @article{sloop_boyle_fountain_pearman_swann_2011, title={Electron-Deficient Aryl beta-Diketones: Synthesis and Novel Tautomeric Preferences}, volume={2011}, ISSN={["1099-0690"]}, DOI={10.1002/ejoc.201001451}, abstractNote={Abstract}, number={5}, journal={EUROPEAN JOURNAL OF ORGANIC CHEMISTRY}, author={Sloop, Joseph C. and Boyle, Paul D. and Fountain, Augustus W. and Pearman, William F. and Swann, Jacob A.}, year={2011}, month={Feb}, pages={936–941} } @article{vezzu_ravindranathan_gamer_bartolotti_smith_boyle_huo_2011, title={Highly Luminescent Tridentate (NC)-C-boolean AND*N Platinum(II) Complexes Featured in Fused Five-Six-Membered Metallacycle and Diminishing Concentration Quenching}, volume={50}, ISSN={["1520-510X"]}, DOI={10.1021/ic200794b}, abstractNote={A series of cyclometalating ligands, N-phenyl-N-(3-(pyridin-2-yl)phenyl)pyridin-2-amine (L1), N-(3-(1H-pyrazol-1-yl)phenyl)-N-phenylpyridin-2-amine (L2), N-phenyl-N-(3-(quinolin-2-yl)phenyl)pyridin-2-amine (L3), N-phenyl-N-(3-(pyridin-2-yl)phenyl)quinolin-2-amine (L4), N-(3-(isoquinolin-1-yl)phenyl)-N-phenylpyridin-2-amine (L5), and N-phenyl-N-(3-(pyridin-2-yl)phenyl)isoquinolin-1-amine (L6), were synthesized, which reacted with K(2)PtCl(4) in glacial acetic acid to produce N^C*N-coordinated platinum(II) complexes featured in a fused five-six-membered metallacycle, 1-6, respectively. The structures of 1, 3, 4, and 6 were determined by single crystal X-ray crystallography. The square geometries of the complexes are improved when compared with those of the N^C^N-coordinated complexes as the bite angles for the platinum in N^C*N-coordinated complexes 1, 3, and 4 are increased. The Pt-C bonds (1.94-1.95 Å) are shorter than those of C^N^N-coordinated platinum complexes but longer than those found for N^C^N-coordinated platinum complexes. With the increase of the steric interaction, the distortion of the molecules from a planar coordination geometry becomes more and more severe from 1 to 3 to 4 and 6, and in 6, the N-phenyl ring has to stand up on the coordination sphere to minimize the steric interaction with the N-isoquinolyl ring. The photophysical properties of the complexes were studied, and their absorption and emission spectra were interpreted by relating to the structural features revealed by the X-ray crystal structures and the orbital characters predicted by DFT calculations. All complexes are emissive in fluid at room temperature, and the quantum yields (up to 0.65) are comparable to those of highly emissive N^C^N-coordinated platinum complexes. Self-quenching was not observed in the concentration range of 10(-6) to 10(-4) M. Large rigidochromic shifts for the emissions of 2, 4, and 6 upon cooling from room temperature to rigid glass (77 K) were observed. Two different triplet states that control the emissions were proposed to account for the photophysical properties of 6.}, number={17}, journal={INORGANIC CHEMISTRY}, author={Vezzu, Dileep A. K. and Ravindranathan, Deepak and Gamer, Alexander W. and Bartolotti, Libero and Smith, Meredith E. and Boyle, Paul D. and Huo, Shouquan}, year={2011}, month={Sep}, pages={8261–8273} } @article{allen_boyle_henderson_2011, title={Lithium difluoro(oxalato)borate tetramethylene sulfone disolvate}, volume={67}, journal={Acta Crystallographica. Section E, Structure Reports Online}, author={Allen, J. L. and Boyle, P. D. and Henderson, W. A.}, year={2011}, pages={M533–305} } @article{smeltz_boyle_ison_2011, title={Mechanism for the Activation of Carbon Monoxide via Oxorhenium Complexes}, volume={133}, ISSN={["0002-7863"]}, DOI={10.1021/ja205477w}, abstractNote={Activation of CO by the rhenium(V) oxo complex [(DAAm)Re(O)(CH(3))] (1) [DAAm = N,N-bis(2-arylaminoethyl)methylamine; aryl = C(6)F(5), Mes] resulted in the isolation of the rhenium(III) acetate complex [(DAAm)Re(O(2)CCH(3))(CO)] (3). The mechanistic details of this reaction were explored experimentally. The novel oxorhenium(V) acyl intermediate [(DAAm)Re(O)(C(O)CH(3))] (2) was isolated, and its reactivity with CO was investigated. An unprecedented mechanism is proposed: CO is activated by the metal oxo complex 1 and inserted into the rhenium-methyl bond to yield acyl complex 2, after which subsequent migration of the acyl ligand to the metal oxo ligand yields acetate complex 3. X-ray crystal structures of 2 and 3 are reported.}, number={34}, journal={JOURNAL OF THE AMERICAN CHEMICAL SOCIETY}, author={Smeltz, Jessica L. and Boyle, Paul D. and Ison, Elon A.}, year={2011}, month={Aug}, pages={13288–13291} } @article{zhou_boyle_malpezzi_mele_shin_passerini_henderson_2011, title={Phase Behavior of Ionic Liquid-LiX Mixtures: Pyrrolidinium Cations and TFSI- Anions - Linking Structure to Transport Properties}, volume={23}, ISSN={["0897-4756"]}, DOI={10.1021/cm201427k}, abstractNote={The thermal phase behavior and ionic conductivity of mixtures of N-alkyl-N-methylpyrrolidinium bis(trifluoromethanesulfonyl)imide ionic liquids (PY1RTFSI where R = 4 or 5 for butyl or pentyl) with LiTFSI have been examined as model systems for ionic liquid-based lithium battery electrolytes. Several mixed salt crystalline phases form. The ionic conductivity variability of the mixtures correlates well with the reported phase behavior. The crystal structures for the (1-x) PY1RTFSI-(x) LiTFSI (x = 0.67) (or 1/2 PY1RTFSI/LiTFSI with R = 4 or 5) phases have been determined. These phases are isostructural, consisting of ionic planar layers formed by Li+···TFSI– linkages in which the uncoordinated PY1R+ cations reside in cavities within the layers. These structures have been used to aid in correlating Raman data with cation–anion coordination in the IL-LiTFSI binary salt mixtures.}, number={19}, journal={CHEMISTRY OF MATERIALS}, author={Zhou, Qian and Boyle, Paul D. and Malpezzi, Luciana and Mele, Andrea and Shin, Joon-Ho and Passerini, Stefano and Henderson, Wesley A.}, year={2011}, month={Oct}, pages={4331–4337} } @article{seo_boyle_henderson_2011, title={Poly[[(acetonitrile)lithium(I)]-mu(3)-tetrafluoridoborato]}, volume={67}, ISSN={["1600-5368"]}, DOI={10.1107/s1600536811012141}, abstractNote={The structure of the title compound, [Li(BF4)(CH3CN)]n, consists of a layered arrangement parallel to (100) in which the Li+ cations are coordinated by three F atoms from three tetrafluoridoborate (BF4 −) anions and an N atom from an acetonitrile molecule. The BF4 − anion is coordinated to three different Li+ cations though three F atoms. The structure can be described as being built from vertex-shared BF4 and LiF3(NCCH3) tetrahedra. These tetrahedra reside around a crystallographic inversion center and form 8-membered rings.}, journal={ACTA CRYSTALLOGRAPHICA SECTION E-STRUCTURE REPORTS ONLINE}, author={Seo, Daniel M. and Boyle, Paul D. and Henderson, Wesley A.}, year={2011}, month={May}, pages={M547–U443} } @article{seo_boyle_henderson_2011, title={Poly[bis(acetonitrile-kappa N)bis[mu(3)-bis(trifluoromethanesulfonyl)imido-kappa O-4,O ':O '':O ''']dilithium]}, volume={67}, journal={Acta Crystallographica. Section E, Structure Reports Online}, author={Seo, D. M. and Boyle, P. D. and Henderson, W. A.}, year={2011}, pages={M534–317} } @article{allen_boyle_henderson_2011, title={Poly[diacetonitrile[mu(3)-difluoro(oxalato)borato]sodium]}, volume={67}, journal={Acta Crystallographica. Section E, Structure Reports Online}, author={Allen, J. L. and Boyle, P. D. and Henderson, W. A.}, year={2011}, pages={M678–388} } @article{webb_munro-leighton_pierpont_gurkin_gunnoe_cundari_sabat_petersen_boyle_2011, title={Pt(II) and Pt(IV) Amido, Aryloxide, and Hydrocarbyl Complexes: Synthesis, Characterization, and Reaction with Dihydrogen and Substrates that Possess C-H Bonds}, volume={50}, ISSN={["1520-510X"]}, DOI={10.1021/ic200153n}, abstractNote={The Pt(II) amido and phenoxide complexes ((t)bpy)Pt(Me)(X), ((t)bpy)Pt(X)(2), and [((t)bpy)Pt(X)(py)][BAr'(4)] (X = NHPh, OPh; py = pyridine) have been synthesized and characterized. To test the feasibility of accessing Pt(IV) complexes by oxidizing their Pt(II) precursors, the previously reported ((t)bpy)Pt(R)(2) (R = Me and Ph) systems were oxidized with I(2) to yield ((t)bpy)Pt(R)(2)(I)(2). The analogous reaction with ((t)bpy)Pt(Me)(NHPh) and MeI yields the corresponding ((t)bpy)Pt(Me)(2)(NHPh)(I) complex. Reaction of ((t)bpy)Pt(Me)(NHPh) and phenylacetylene at 80 °C results in the formation of the Pt(II) phenylacetylide complex ((t)bpy)Pt(Me)(C≡CPh). Kinetic studies indicate that the reaction of ((t)bpy)Pt(Me)(NHPh) and phenylacetylene occurs via a pathway that involves [((t)bpy)Pt(Me)(NH(2)Ph)][TFA] as a catalyst. The reaction of H(2) with ((t)bpy)Pt(Me)(NHPh) ultimately produces aniline, methane, (t)bpy, and elemental Pt. For this reaction, mechanistic studies reveal that 1,2-addition of dihydrogen across the Pt-NHPh bond to initially produce ((t)bpy)Pt(Me)(H) and free aniline is catalyzed by elemental Pt. Heating the cationic complexes [((t)bpy)Pt(NHPh)(py)][BAr'(4)] and [((t)bpy)Pt(OPh)(py)][BAr'(4)] in C(6)D(6) does not result in the production of aniline and phenol, respectively. Attempted synthesis of a cationic system analogous to [((t)bpy)Pt(NHPh)(py)][BAr'(4)] with ligands that are more labile than pyridine (e.g., NC(5)F(5)) results in the formation of the dimer [((t)bpy)Pt(μ-NHPh)](2)[BAr'(4)](2). Solid-state X-ray diffraction studies of the complexes ((t)bpy)Pt(Me)(NHPh), [((t)bpy)Pt(NH(2)Ph)(2)][OTf](2), ((t)bpy)Pt(NHPh)(2), ((t)bpy)Pt(OPh)(2), ((t)bpy)Pt(Me)(2)(I)(2), and ((t)bpy)Pt(Ph)(2)(I)(2) are reported.}, number={9}, journal={INORGANIC CHEMISTRY}, author={Webb, Joanna R. and Munro-Leighton, Colleen and Pierpont, Aaron W. and Gurkin, Joshua T. and Gunnoe, T. Brent and Cundari, Thomas R. and Sabat, Michal and Petersen, Jeffrey L. and Boyle, Paul D.}, year={2011}, month={May}, pages={4195–4211} } @article{lilly_boyle_ison_2011, title={Synthesis and characterization of oxorhenium(V) diamido pyridine complexes that catalyze oxygen atom transfer reactions}, volume={40}, ISSN={["1477-9226"]}, DOI={10.1039/c1dt11143d}, abstractNote={The detailed syntheses of complexes 1-4, Re(O)(X)(DAP) (X = Me, 1; Cl, 2; I, 3; OTf (triflate), 4) incorporating the diamido pyridine (DAP) ancillary ligand (2,6-bis((mesitylamino)methyl)pyridine) are described and shown to be effective catalysts for oxygen atom transfer (OAT) reactions of PyO to PPh(3). The catalytic activities are as follows: 4≈3 > 2 > 1. The observed electronic trend is consistent with the turnover limiting reduction of the proposed Re(VII) dioxo intermediate, Re(O)(2)(X)(DAP), during the catalytic cycle. The catalytic activity of complexes 1-3 was compared to previously published diamido amine (DAAm) oxorhenium complexes of the type Re(O)(X)(DAAm) (X = Me, 5; Cl, 6; I, 7 and DAAm = N,N-bis(2-arylaminoethyl)methylamine) which exhibit hydrolytic degradation during the catalytic reaction. Complexes 1-3 displayed higher turnover frequencies compared to 5-7. This higher catalytic activity was attributed to the more rigid DAP ligand backbone, which makes the complexes less susceptible to decomposition. However, another decomposition pathway was proposed for this catalytic system due to the observation of Re(O)(3)((MesNCH(2))(MesNCH)NC(5)H(3)) 8 in which one arm of the DAP ligand is oxidized.}, number={44}, journal={DALTON TRANSACTIONS}, author={Lilly, Cassandra P. and Boyle, Paul D. and Ison, Elon A.}, year={2011}, pages={11815–11821} } @article{sloop_shultz_coote_shepler_sullivan_kampf_boyle_2011, title={Synthesis of and structure-property relationships in zinc complexes of bis-metaphenylene semiquinone biradical species}, volume={25}, ISSN={0894-3230}, url={http://dx.doi.org/10.1002/poc.1917}, DOI={10.1002/poc.1917}, abstractNote={The design of novel, functionalized bis‐metaphenylene semiquinone (SQ) ligands and their corresponding metal complexes which combine conformational flexibility and electron‐withdrawing, electron‐donating, and conjugating substituents enable investigation of multiple structure–property relationships. Along these lines, we report the synthesis of three new bis(ZnII(SQ)TpCum,Me) complexes containing the bis‐metaphenylene coupling fragment. Using electron paramagnetic resonance spectroscopy, ab initio computations and superconducting quantum interference device magnetometry, we show how spin‐density is affected by the bis‐metaphenylene system substituents. Copyright © 2011 John Wiley & Sons, Ltd.}, number={4}, journal={Journal of Physical Organic Chemistry}, publisher={Wiley}, author={Sloop, Joseph C. and Shultz, David A. and Coote, Tashni and Shepler, Benjamin and Sullivan, Ubie and Kampf, Jeff W. and Boyle, Paul D.}, year={2011}, month={Aug}, pages={314–321} } @article{seo_boyle_henderson_2011, title={Tetrakis(acetonitrile-kappa N) lithium hexafluoridophosphate acetonitrile monosolvate}, volume={67}, journal={Acta Crystallographica. Section E, Structure Reports Online}, author={Seo, D. M. and Boyle, P. D. and Henderson, W. A.}, year={2011}, pages={M1148–1301} } @article{jordan_boyle_sargent_allen_2010, title={Binding of Carboxylic Acids by Fluorescent Pyridyl Ureas}, volume={75}, ISSN={["1520-6904"]}, DOI={10.1021/jo101730w}, abstractNote={Fluorescent pyrid-2-yl ureas were prepared by treating halogenated 2-aminopyridines with hexyl isocyanate, followed by Sonogashira coupling with arylacetylenes. The sensors emit light of ∼360 nm with quantum yields of 0.05-0.1 in acetonitrile solution. Addition of strong organic acids (pK(a) < 13 in CH(3)CN) shifts the fluorescence band to lower energy, and clean isoemissive behavior is observed. Fluorescence response curves (i.e., F/F(0) vs [acid](total)) are hyperbolic in shape for CCl(3)COOH and CF(3)COOH, with association constants on the order of 10(3) M(-1) for both acids. (1)H NMR titrations and DFT analyses indicate that trihaloacetic acids bind in ionized form to the receptors. Pyridine protonation disrupts an intramolecular H-bond, thereby unfolding an array of ureido NH donors for recognition of the corresponding carboxylates. Methanesulfonic acid protonates the sensors, but no evidence for conjugate base binding at the urea moiety is found by NMR. An isosteric control compound that lacks an integrated pyridine does not undergo significant fluorescence changes upon acidification.}, number={24}, journal={JOURNAL OF ORGANIC CHEMISTRY}, author={Jordan, Lisa M. and Boyle, Paul D. and Sargent, Andrew L. and Allen, William E.}, year={2010}, month={Dec}, pages={8450–8456} } @article{schmidt_shultz_martin_boyle_2010, title={Goldilocks Effect in Magnetic Bistability: Remote Substituent Modulation and Lattice Control of Photoinduced Valence Tautomerism and Light-Induced Thermal Hysteresis}, volume={132}, ISSN={0002-7863 1520-5126}, url={http://dx.doi.org/10.1021/ja101957c}, DOI={10.1021/ja101957c}, abstractNote={The thermal-induced and photoinduced valence tautomerism of a series of Co(dioxolene)(2)(4-X-py)(2) complexes (dioxolene = 3,5-di-tert-butylcatecholate or 3,5-di-tert-butylsemiquinonate; 4-X-py = 4-(X)pyridine, X = H (1), OMe (2), Me (3), CN (4), Br (5), NO(2) (6)) is described. The thermal valence tautomerism (ls-Co(III)(SQ)(Cat)(4-X-py)(2) <--> hs-Co(II)(SQ)(SQ)(4-X-py)(2)) is only observed for complexes 4, 5, and 6 where each is accompanied by a hysteresis loop of ca. 5 K. When a crystalline sample of 4-6 is held at 10 K in a SQUID magnetometer and irradiated with white light (lambda = 400-850 nm), the hs-Co(II) tautomer is formed. When the light source is removed, and the sample is slowly heated, the hs-Co(II) tautomer persists until ca. 90 K, approximately 40 K higher than the thermal stability of previously reported complexes. Heating and cooling the sample while maintaining irradiation results in the appearance of a new light-induced thermal hysteresis loop below 90 K (DeltaT = ca. 25 K). Below 50 K, the hs-Co(II) tautomer displays temperature-independent relaxation to the ls-Co(III) form, and above 50 K, the relaxation is thermally activated with an activation energy E(a) > ca. 1500 cm(-1). The coordination geometry (trans-pyridines), pyridine substitution, and crystal packing forces conspire to create the comparatively thermally stable photogenerated hs-Co(II) tautomer, thus providing an excellent handle for molecular and crystal engineering studies.}, number={17}, journal={Journal of the American Chemical Society}, publisher={American Chemical Society (ACS)}, author={Schmidt, R. D. and Shultz, D. A. and Martin, J. D. and Boyle, P. D.}, year={2010}, month={May}, pages={6261–6273} } @article{ravindranathan_vezzu_bartolotti_boyle_huo_2010, title={Improvement in Phosphorescence Efficiency through Tuning of Coordination Geometry of Tridentate Cyclometalated Platinum(II) Complexes}, volume={49}, ISSN={["1520-510X"]}, DOI={10.1021/ic101109h}, abstractNote={A series of tridentate cyclometalated platinum(II) complexes (C(∧)N*N)PtL (L = Cl or acetylide) featuring a fused five-six-membered metallacycle were synthesized. The structure of the complexes was confirmed by X-ray crystallography. In contrast to the C(∧)N(∧)N platinum complexes with a fused five-five-membered metallacycle, the platinum coordination in C(∧)N*N complexes is much closer to a square planar geometry. The photophysical properties of the complexes were studied. The geometrical change from C(∧)N(∧)N to C(∧)N*N led to a substantial improvement in phosphorescence efficiency of the complexes with an acetylide ligand in solution at room temperature. For example, the quantum yield of (C(∧)N*N)PtCCPh was measured to be 56%, demonstrating a big jump from 4% reported for (C(∧)N(∧)N)PtCCPh.}, number={19}, journal={INORGANIC CHEMISTRY}, author={Ravindranathan, Deepak and Vezzu, Dileep A. K. and Bartolotti, Libero and Boyle, Paul D. and Huo, Shouquan}, year={2010}, month={Oct}, pages={8922–8928} } @article{kizewski_boyle_hesterberg_martin_2010, title={Mixed Anion (Phosphate/Oxalate) Bonding to Iron(III) Materials}, volume={132}, ISSN={["1520-5126"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-77749237327&partnerID=MN8TOARS}, DOI={10.1021/ja908807b}, abstractNote={A novel phosphate/oxalate inorganic-organic hybrid material has been prepared to elucidate synthesis and bonding characteristics of iron(III) with both phosphate and organic matter (OM). Such mixed anion bonding of inorganic oxyanions and OM to iron(III) and aluminum(III) in environmental systems has been proposed but not proven, mainly because of the complexity of natural geochemical matrices. The compound reported here with the molecular formula of [C(3)H(12)N(2)](2)[Fe(5)(C(2)O(4))(2)(H(x)PO(4))(8)] (I) was hydrothermally synthesized and characterized by single crystal X-ray diffraction and X-ray absorption spectroscopy (XAS). In this new structure, Fe-O octahedra and P-O tetrahedra are connected by corner-sharing to form a 2-D network in the a-b plane. Oxalate anions cross-link these Fe-P layers constructing a 3-D anionic framework. A diprotonated structure-directing template, DAP (1,3-diaminopropane), resides in the oxalate layer of the structure and offsets the negative charge of the anionic framework. Iron K-edge XANES spectra confirmed that the iron in I is Fe(III). The crystal structure of I is used to successfully fit its Fe K-edge EXAFS spectrum, which exhibits spectral signatures that unambiguously identify iron-phosphate and iron-OM bonding. Such molecular spectroscopic features will be invaluable for the evaluation of complex environmental systems. Furthermore, syntheses demonstrated the critical role of the templating amine to mediate whether or not the iron(III) is reduced by the organic acid.}, number={7}, journal={JOURNAL OF THE AMERICAN CHEMICAL SOCIETY}, author={Kizewski, Fiona R. and Boyle, Paul and Hesterberg, Dean and Martin, James D.}, year={2010}, month={Feb}, pages={2301–2308} } @article{webb_pierpont_munro-leighton_gunnoe_cundari_boyle_2010, title={Net Hydrogenation of Pt-NHPh Bond Is Catalyzed by Elemental Pt}, volume={132}, ISSN={["1520-5126"]}, DOI={10.1021/ja9102309}, abstractNote={Synthesis and characterization of the monomeric complex ((t)bpy)Pt(Me)(NHPh) ((t)bpy = 4,4'-di-tert-butyl-2,2'-dipyridyl) has been accomplished. Mechanistic studies reveal that 1,2-addition of dihydrogen across the Pt-anilido bond to initially produce ((t)bpy)Pt(Me)(H) and free aniline is catalyzed by elemental Pt rather than through a pathway that involves direct activation of H(2) by Pt and 1,2-addition across the Pt-NHPh bond.}, number={13}, journal={JOURNAL OF THE AMERICAN CHEMICAL SOCIETY}, author={Webb, Joanna R. and Pierpont, Aaron W. and Munro-Leighton, Colleen and Gunnoe, T. Brent and Cundari, Thomas R. and Boyle, Paul D.}, year={2010}, month={Apr}, pages={4520-+} } @article{zhou_fitzgerald_boyle_henderson_2010, title={Phase Behavior and Crystalline Phases of Ionic Liquid-Lithium Salt Mixtures with 1-Alkyl-3-methylimidazolium Salts}, volume={22}, ISSN={["1520-5002"]}, DOI={10.1021/cm902691v}, abstractNote={The thermal phase behavior of 1-alkyl-3-methylimidazolium bis(trifluoromethanesulfonyl)imide (IM10RTFSI where R = 1, 2, or 4 for methyl, ethyl or butyl, respectively) ionic liquid binary mixtures with LiTFSI have been investigated as models for electrolytes for lithium batteries. Diverse phase behavior is found with significant variations noted from similar mixtures in which the imidazolium cations are replaced with N-alkyl-N-methylpyrrolidinium cations. The crystal structure for a (1−x) IM101TFSI-(x) LiTFSI (x = 0.50) (or 1/1 IM101TFSI/LiTFSI) phase is reported to further clarify the molecular level interactions occurring in these binary salt mixtures.}, number={3}, journal={CHEMISTRY OF MATERIALS}, author={Zhou, Qian and Fitzgerald, Kendall and Boyle, Paul D. and Henderson, Wesley A.}, year={2010}, month={Feb}, pages={1203–1208} } @article{taniguchi_mass_boyle_tang_diers_bocian_holten_lindsey_2010, title={Structural studies of sparsely substituted synthetic chlorins and phorbines establish benchmarks for changes in the ligand core and framework of chlorophyll macrocycles}, volume={979}, ISSN={["1872-8014"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-77955427819&partnerID=MN8TOARS}, DOI={10.1016/j.molstruc.2010.05.035}, abstractNote={Understanding the changes in molecular structure of tetrapyrrole macrocycles upon derivatization of the organic framework is essential for diverse studies ranging from metal complexation to formation of supramolecular assemblies. New, sparsely substituted free base chlorin, 17-oxochlorin, phorbine and 131-oxophorbine macrocycles provide benchmarks for naturally occurring hydroporphyrins and have been examined here by X-ray crystallography, resonance Raman spectroscopy, and density functional theoretical (DFT) calculations. The macrocycles contain no substituents other than a geminal-dimethyl group attached to the reduced, pyrroline ring. The X-ray studies indicate that the benchmark compounds exhibit only slight distortion from planarity, which increases along the series porphine < chlorin < oxochlorin < phorbine < oxophorbine. The elongated CβCβ bond distance due to sp3 versus sp2 hybridization in the pyrroline ring (ring D) of the (oxo)chlorins and (oxo)phorbines (1.52–1.54 Å) versus that of porphine (1.35 Å) is accompanied by altered bond angles in ring D. Introduction of ring E (exocyclic ring) in a chlorin to give the phorbine or oxophorbine causes alteration of the bond angles at many sites in the framework of the macrocycle; for example, the bond angles of N3C14C15 in the (oxo)phorbine are widened by ∼11° compared to those of porphine or the analogous sparsely substituted chlorin. As a result, the shape of the macrocycle core changes along the series of porphine (nearly square), (oxo)chlorin (kite-shaped), and (oxo)phorbine (right-angled trapezoid), and the core size increases in the order porphine < phorbine ∼ oxophorbine < oxochlorin ∼ chlorin. Comparison of the bond distances and angles in ring E of the phorbine versus oxophorbine indicates that the shortening of the C13C131 bond owing to the presence of the oxo-group is quite small, only 0.024 Å; thus, the unsymmetrical structure of ring E does not appear to be due to conjugation with the C131O group but may be a characteristic feature of the (oxo)phorbine framework. The X-ray data further indicate that the lengths of the oxochlorin C17O and oxophorbine C131O groups are essentially identical, a result also predicted by DFT calculations. Regardless, the observed frequencies for the stretching vibrations of the C17O (1721 cm−1) and C131O (1701 cm−1) groups are different and suggest that conjugation of the latter group with the π-system of the macrocycle is greater than that of the former group. Collectively, the studies provide new insights into the individual factors that give rise to the overall structural characteristics of various macrocycles.}, number={1-3}, journal={JOURNAL OF MOLECULAR STRUCTURE}, publisher={Elsevier BV}, author={Taniguchi, Masahiko and Mass, Olga and Boyle, Paul D. and Tang, Qun and Diers, James R. and Bocian, David F. and Holten, Dewey and Lindsey, Jonathan S.}, year={2010}, month={Aug}, pages={27–45} } @article{wan_niu_behof_wang_boyle_gorman_2009, title={Aminoisoquinolines as colorimetric Hg2+ sensors: the importance of molecular structure and sacrificial base}, volume={65}, ISSN={["0040-4020"]}, DOI={10.1016/j.tet.2009.03.063}, abstractNote={Here it is shown that 3-phenyl-2-amino isoquinoline acts as a simple mercury sensor. It is simple to synthesize. The molecule requires base/buffer to bind in a 1:1 stoichiometry with mercury ion, however. Otherwise, it acts as a sacrificial base, presumably to pick up a proton liberated during binding. This binding is characterized quantitatively.}, number={22}, journal={TETRAHEDRON}, author={Wan, Yanjun and Niu, Weijun and Behof, William J. and Wang, Yifei and Boyle, Paul and Gorman, Christopher B.}, year={2009}, month={May}, pages={4293–4297} } @article{foley_abernethy_gunnoe_hill_boyle_sabat_2009, title={Chlorination of Boron on a Ruthenium-Coordinated Hydridotris(pyrazolyl)borate (Tp) Ligand: A Caveat for the Use of TpRu(PPh3)(2)Cl}, volume={28}, ISSN={["1520-6041"]}, DOI={10.1021/om8008074}, abstractNote={A side-product that accompanies the synthesis of the widely utilized starting material TpRu(PPh3)2Cl (Tp = hydridotris(pyrazolyl)borate) has been identified as the complex ClTpRu(PPh3)2H (ClTp = chlorotris(pyrazolyl)borate), which provides a rare example of a boron-halogenated pyrazolylborate ligand. The reaction of ClTpRu(PPh3)2H with dichloromethane or chloroform quantitatively produce ClTpRu(PPh3)2Cl.}, number={1}, journal={ORGANOMETALLICS}, author={Foley, Nicholas A. and Abernethy, Robyn J. and Gunnoe, T. Brent and Hill, Anthony F. and Boyle, Paul D. and Sabat, Michal}, year={2009}, month={Jan}, pages={374–377} } @article{lee_ke_ramirez_gunnoe_cundari_boyle_petersen_2009, title={Six-, Five-, and Four-Coordinate Ruthenium(II) Hydride Complexes Supported by N-Heterocyclic Carbene Ligands: Synthesis, Characterization, Fundamental Reactivity, and Catalytic Hydrogenation of Olefins, Aldehydes, and Ketones}, volume={28}, ISSN={["1520-6041"]}, DOI={10.1021/om801111c}, abstractNote={The Ru(II) hydride complex (IMes)2Ru(Cl)(H)(CO) (1) {IMes = 1,3-bis-(2,4,6-trimethylphenyl)imidazol-2-ylidene} was synthesized from [Ru(CO)2Cl2]n and free IMes. Complex 1 rapidly reacts with CO to produce the cis-dicarbonyl Ru(II) complex (IMes)2Ru(Cl)(H)(CO)2 (2). The reaction of 1 with NaBAr′4 {Ar′ = 3,5-(CF3)C6H3} produces the four-coordinate Ru(II) cationic complex [(IMes)2Ru(H)(CO)][BAr′4] (4), which can be trapped by two equivalents of tert-butylisonitrile to produce [(IMes)2Ru(H)(CO)(CNtBu)2][BAr′4] (5). Experimental and computational studies suggest that complex 4 is a diamagnetic system that adopts a sawhorse structure. The hydride ligand of complex 2 is readily displaced as dihydrogen upon reaction with HCl to produce (IMes)2Ru(Cl)2(CO)2 (3). Both complex 1 and 4 were found to react with D2 (30 psi) at room temperature to produce the isotopomers (IMes)2Ru(Cl)(D)(CO) (1-d1) and [(IMes)2Ru(D)(CO)][BAr′4] (4-d1), respectively, with the rate of formation of 4-d1 at least 28 times faster than the con...}, number={6}, journal={ORGANOMETALLICS}, author={Lee, John P. and Ke, Zhuofeng and Ramirez, Magaly A. and Gunnoe, T. Brent and Cundari, Thomas R. and Boyle, Paul D. and Petersen, Jeffrey L.}, year={2009}, month={Mar}, pages={1758–1775} } @article{feng_aponte_houseworth_boyle_ison_2009, title={Synthesis of Oxorhenium(V) Complexes with Diamido Amine Ancillary Ligands and Their Role in Oxygen Atom Transfer Catalysis}, volume={48}, ISSN={["1520-510X"]}, DOI={10.1021/ic901434u}, abstractNote={The detailed syntheses of complexes of the form [Re(O)(X)(RNCH(2)CH(2))(2)N(Me)] (X = Me, Cl, I, R = mesityl, C(6)F(5)), 1-3, incorporating diamidoamine ancillary ligands are described. X-ray crystal structures for the complexes [Re(O)(Me)((C(6)F(5))NCH(2)CH(2))(2)N(Me)], 1a, [Re(O)(I)((C(6)F(5))N CH(2)CH(2))(2)N(Me)], 3a, and [Re(O)(I)((Mes)NCH(2))(2)N(Me)], 3b, are reported. The geometry about the metal center in 1a is best described as a severely distorted square pyramid with the oxo ligand in the apical position. In contrast, the geometry about the metal center in 3a is best described as a severely distorted trigonal bipyramid, with the iodo ligand occupying the apical position and the diamido nitrogens and the oxo ligand occupying the equatorial plane. The catalytic activities of these complexes for oxygen atom transfer, OAT, from pyridine-N-oxides, PyO, to PPh(3) were also examined. The reactions exhibited a clear dependence on the diamido ligand substituent and the X ligand (Me, I, Cl) attached to the metal, with the combined effect that electron-withdrawing substituents on the diamido ligand and poor sigma donors directly attached to the metal center increases the rate of catalytic activity. The kinetics of OAT from pyridine-N-oxides to Re were also investigated. The reactions displayed clean first order kinetics in Re and saturation kinetics for the dependence on PyO. Changing the PyO substrate had no effect on the saturation value, k(sat), suggesting that the OAT reaction in these five-coordinate complexes appears to be governed by isomerization of the starting complex. Attempts to isolate a postulated Re(VII) intermediate were not successful because of hydrolytic degradation. The product of hydrolytic degradation [((C(6)F(5))N(H)CH(2)CH(2)))(2)NH(Me)][X], (X = ReO(4)(-), or I(-)), 4 can be isolated, and its X-ray crystal structure is reported. Although the Re(VII) intermediate could not be isolated, its activity in OAT reactions was probed by competition experiments with PPh(3) and four para-substituted triarylphosphines (p-X-Ph)(3)P (X = OMe, Me, Cl, CF(3)). These experiments led to a Hammett that yielded a reaction constant of rho = -0.30 +/- 0.01. This data suggests a positive charge buildup on phosphorus for the OAT reaction and is consistent with the nucleophilic attack of phosphorus on an electrophilic metal oxo.}, number={23}, journal={INORGANIC CHEMISTRY}, author={Feng, Yuee and Aponte, Joel and Houseworth, Paul J. and Boyle, Paul D. and Ison, Elon A.}, year={2009}, month={Dec}, pages={11058–11066} } @article{he_el-shafei_freeman_boyle_2009, title={X-ray and molecular modelling studies of 4-[N-alkylamino]azobenzene dyes}, volume={82}, ISSN={["1873-3743"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-64849101647&partnerID=MN8TOARS}, DOI={10.1016/j.dyepig.2009.01.013}, abstractNote={Abstract 4-[N,N-Bisalkyl]amino-2′-chloro-4′-nitroazobenzenes were recrystallized from acetone in either triclinic or monoclinic cells with the space group P-1 or P21/c. The asymmetric unit cell of dyes having at least one N-cyanoethyl group contained two molecules that were symmetrically unequivalent. The aromatic rings in the azobenzene skeleton were essentially planar with respect to the plane of the azo group, although the C6–C1–N1–N2 torsion angle was 13.2° when the N,N-dicyanoethyl group was employed. X-ray studies were used as a basis for assessing the utility of nonlocal DFT calculations in predicting the equilibrium molecular geometry and solvatochromic properties of the compounds using MM3/ZINDO-S and the COSMO Solvation Model. Although B3LYP and PBE energy functionals were comparable in predicting bond lengths, PBE was slightly better than B3LYP in predicting torsion angles. Furthermore, the dipolarity/polarizability index (π∗) was the preferred solvent parameter for predicting the effects of solvents on λmax.}, number={3}, journal={DYES AND PIGMENTS}, author={He, Liang and El-Shafei, Ahmed and Freeman, Harold S. and Boyle, Paul}, year={2009}, month={Sep}, pages={299–306} } @article{foley_gunnoe_cundari_boyle_petersen_2008, title={Activation of sp(3) carbon-hydrogen bonds complex and subsequent metal-mediated formation}, volume={47}, number={4}, journal={Angewandte Chemie [International Edition in English]}, author={Foley, N. A. and Gunnoe, T. B. and Cundari, T. R. and Boyle, P. D. and Petersen, J. L.}, year={2008}, pages={726–730} } @article{munro-leighton_feng_zhang_alsop_gunnoe_boyle_petersen_2008, title={Preparation and reactivity of a monomeric octahedral platinum(IV) amido complex}, volume={47}, DOI={10.1021/icS00843b}, number={14}, journal={Inorganic Chemistry}, author={Munro-Leighton, C. and Feng, Y. and Zhang, J. and Alsop, N. M. and Gunnoe, T. B. and Boyle, P. D. and Petersen, J. L.}, year={2008}, pages={6124–6126} } @article{sumiyoshi_gunnoe_petersen_boyle_2008, title={Ruthenium(II) complexes possessing the eta(6)-p-cymene ligand}, volume={361}, ISSN={["1873-3255"]}, DOI={10.1016/j.ica.2007.10.012}, abstractNote={Complexes possessing a soft donor η6-arene and hard donor acetylacetonate ligand, [(η6-p-cymene)Ru(κ2-O,O-acac-μ-CH)]2[OTf]2 (1) (OTf = trifluoromethanesulfonate; acac = acetylacetonate) and [(η6-p-cymene)Ru(κ2-O,O-acac)(THF)][BAr4′](2) {Ar′ = 3,5-(CF3)–C6H3}, were prepared and fully characterized. The lability of the μ-CH linkage for complex 1 and the THF ligand of 2 allow access to the unsaturated cation [(η6-p-cymene)Ru(κ2-O,O-acac)]+. The reaction of [(η6-p-cymene)Ru(κ2-O,O-acac)(THF)][BAr4′](2) with KTp {Tp = hydridotris(pyrazolyl)borate} produces [TpRu(η6-p-cymene)][BAr4′](5). The azide complex [(η6-p-cymene)Ru(κ2-O,O-acac)(N3Ar)][BAr4′](6) forms upon reaction of [(η6-p-cymene)Ru(κ2-O,O-acac)(THF)][BAr4′](2)with N3Ar (Ar = p-tolyl), and reaction of [(η6-p-cymene)Ru(κ2-O,O-acac)(THF)][BAr4′](2) with CHCl3 at 100 °C yields the chloride-bridged binuclear complex [{(η6-p-cymene)Ru}2(μ-Cl)3][BAr4′](7). The details of solid-state structures of [(η6-p-cymene)Ru(κ2-O,O-acac-μ-CH)]2[OTf]2 (1), [TpRu(η6-p-cymene)][BAr4′](5) and [{(η6-p-cymene)Ru}2(μ-Cl)3][BAr4′](7) are disclosed.}, number={11}, journal={INORGANICA CHIMICA ACTA}, author={Sumiyoshi, Taisuke and Gunnoe, T. Brent and Petersen, Jeffrey L. and Boyle, Paul D.}, year={2008}, month={Jul}, pages={3254–3262} } @article{delp_munro-leighton_goj_ramirez_gunnoe_petersen_boyle_2007, title={Addition of s-h bonds across electron-deficient olefins catalyzed by well-defined copper(I) thiolate complexes}, volume={46}, ISSN={["1520-510X"]}, DOI={10.1021/ic070268s}, abstractNote={A series of monomeric (NHC)Cu(SR) (R = Ph or CH2Ph; NHC = N-heterocyclic carbene) complexes have been synthesized and fully characterized including single-crystal X-ray diffraction studies. These complexes catalyze the addition of S-H bonds across electron-deficient olefins to regioselectively produce "anti-Markovnikov" products.}, number={7}, journal={INORGANIC CHEMISTRY}, author={Delp, Samuel A. and Munro-Leighton, Colleen and Goj, Laurel A. and Ramirez, Magaly A. and Gunnoe, T. Brent and Petersen, Jeffrey L. and Boyle, Paul D.}, year={2007}, month={Apr}, pages={2365–2367} } @article{foley_lail_gunnoe_cundari_boyle_petersen_2007, title={Combined experimental and computational study of TpRu{P(pyr)(3)}(NCMe)Me (pyr = N-pyrrolyl): Inter- and intramolecular activation of C-H bonds and the impact of sterics on catalytic hydroarylation of olefins}, volume={26}, ISSN={["1520-6041"]}, DOI={10.1021/om700666y}, abstractNote={Complexes of the type TpRu{P(pyr)3}(L)R {L = PPh3 or NCMe; R = Cl, OTf (OTf = trifluoromethanesulfonate), Me, or Ph; Tp = hydridotris(pyrazolyl)borate; pyr = N-pyrrolyl} and TpRu{κ2-P,C-P(pyr)2(NC4H3)}NCMe have been synthesized and isolated. TpRu{P(pyr)3}(NCMe)Me initiates intermolecular C−H activation of benzene to form TpRu{P(pyr)3}(NCMe)Ph and, in the absence of benzene, intramolecular C−H activation of a pyrrolyl ring to form the cyclometalated species TpRu{κ2-P,C-P(pyr)2(NC4H3)}NCMe. TpRu{P(pyr)3}(NCMe)Ph catalyzes the hydrophenylation of ethylene in benzene to produce ethylbenzene in low yields. Experimental and computational analyses of the hydrophenylation of ethylene by TpRu{P(pyr)3}(NCMe)Ph suggest that inefficient catalysis is not due to difficulty in the C−H activation of benzene by the active catalyst species, but rather likely arises from the steric bulk of the tris-N-pyrrolyl phosphine ligand, which inhibits coordination of ethylene and thus thwarts C−C bond formation.}, number={23}, journal={ORGANOMETALLICS}, author={Foley, Nicholas A. and Lail, Marty and Gunnoe, T. Brent and Cundari, Thomas R. and Boyle, Paul D. and Petersen, Jeffrey L.}, year={2007}, month={Nov}, pages={5507–5516} } @article{taniguchi_ptaszek_mcdowell_boyle_lindsey_2007, title={Sparsely substituted chlorins as core constructs in chlorophyll analogue chemistry. Part 3: Spectral and structural properties}, volume={63}, ISSN={["0040-4020"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-33947694765&partnerID=MN8TOARS}, DOI={10.1016/j.tet.2007.02.040}, abstractNote={The availability of stable chlorins bearing few or no substituents has enabled a variety of fundamental studies. The studies described herein report absorption spectra of diverse chlorins, comparative NMR features of chlorins bearing 0-3 meso-aryl substituents, and X-ray structures of the fully unsubstituted chlorin and the oxochlorin.}, number={18}, journal={TETRAHEDRON}, publisher={Elsevier BV}, author={Taniguchi, Masahiko and Ptaszek, Marcin and McDowell, Brian E. and Boyle, Paul D. and Lindsey, Jonathan S.}, year={2007}, month={Apr}, pages={3850–3863} } @article{shultz_sloop_coote_beikmohammadi_kampf_boyle_2007, title={Substituent Effects on Exchange Coupling:  5-Aryl-Substituted Semiquinones and Their Complexes with MnIIand CuII}, volume={46}, ISSN={0020-1669 1520-510X}, url={http://dx.doi.org/10.1021/ic061807g}, DOI={10.1021/ic061807g}, abstractNote={A series of functionalized radical anion semiquinone (SQ-Ar) ligands and their MnII- and CuIIhydro-tris(3-cumenyl-5-methylpyrazolyl)borate (TpCum,MeMII) complexes were prepared and characterized. The semiquinone ligands have substituted phenyl rings (Ar = -C6H5NO2, -C6H5OMe, -C6H5-tert-Bu, etc.) attached to the SQ 5-position. Despite the "remoteness" of the phenyl ring substituents, the MII-SQ exchange parameters, J, were found to vary nearly 3-fold. Attempts to quantify the substituent effects on J are complicated by the fact that not all complexes could be structurally characterized. As such, substituent effects and phenyl-ring torsion angles could conspire to produce the observed variation in J values. Although there is no clear trend in the J values as a function of SQ substituent for the MnII complexes, for the CuII complexes, electron-withdrawing substituents on the phenyl ring have greater ferromagnetic J values than the CuII complexes of SQ ligands with electron-donating substituents. This trend suggests a FM contribution from MLCT excited states in the copper complexes.}, number={1}, journal={Inorganic Chemistry}, publisher={American Chemical Society (ACS)}, author={Shultz, David A. and Sloop, Joseph C. and Coote, Tashni-Ann and Beikmohammadi, Mithra and Kampf, Jeff and Boyle, Paul D.}, year={2007}, month={Jan}, pages={273–277} } @article{ptaszek_yao_savithri_boyle_lindsey_2007, title={Synthesis and structural properties of porphyrin analogues of bacteriochlorophyll c}, volume={63}, ISSN={["0040-4020"]}, DOI={10.1016/j.tet.2007.10.023}, abstractNote={The self-assembling photosynthetic pigment bacteriochlorophyll c contains α-hydroxyethyl and keto groups on opposite sides of the macrocycle. A porphyrin has been synthesized that contains a 3-hydroxymethyl group and a 15-ethoxycarbonyl group (ZnP2-OH). X-ray analysis of ZnP2-OH and the related porphyrin containing 5-hydroxymethyl and 15-ethoxycarbonyl groups (ZnP1-OH) in each case revealed infinite coordination polymers wherein the zinc porphyrins are bound by Zn–O coordination and are cofacially offset in a staircase architecture.}, number={51}, journal={TETRAHEDRON}, author={Ptaszek, Marcin and Yao, Zhen and Savithri, Dhanalekshmi and Boyle, Paul D. and Lindsey, Jonathan S.}, year={2007}, month={Dec}, pages={12629–12638} } @article{goj_blue_delp_gunnoe_cundari_pierpont_petersen_boyle_2006, title={Chemistry surrounding monomeric copper(I) methyl, phenyl, anilido, ethoxide, and phenoxide complexes supported by N-heterocyclic carbene ligands: Reactivity consistent with both early and late transition metal systems}, volume={45}, DOI={10.1021/ic0611995}, abstractNote={Monomeric copper(I) alkyl complexes that possess the N-heterocyclic carbene (NHC) ligands IPr, SIPr, and IMes [IPr = 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene, SIPr = 1,3-bis(2,6-diisopropylphenyl)imidazolin-2-ylidene, IMes = 1,3-bis(2,4,6-trimethylphenyl)imidazol-2-ylidene] react with amines or alcohols to release alkane and form the corresponding monomeric copper(I) amido, alkoxide, or aryloxide complexes. Thermal decomposition reactions of (NHC)Cu(I) methyl complexes at temperatures between 100 and 130 degrees C produce methane, ethane, and ethylene. The reactions of (NHC)Cu(NHPh) complexes with bromoethane reveal increasing nucleophilic reactivity at the anilido ligand in the order (SIPr)Cu(NHPh) < (IPr)Cu(NHPh) < (IMes)Cu(NHPh) < (dtbpe)Cu(NHPh) [dtbpe = 1,2-bis(di-tert-butylphosphino)ethane]. DFT calculations suggest that the HOMO for the series of Cu anilido complexes is localized primarily on the amido nitrogen with some ppi(anilido)-dpi(Cu) pi-character. [(IPr)Cu(mu-H)]2 and (IPr)Cu(Ph) react with aniline to quantitatively produce (IPr)Cu(NHPh)/dihydrogen and (IPr)Cu(NHPh)/benzene, respectively. Analysis of the DFT calculations reveals that the conversion of [(IPr)Cu(mu-H)]2 and aniline to (IPr)Cu(NHPh) and dihydrogen is favorable with DeltaH approximately -7 kcal/mol and DeltaG approximately -9 kcal/mol.}, number={22}, journal={Inorganic Chemistry}, author={Goj, L. A. and Blue, E. D. and Delp, S. A. and Gunnoe, T. B. and Cundari, T. R. and Pierpont, A. W. and Petersen, J. L. and Boyle, P. D.}, year={2006}, pages={9032–9045} } @article{blue_gunnoe_petersen_boyle_2006, title={Protonation of N-heterocyclic carbene ligand coordinated to copper(I): Coordination mode of imidazolium cation as a function of counterion as determined by solid-state structures}, volume={691}, ISSN={["1872-8561"]}, DOI={10.1016/j.jorganchem.2006.09.051}, abstractNote={Reactions of (IPr)Cu(X) (X = Cl or trifluoromethanesulfonate, IPr = 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene) complexes with the strong acids HOTf or HCl result in protonation of the C2 carbon of the IPr ligand to form imidazolium cations. Coordination of the imidazolium to the resulting CuI system depends upon the identity of the two counterions (chloride or triflate). The copper complexes [(IPrH)Cu(OTf)(μ-OTf)]2 and [IPrH][CuCl2] as well as the imidazolium salt [IPrH][OTf] have been characterized by NMR spectroscopy and single crystal X-ray diffraction studies.}, number={26}, journal={JOURNAL OF ORGANOMETALLIC CHEMISTRY}, author={Blue, Elizabeth D. and Gunnoe, T. Brent and Petersen, Jeffrey L. and Boyle, Paul D.}, year={2006}, month={Dec}, pages={5988–5993} } @article{lin_yan_boyle_maggard_2006, title={Synthesis and Properties of Pyrazine-Pillared Ag3Mo2O4F7 and AgReO4 Layered Phases}, volume={179}, ISSN={["1095-726X"]}, DOI={10.1016/j.jssc.2005.10.037}, abstractNote={The new pyrazine-pillared solids, AgReO4(C4H4N2) (I) and Ag3Mo2O4F7(C4H4N2)3 (C4H4N2=pyrazine, pyz) (II), were synthesized by hydrothermal methods at 150 °C and characterized using single crystal X-ray diffraction (I—P21/c, No. 14, Z=4, a=7.2238(6) Å, b=7.4940(7) Å, c=15.451(1) Å, β=92.296(4)°; II—P2/n, No. 13, Z=2, a=7.6465(9) Å, b=7.1888(5) Å, c=19.142(2) Å, β=100.284(8)°), thermogravimetric analysis, UV-Vis diffuse reflectance, and photoluminescence measurements. Individual Ag(pyz) chains in I are bonded to three perrhenate ReO4– tetrahedra per layer, while each layer in II contains sets of three edge-shared Ag(pyz) chains (π–π stacked) that are edge-shared to four Mo2O4F73– dimers. A relatively small interlayer spacing results from the short length of the pyrazine pillars, and which can be removed at just slightly above their preparation temperature, at >150–175 °C, to produce crystalline AgReO4 for I, and Ag2MoO4 and an unidentified product for II. Both pillared solids exhibit strong orange-yellow photoemission, at 575 nm for I and 560 nm for II, arising from electronic excitations across (charge transfer) band gaps of 2.91 and 2.76 eV in each, respectively. Their structures and properties are analyzed with respect to parent ‘organic free’ silver perrhenate and molybdate solids which manifest similar photoemissions, as well as to the calculated electronic band structures.}, number={1}, journal={Journal of Solid State Chemistry}, author={Lin, H. and Yan, B. and Boyle, P.D. and Maggard, P}, year={2006}, pages={217–225} } @article{tang_boyle_novak_2005, title={Chiroptical switching polyguanidine synthesized by helix-sense-selective polymerization using [(R)-3,3 '-dibromo-2,2 '-binaphthoxy](di-tert-butoxy)titanium(IV) catalyst}, volume={127}, ISSN={["1520-5126"]}, DOI={10.1021/ja0453533}, abstractNote={A series of chiral binaphthyl titanium alkoxide complexes were synthesized. Among them, chiral titanium complex [(R)-3,3'-dibromo-2,2'-binaphthoxy](di-tert-butoxy)titanium(IV) (R-3) exists as a crystallographic C2 dimer in the solid state but a monomer in solution at room temperature. Application of R-3 in the helix-sense-selective polymerization of achiral carbodiimide, N-(1-anthryl)-N'-octadecylcarbodiimide (1), yielded a well-defined regioregular, stereoregular poly[N-(1-anthryl)-N'-octadecylguanidine] (poly-1b) with a relatively narrow polymer dispersity index of 2.7. Full racemization of poly-1b at +80 degrees C in toluene requires more than 100 h. Interestingly, poly-1b was found to undergo fast reversible chiroptical switching at +38.5 degrees C in toluene. Furthermore, at room temperature, poly-1b shows a positively signed Cotton effect in toluene, but negative ones in THF and chloroform, respectively. The chiroptical switching takes place around the toluene content of 90% (vol) in the mixed toluene/THF solvents. This is the first example of chiroptical switching phenomenon occurring in a helical polymer possessing no chiral moieties in the polymer chains. We believe this reversible chiroptical switching phenomenon occurs by reorientation of anthracene rings relative to the chain director.}, number={7}, journal={JOURNAL OF THE AMERICAN CHEMICAL SOCIETY}, author={Tang, HZ and Boyle, PD and Novak, BM}, year={2005}, month={Feb}, pages={2136–2142} } @article{zhang_barakat_cundari_gunnoe_boyle_petersen_day_2005, title={Synthesis of the five-coordinate ruthenium(II) Complexes [(PCP)Ru(CO)(L)][BAr'(4)] {PCP=2,6-((CH2PBu2)-Bu-t)(2)C6H3, BAr'(4) 3,5-(CF3)(2)C6H3, L = eta(1)-ClCH2Cl, eta-N-2, or mu-Cl-Ru(PCP)(CO)}: Reactions with phenyldiazomethane and phenylacetylene}, volume={44}, DOI={10.1012/ic051074k}, number={23}, journal={Inorganic Chemistry}, author={Zhang, J. B. and Barakat, K. A. and Cundari, T. R. and Gunnoe, T. B. and Boyle, P. D. and Petersen, J. L. and Day, C. S.}, year={2005}, pages={8379–8390} } @article{nodono_novak_boyle_2004, title={Ethylene polymerization catalyzed by neutral nickel(II) complex with ON-chelating ligand}, volume={36}, ISSN={["0032-3896"]}, DOI={10.1295/polymj.36.140}, abstractNote={We have achieved the synthesis of neutral nickel catalyst with the modified O^N-chelating ligand for the ethylene polymerization. The activity of the catalyst, and the molecular weight and the branching structure of the polymer obtained strongly depend on the ligand structure as well as the presence of Ni(COD)2 which is used as an activator. The crystal structure of the catalyst was determined and the long nickel–phosphine bond length seems to play an important role for increasing the polymerization activity.}, number={2}, journal={POLYMER JOURNAL}, author={Nodono, M and Novak, BM and Boyle, PT}, year={2004}, pages={140–145} } @article{el-shafei_hinks_boyle_freeman_2004, title={Non-mutagenic organic pigment intermediates. II. Isomorphous 2,2 '-dichloro-5,5 '-dipropoxybenzidine and 2,2 '-dimethyl-5,5 '-dipropoxybenzidine}, volume={60}, ISSN={["2053-2296"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-4444237632&partnerID=MN8TOARS}, DOI={10.1107/S0108270104014763}, abstractNote={The title compounds, C 18 H 22 Cl 2 N 2 O 2 and C 20 H 28 N 2 O 2 , respectively, are isomorphous. The molecules lie at general positions in the unit cell. In each structure, chemically equivalent but crystallographically inequivalent amine N atoms exhibit different degrees of pyramidalization. The structures exhibit weak N-H...N hydrogen bonding, which is influenced by the differences in hybridization around the amine N atoms. The torsion angles across the biphenyl linkage for the two compounds are 67.2 (2) and 68.3 (3)°.}, number={8}, journal={ACTA CRYSTALLOGRAPHICA SECTION C-STRUCTURAL CHEMISTRY}, author={El-Shafei, A and Hinks, D and Boyle, PD and Freeman, HS}, year={2004}, month={Aug}, pages={O569–O571} } @article{zhang_gunnoe_boyle_2004, title={Ruthenium(II) anilido complex containing a bisphosphine pincer ligand: Reversible formation of amidinate ligands via intramolecular C-N bond formation}, volume={23}, ISSN={["1520-6041"]}, DOI={10.1021/om0497543}, abstractNote={The reaction of the octahedral anilido complex (PCP)Ru(CO)(NHPh)(PMe3) with acetonitrile produces the amidinate complex (PCP)Ru(CO){N(H)C(Me)N(Ph)}. Mechanistic studies indicate that the reaction proceeds through coordination of the nitrile to the Ru(II) metal center, followed by intramolecular nucleophilic addition of the amido ligand.}, number={13}, journal={ORGANOMETALLICS}, author={Zhang, J and Gunnoe, TB and Boyle, PD}, year={2004}, month={Jun}, pages={3094–3097} } @article{martin_hess_boyle_2004, title={Synthesis of [NH4]MnCl2(OAc) and [NH4](2)MnCl4(H2O)(2) by solvothermal dehydration and structure/property correlations in a one-dimensional antiferromagnet}, volume={43}, ISSN={["0020-1669"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-2442565694&partnerID=MN8TOARS}, DOI={10.1021/ic049881r}, abstractNote={The utility of the solvothermal dehydration strategy whereby superheated acetonitrile reacts with water of hydration to form ammonium acetate is demonstrated in the synthesis of [NH(4)]MnCl(2)(OAc), I, and [NH(4)](2)MnCl(4)(H(2)O)(2), II, from MnCl(2).4H(2)O. The structure of I is shown to crystallize in the monoclinic space group C2/c (No. 15) with a = 15.191(6) A, b = 7.044(2) A, c = 13.603(6) A, beta = 107.31 degrees, V = 1389.7(9) cm(-)(1), and Z = 8. The structure of II crystallizes in the space group I4/mmm (No. 139) with a = 7.5250(5) A, b = 8.276(2) A, V = 468.6(1) cm(-)(1), and Z = 2. Both structures exhibit extensive hydrogen bonding that controls both local Mn-Cl bonding and the interchain organization. I is shown to be a one-dimensional Heisenberg antiferromagnet with an intrachain exchange constant J/k = -2.39 K. This structure exhibits exchange coupling intermediate between the well-studied triply and doubly chloride-bridged one-dimensional manganese Heisenberg antiferromagnets. The structure/property correlation demonstrates a linear dependence of the exchange constant on the Mn-Cl-Mn bond angle, alpha, for alpha < 94 degrees.}, number={10}, journal={INORGANIC CHEMISTRY}, author={Martin, JD and Hess, RF and Boyle, PD}, year={2004}, month={May}, pages={3242–3247} } @article{wolak_knutson_martin_boyle_sargent_white_2003, title={Dynamic disorder and conformer exchange in the crystalline monomer of polycarbonate}, volume={107}, ISSN={["1520-5207"]}, DOI={10.1021/jp036527q}, abstractNote={Direct measurement of chemical exchange events in the crystalline polycarbonate monomer 4,4‘- isopropylidenediphenol (bisphenol A, BPA) via 2D 13C solid-state NMR reveals slow, large-amplitude aromatic ring reorientations. X-ray diffraction, however, indicates a static crystalline structure. Experiments with multiple exchange times show that ring flips occur in all of the three unique conformers found in the crystalline unit cell of this compound, but in specific cases, two of the three unique molecules actually switch conformations. Kinetic analysis of the exchange data indicates that the average rate constant kex = 0.01 s-1 for ring flips and conformer interchange at room temperature. Differential scanning calorimetry and variable-temperature powder diffraction studies indicate a systematic volume expansion that accompanies this motion but no first-order phase transition. All room-temperature exchange events may be quenched at 213 K, at least on the time scale (up to several seconds) probed in this work...}, number={48}, journal={JOURNAL OF PHYSICAL CHEMISTRY B}, author={Wolak, JE and Knutson, J and Martin, JD and Boyle, P and Sargent, AL and White, JL}, year={2003}, month={Dec}, pages={13293–13299} } @article{yu_muthukumaran_sazanovich_kirmaier_hindin_diers_boyle_bocian_holten_lindsey_2003, title={Excited-state energy-transfer dynamics in self-assembled triads composed of two porphyrins and an intervening bis(dipyrrinato)metal complex}, volume={42}, ISSN={["1520-510X"]}, DOI={10.1021/ic034559m}, abstractNote={The synthesis and characterization of various triads composed of a linear array of two zinc porphyrins joined via an intervening bis(dipyrrinato)metal(II) complex are reported. The preparation exploits the facile complexation of dipyrrins with divalent metal ions to give bis(dipyrrinato)metal(II) complexes [abbreviated (dp)(2)M]. Copper(II) and palladium(II) chelates of dipyrrins (available by oxidation of dipyrromethanes) were prepared in 50-80% yield. A one-flask synthesis of bis(dipyrrinato)zinc(II) complexes was developed by oxidation of a dipyrromethane with DDQ or p-chloranil in the presence of Zn(OAc)(2).2H(2)O in THF ( approximately 80% yield). Three routes were developed for preparing porphyrin-dipyrrins: (1). Suzuki coupling of a boronate-substituted zinc porphyrin (ZnP) and bis[5-(4-iodophenyl)dipyrrinato]Pd(II) to give the (ZnP-dp)(2)Pd triad (50% yield), followed by selective demetalation of the (dp)(2)Pd unit by treatment with 1,4-dithiothreitol under neutral conditions (71% yield); (2). oxidation of a porphyrin-dipyrromethane with p-chloranil in the presence of Zn(OAc)(2).2H(2)O followed by chromatography on silica gel (71% yield); and (3). condensation of a dipyrrin-dipyrromethane and a dipyrromethane-dicarbinol under InCl(3) catalysis followed by oxidation with DDQ (10-16% yield). Four triads of form (ZnP-dp)(2)Zn were prepared in 83-97% yield by treatment of a porphyrin-dipyrrin with Zn(OAc)(2).2H(2)O at room temperature. Free base dipyrrins typically absorb at 430-440 nm, while the bis(dipyrrinato)metal complexes absorb at 460-490 nm. The fluorescence spectra/yields and excited-state lifetimes of the (ZnP-dp)(2)Zn triad in toluene show (1). efficient energy transfer from the bis(dipyrrinato)zinc(II) chromophore to the zinc porphyrins (98.5% yield), and (2). little or no quenching of the resulting excited zinc porphyrin relative to the isolated chromophore. Taken together, these results indicate that bis(dipyrrinato)zinc(II) complexes can serve as self-assembling linkers that further function as secondary light-collection elements in porphyrin-based light-harvesting arrays.}, number={21}, journal={INORGANIC CHEMISTRY}, author={Yu, LH and Muthukumaran, K and Sazanovich, IV and Kirmaier, C and Hindin, E and Diers, JR and Boyle, PD and Bocian, DF and Holten, D and Lindsey, JS}, year={2003}, month={Oct}, pages={6629–6647} } @article{shultz_fico_lee_kampf_kirschbaum_pinkerton_boyle_2003, title={Mechanisms of Exchange Modulation in Trimethylenemethane-type Biradicals:  The Roles of Conformation and Spin Density}, volume={125}, ISSN={["0002-7863"]}, DOI={10.1021/ja037787o}, abstractNote={The molecular structures and magnetic properties of six dinitroxide biradicals are described. Five of the dinitroxides are trimethylenemethane-type (TMM-type) biradicals; that is, the intramolecular exchange parameter, J, is modulated by a carbon-carbon double bond. However, the efficacy of the carbon-carbon double bond as an exchange coupler is determined by the molecular conformation. Our results show that the exchange parameters correlate with phenyl-ring torsion angles (phi) via a simple Karplus-Conroy-type relation: J = 44 cos(2) phi - 17. Comparison of these results to those obtained for our isostructural series of bis(semiquinone) biradicals shows that both the magnitude of J and the resistance of ferromagnetic J to bond torsions is proportional to the spin density adjacent to the exchange coupler.}, number={50}, journal={Journal of the American Chemical Society}, author={Shultz, D.A. and Fico, R.M., Jr. and Lee, H. and Kampf, J.W. and Kirschbaum, K. and Pinkerton, A.A. and Boyle, P.D.}, year={2003}, pages={15426–15432} } @article{el-shafei_boyle_hinks_freeman_2003, title={Non-mutagenic organic pigment intermediates. I. 3,3 '-Dipropoxybenzidine}, volume={59}, ISSN={["0108-2701"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0037295760&partnerID=MN8TOARS}, DOI={10.1107/S0108270102021303}, abstractNote={Crystals of the title compound, C(18)H(24)N(2)O(2), were grown from ethanol by slow evaporation and the structure has been determined. The molecule resides on a crystallographic inversion center and the biphenyl moiety is essentially planar. The structure forms an infinite two-dimensional array of N-H.pi(arene) interactions parallel to the (101) direction.}, number={2}, journal={ACTA CRYSTALLOGRAPHICA SECTION C-CRYSTAL STRUCTURE COMMUNICATIONS}, author={El-Shafei, A and Boyle, PD and Hinks, D and Freeman, HS}, year={2003}, month={Feb}, pages={O71–O73} } @article{conner_jayaprakash_wells_manzer_gunnoe_boyle_2003, title={Octahedral Ru(II) amido complexes TpRu(L)(L ')(NHR) (Tp = hydridotris(pyrazolyl)borate; L = L ' = P(OMe)(3) or PMe3 or L = CO and L ' = PPh3; R = H, Ph, or Bu-t): Synthesis, characterization, and reactions with weakly acidic C-H bonds}, volume={42}, ISSN={["1520-510X"]}, DOI={10.1021/ic034410a}, abstractNote={The octahedral Ru(II) amine complexes [TpRu(L)(L')(NH(2)R)][OTf] (L = L' = PMe(3), P(OMe)(3) or L = CO and L' = PPh(3); R = H or (t)Bu) have been synthesized and characterized. Deprotonation of the amine complexes [TpRu(L)(L')(NH(3))][OTf] or [TpRu(PMe(3))(2)(NH(2)(t)Bu)][OTf] yields the Ru(II) amido complexes TpRu(L)(L')(NH(2)) and TpRu(PMe(3))(2)(NH(t)Bu). Reactions of the parent amido complexes or TpRu(PMe(3))(2)(NH(t)Bu) with phenylacetylene at room temperature result in immediate deprotonation to form ruthenium-amine/phenylacetylide ion pairs, and heating a benzene solution of the [TpRu(PMe(3))(2)(NH(2)(t)Bu)][PhC(2)] ion pair results in the formation of the Ru(II) phenylacetylide complex TpRu(PMe(3))(2)(C[triple bond]CPh) in >90% yield. The observation that [TpRu(PMe(3))(2)(NH(2)(t)Bu)][PhC(2)] converts to the Ru(II) acetylide with good yield while heating the ion pairs [TpRu(L)(L')(NH(3))][PhC(2)] yields multiple products is attributed to reluctant dissociation of ammonia compared with the (t)butylamine ligand (i.e., different rates for acetylide/amine exchange). These results are consistent with ligand exchange reactions of Ru(II) amine complexes [TpRu(PMe(3))(2)(NH(2)R)][OTf] (R = H or (t)Bu) with acetonitrile. The previously reported phenyl amido complexes TpRuL(2)(NHPh) [L = PMe(3) or P(OMe)(3)] react with 10 equiv of phenylacetylene at elevated temperature to produce Ru(II) acetylide complexes TpRuL(2)(C[triple bond]CPh) in quantitative yields. Kinetic studies indicate that the reaction of TpRu(PMe(3))(2)(NHPh) with phenylacetylene occurs via a pathway that involves TpRu(PMe(3))(2)(OTf) or [TpRu(PMe(3))(2)(NH(2)Ph)][OTf] as catalyst. Reactions of 1,4-cyclohexadiene with the Ru(II) amido complexes TpRu(L)(L')(NH(2)) (L = L' = PMe(3) or L = CO and L' = PPh(3)) or TpRu(PMe(3))(2)(NH(t)Bu) at elevated temperatures result in the formation of benzene and Ru hydride complexes. TpRu(PMe(3))(2)(H), [Tp(PMe(3))(2)Ru[double bond]C[double bond]C(H)Ph][OTf], [Tp(PMe(3))(2)Ru=C(CH(2)Ph)[N(H)Ph]][OTf], and [TpRu(PMe(3))(3)][OTf] have been independently prepared and characterized. Results from solid-state X-ray diffraction studies of the complexes [TpRu(CO)(PPh(3))(NH(3))][OTf], [TpRu(PMe(3))(2)(NH(3))][OTf], and TpRu(CO)(PPh(3))(C[triple bond]CPh) are reported.}, number={15}, journal={INORGANIC CHEMISTRY}, author={Conner, D and Jayaprakash, KN and Wells, MB and Manzer, S and Gunnoe, TB and Boyle, PD}, year={2003}, month={Jul}, pages={4759–4772} } @article{arrowood_lail_gunnoe_boyle_2003, title={Radical polymerization of styrene and methyl methacrylate with Ruthenium(II) complexes}, volume={22}, ISSN={["0276-7333"]}, DOI={10.1021/om020871c}, abstractNote={Methylation of [TpRu(CO)2(THF)][PF6] yields the Ru(II) complex TpRu(CO)2(Me), and reaction of TpRu(CO)2(Me) with Me3NO in refluxing acetonitrile yields TpRu(CO)(CH3)(NCMe) (Tp = hydridotris(pyrazolyl)borate). Reactions of TpRu(CO)(CH3)(NCCH3) with CD3CN indicate that the bound acetonitrile ligand undergoes exchange to yield TpRu(CO)(CH3)(NCCD3), and the rates of the exchange reactions are independent of CD3CN concentration. At elevated temperatures catalytic quantities of TpRu(CO)(CH3)(NCMe) in the presence of styrene result in the production of polystyrene. The dependence of polystyrene molecular weight on benzene/cumene molar ratios indicates a radical polymerization mechanism. Styrene polymerization also occurs in the presence of carbon tetrachloride or methyl dichloroacetate. In addition, the polymerization of methyl methacrylate in the presence of TpRu(CO)(CH3)(NCMe) with carbon tetrachloride or methyl dichloroacetate has been observed at 90 °C.}, number={23}, journal={ORGANOMETALLICS}, author={Arrowood, BN and Lail, M and Gunnoe, TB and Boyle, PD}, year={2003}, month={Nov}, pages={4692–4698} } @article{maggard_boyle_2003, title={Synthesis and properties of V6O16CU(C4H4N2)(2)(.)(H2O)(0.22(1)): Charge density matching of a metal-segregated layer structure}, volume={42}, ISSN={["0020-1669"]}, DOI={10.1021/ic0342649}, abstractNote={A metal-segregated layered compound, containing square nets of Cu(pyz)(2)(2+) and buckled V(6)O(16)(2)(-) layers, has been synthesized using hydrothermal techniques to have the composition V(6)O(16)Cu(C(4)H(4)N(2))(2) x (H(2)O)(0.22(1)) (C(4)H(4)N(2) = pyrazine, pyz). The Cu(II) square nets are nearly regular and undergo an antiferromagnetic transition at 8 K. In contrast to the plethora of recently synthesized metal-oxide clusters, chains, and networks in the VO(x)/M/L (M = late transition element; L = organonitrogen ligand) system, this compound is a relatively rare example that contains two different metals distributed into distinct layers. An application of charge density matching to form layered structures is postulated.}, number={14}, journal={INORGANIC CHEMISTRY}, author={Maggard, PA and Boyle, PD}, year={2003}, month={Jul}, pages={4250–4252} } @article{shultz_fico_bodnar_kumar_vostrikova_kampf_boyle_2003, title={Trends in exchange coupling for trimethylenemethane-type bis(semiquinone) biradicals and correlation of magnetic exchange with mixed valency for cross-conjugated systems}, volume={125}, ISSN={["0002-7863"]}, DOI={10.1021/ja0367849}, abstractNote={A magnetostructural correlation (conformational electron spin exchange modulation) within an isostructural series of biradical complexes is presented. X-ray crystal structures, variable-temperature electron paramagnetic resonance spectroscopy, zero-field splitting parameters, and variable-temperature magnetic susceptibility measurements were used to evaluate molecular conformation and electron spin exchange coupling in this series of molecules. Our combined results indicate that the ferromagnetic portion of the exchange couplings occurs via the cross-conjugated pi-systems, while the antiferromagnetic portion occurs through space and is equivalent to incipient bond formation. Thus, molecular conformation controls the relative amounts of ferro- and antiferromagnetic contributions to exchange coupling. In fact, the exchange parameter correlates with average semiquinone ring torsion angles via a Karplus-Conroy-type relation. Because of the natural connection between electron spin exchange coupling and electronic coupling related to electron transfer, we also correlate the exchange parameters in the biradical complexes to mixed valency in the corresponding quinone-semiquinone radical anions. Our results suggest that delocalization in the cross-conjugated, mixed-valent radical anions is proportional to the ferromagnetic contribution to the exchange coupling in the biradical oxidation states.}, number={38}, journal={JOURNAL OF THE AMERICAN CHEMICAL SOCIETY}, author={Shultz, DA and Fico, RM and Bodnar, SH and Kumar, RK and Vostrikova, KE and Kampf, JW and Boyle, PD}, year={2003}, month={Sep}, pages={11761–11771} } @article{yu_ballard_boyle_wang_2002, title={An inexpensive carbohydrate derivative used as a chiral auxiliary in the synthesis of alpha-hydroxy carboxylic acids}, volume={58}, ISSN={["0040-4020"]}, DOI={10.1016/S0040-4020(02)00871-2}, abstractNote={Protected α-hydroxy carboxylic acids were synthesized in moderate yield and high diastereoselectivity by alkylation of glycolate (α-hydroxy acetate) enolates using a d-fructose-derived chiral auxiliary. The new chiral center was assigned the R configuration based on comparisons of optical rotations and on one crystal structure analysis. This alkylation methodology is compatible with several hydroxyl protecting groups. The free hydroxy acids were obtained upon removal of the protecting group from the hydroxyl functionality followed by saponification.}, number={38}, journal={TETRAHEDRON}, author={Yu, HW and Ballard, CE and Boyle, PD and Wang, BH}, year={2002}, month={Sep}, pages={7663–7679} } @article{conner_jayaprakash_gunnoe_boyle_2002, title={Influence of filled d pi-manifold and L/L ' ligands on the structure, basicity, and bond rotations of the octahedral and d(6) amido complexes TpRu(L)(L ')(NHPh) (Tp = hydridotris(pyrazolyl)borate; L= L ' = PMe3 or P(OMe)(3), or L = CO and L ' = PPh3): Solid-state structures of [TpRu(PMe3)(2)(NH2Ph)][OTf], [TpRu{P(OMe)(3)}(2)(NH2Ph)][OTf], and TpRu{P(OMe)(3)}(2)(NHPh)}, volume={41}, ISSN={["1520-510X"]}, DOI={10.1021/ic020163j}, abstractNote={It has been suggested that the reactivity of π-donating ligands bound to late-transition-metal complexes is heightened due to high d-electron counts. Herein, the synthesis and characterization of the Ru(II) amine and Ru(II) amido complexes [TpRuL2(NH2Ph)][OTf] (OTf = trifluoromethanesulfonate) and TpRuL2(NHPh) (L = PMe3 or P(OMe)3) are presented, including solid-state X-ray diffraction studies of [TpRu(PMe3)2(NH2Ph)][OTf], [TpRu{P(OMe)3}2(NH2Ph)][OTf], and TpRu{P(OMe)3}2(NHPh). The pKa's of the Ru(II) amine complexes and the previously reported [TpRu(CO)(PPh3)(NH2Ph)]+ have been estimated to be comparable to that of malononitrile in methylene chloride. In addition, the impact of the filled dπ-manifold (i.e., Ru(II) and d6 octahedral systems) on barriers to rotation of the Ru−NHPh moieties has been studied. For TpRu(PMe3)2(NHPh) and TpRu{P(OMe)3}2(NHPh), evidence for hindered rotation about the amido nitrogen and phenyl ipso carbon has been observed, and the relative N−C and Ru−N bond rotational barriers for the series of three amido complexes are discussed in terms of the π-conflict.}, number={11}, journal={INORGANIC CHEMISTRY}, author={Conner, D and Jayaprakash, KN and Gunnoe, TB and Boyle, PD}, year={2002}, month={Jun}, pages={3042–3049} } @article{conner_jayaprakash_gunnoe_boyle_2002, title={Ruthenium(II) anilido complexes TpRuL(2)(NHPh): Oxidative 4,4 '-aryl coupling reactions (Tp = hydridotris(pyrazolylborate); L = PMe3, P(OMe)(3), or CO)}, volume={21}, ISSN={["1520-6041"]}, DOI={10.1021/om0206406}, abstractNote={Reactions of the Ru(II) amido complexes TpRuL2(NHPh) (L = CO, PMe3, or P(OMe)3) with AgOTf (OTf = trifluoromethanesulfonate) yield the binuclear complexes [TpRuL2NH(C6H4−)]2[OTf]2 along with the Ru(II) amine complexes [TpRuL2(NH2Ph)][OTf] in an approximate 1:1 molar ratio. In these reactions, the two ruthenium fragments are coupled via C−H bond cleavage and C−C bond formation at the para position of anilido ligands. A resonance structure corresponding to Ru(II) metal centers linked by a diimine ligand contributes significantly to the bonding. Evidence for such a contribution comes from the diamagnetic nature of the binuclear complexes and a solid-state X-ray crystallographic study of [TpRu{P(OMe)3}2NH(C6H4−)]2[OTf]2. It is proposed that the coupled products are formed via initial single-electron oxidation followed by C−C bond formation. Variable-temperature NMR spectra of the aryl-coupled complexes are consistent with two geometrical isomers around the rigid HN−C6H4−C6H4NH bridges.}, number={24}, journal={ORGANOMETALLICS}, author={Conner, D and Jayaprakash, KN and Gunnoe, TB and Boyle, PD}, year={2002}, month={Nov}, pages={5265–5271} } @article{tian_boyle_novak_2002, title={Synthesis and crystal structure of a dinuclear palladium complex containing C,O-bridging ester-enolato moieties}, volume={21}, ISSN={["1520-6041"]}, DOI={10.1021/om010968j}, abstractNote={Palladium ester enolates have been prepared and characterized using IR, NMR, and single-crystal X-ray diffraction techniques. The X-ray diffraction established a dimeric crystal structure for complex {[N∧N]PdCH2C(O)OCH3}2 (N∧N = 1-[1-(5-methylpyrrole-2-yl)ethylidne]amino-2,6-diisopropylbenzene), 2, which contains C,O-bridging enolato groups. Upon reaction with donor molecules (acetonitrile, phosphines), the dimeric 2 cleaves to form monometallic C-bound enolate complexes, 3‘ and 4.}, number={7}, journal={ORGANOMETALLICS}, author={Tian, GL and Boyle, PD and Novak, BM}, year={2002}, month={Apr}, pages={1462–1465} } @article{shultz_fico_boyle_kampf_2001, title={Observation of a hysteretic phase transition in a crystalline dinitroxide biradical that leads to magnetic bistability}, volume={123}, ISSN={["0002-7863"]}, DOI={10.1021/ja011341v}, abstractNote={ADVERTISEMENT RETURN TO ISSUEPREVCommunicationNEXTObservation of a Hysteretic Phase Transition in a Crystalline Dinitroxide Biradical That Leads to Magnetic BistabilityDavid A. Shultz, Rosario M. Fico,, Paul D. Boyle, and Jeff W. KampfView Author Information Department of Chemistry North Carolina State University Raleigh, North Carolina 27695-8204 Department of Chemistry University of Michigan Ann Arbor, Michigan 48109-1055 Cite this: J. Am. Chem. Soc. 2001, 123, 42, 10403–10404Publication Date (Web):September 27, 2001Publication History Received1 June 2001Published online27 September 2001Published inissue 1 October 2001https://doi.org/10.1021/ja011341vCopyright © 2001 American Chemical SocietyRIGHTS & PERMISSIONSArticle Views378Altmetric-Citations28LEARN ABOUT THESE METRICSArticle Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated. Share Add toView InAdd Full Text with ReferenceAdd Description ExportRISCitationCitation and abstractCitation and referencesMore Options Share onFacebookTwitterWechatLinked InReddit Read OnlinePDF (51 KB) Get e-AlertscloseSupporting Info (2)»Supporting Information Supporting Information SUBJECTS:Bistability,Exchange coupling,Hysteresis,Molecules,Phenyls Get e-Alerts}, number={42}, journal={JOURNAL OF THE AMERICAN CHEMICAL SOCIETY}, author={Shultz, DA and Fico, RM and Boyle, PD and Kampf, JW}, year={2001}, month={Oct}, pages={10403–10404} } @article{jayaprakash_gunnoe_boyle_2001, title={Preparation of the octahedral d(6) amido complex TpRu(CO)(PPh3)(NHPh) (Tp = hydridotris(pyrazolyl)borate): Solid-state structural characterization and reactivity}, volume={40}, ISSN={["0020-1669"]}, DOI={10.1021/ic010785p}, abstractNote={The reaction of TpRu(CO)(PPh(3))(OTf) (2) with LiNHPh affords the amido complex TpRu(CO)(PPh(3))(NHPh) (3) in 88% isolated yield. The amido complex 3 has been characterized by (1)H NMR, (13)C NMR, (31)P NMR, elemental analysis, cyclic voltammetry, and a solid-state X-ray diffraction study. Variable temperature NMR studies have revealed a rotational barrier around the ruthenium-amido nitrogen bond of 3 of 12 kcal/mol (transformation of the major isomer to the minor isomer). The solid-state structure of 3 discloses a pyramidal amido moiety. Heating benzene solutions of the amido complex 3 and 1,4-cyclohexadiene or 9,10-dihydroanthracene results in no observable reaction. Reaction of complex 2 with excess aniline yields [TpRu(CO)(PPh(3))(NH(2)Ph)][OTf] (4).}, number={25}, journal={INORGANIC CHEMISTRY}, author={Jayaprakash, KN and Gunnoe, TB and Boyle, PD}, year={2001}, month={Dec}, pages={6481–6486} } @article{kockerling_willems_boyle_2001, title={[(Zr6B)Cl11-xI2+x] (0 <= x <= 6): A new mixed-halide structure with zigzag chains of clusters in multiply twinned crystals}, volume={40}, ISSN={["1520-510X"]}, DOI={10.1021/ic991359v}, abstractNote={The new [(Zr6B)Cl11-xI2+x] phase (with 0 < or = x < or = 6) is obtained from reactions of ZrI4, ZrCl4, and elemental Zr and B for 2-4 weeks in sealed Ta tubing at 800-850 degrees C. Single crystals of [(Zr6B)Cl6.44(7)I6.56] have been characterized by X-ray diffraction at room temperature (orthorhombic Pbcn, Z = 4, a = 12.365(2) A, b = 15.485(3) A, c = 13.405(2) A). This structure contains zigzag chains of boron-centered (Zr6B) octahedra that are interconnected by Cl(i-i) halides. Further three-dimensional connectivity is achieved by I(a-a-a) bridges. The noncluster interconnecting two-bonded X(i) sites are occupied statistically by a mixture of Cl and I. For each site both positions were resolved. This structure forms within a phase width of 0 < or = x < or = 6 at temperatures between 800 and 850 degrees C. Crystals of this phase appear to be always multiply twinned.}, number={7}, journal={INORGANIC CHEMISTRY}, author={Kockerling, M and Willems, JB and Boyle, PD}, year={2001}, month={Mar}, pages={1439–1444} } @article{shull_spielvogel_gopalaswamy_sankar_boyle_head_devito_2000, title={Evidence for spontaneous, reversible paracyclophane formation. Aprotic solution structure of the boron neutron capture therapy drug, L-p-boronophenylalanine}, ISSN={["0300-9580"]}, DOI={10.1039/a906038c}, abstractNote={The simple, straightforward 1H NMR spectrum of the neutron capture therapy drug, L-p-boronophenylalanine (L-p-BPA), in D2O changes to a more complex one in DMSO-d6 in which the ratio of new species observed is highly concentration dependent. The new species detected can only be explained by an additional stereocenter being formed at the boron center by intermolecular chelation of the amino acid of another molecule of L-p-BPA. This gives rise to the presence of an oligomeric species as well as another whose aromatic protons appear as a pair of sharp AB quartets centered further upfield at 6.66 and 6.84. Due to these shifts and couplings observed between the benzylic protons and the proton at the stereocenter of the amino acid, it is proposed that this species is a paracyclophane dimer of L-p-BPA in which one molecule of L-p-BPA chelates head-to-tail with another. This cyclophane dimer predominates in low concentrations ( 90 mM) concentrations. The formation of these two species is completely reversible, the addition of D2O completely regenerating L-p-BPA. Variable temperature 1H NMR found that the two pairs of aromatic protons of the cyclophane dimer coalesce at Tc = 141 °C, corresponding to a ΔG‡ = 20.6 kcal mol−1.}, number={3}, journal={JOURNAL OF THE CHEMICAL SOCIETY-PERKIN TRANSACTIONS 2}, author={Shull, BK and Spielvogel, DE and Gopalaswamy, R and Sankar, S and Boyle, PD and Head, G and Devito, K}, year={2000}, pages={557–561} } @article{balasubramanian_strachan_boyle_lindsey_2000, title={Rational synthesis of beta-substituted chlorin building blocks}, volume={65}, ISSN={["0022-3263"]}, DOI={10.1021/jo000913b}, abstractNote={Chlorins provide the basis for plant photosynthesis, but synthetic model systems have generally employed porphyrins as surrogates due to the unavailability of suitable chlorin building blocks. We have adapted a route pioneered by Battersby to gain access to chlorins that bear two meso substituents, a geminal dimethyl group to lock in the chlorin hydrogenation level, and no flanking meso and beta substituents. The synthesis involves convergent joining of an Eastern half and a Western half. A 3,3-dimethyl-2,3-dihydrodipyrrin (Western half) was synthesized in four steps from pyrrole-2-carboxaldehyde. A bromodipyrromethane carbinol (Eastern half) was prepared by sequential acylation and bromination of a 5-substituted dipyrromethane followed by reduction. Chlorin formation is achieved by a two-flask process of acid-catalyzed condensation followed by metal-mediated oxidative cyclization. The latter reaction has heretofore been performed with copper templates. Investigation of conditions for this multistep process led to copper-free conditions (zinc acetate, AgIO(3), and piperidine in toluene at 80 degrees C for 2 h). The zinc chlorin was obtained in yields of approximately 10% and could be easily demetalated to give the corresponding free base chlorin. The synthetic process is compatible with a range of meso substituents (p-tolyl, mesityl, pentafluorophenyl, 4-[2-(trimethylsilyl)ethynyl]phenyl, 4-iodophenyl). Altogether four free base and four zinc chlorins have been prepared. The chlorins exhibit typical absorption spectra, fluorescence spectra, and fluorescence quantum yields. The ease of synthetic access, presence of appropriate substituents, and characteristic spectral features make these types of chlorins well suited for incorporation in synthetic model systems.}, number={23}, journal={JOURNAL OF ORGANIC CHEMISTRY}, author={Balasubramanian, T and Strachan, JP and Boyle, PD and Lindsey, JS}, year={2000}, month={Nov}, pages={7919–7929} } @article{gorman_su_jiang_watson_boyle_1999, title={Hybrid organic-inorganic, hexa-arm dendrimers based on an Mo6Cl8 core}, ISSN={["1359-7345"]}, DOI={10.1039/a901590f}, abstractNote={Dendrons with focal phenoxide groups were shown to substitute for triflate or methoxide ligands around an Mo6Cl8 core to form molecules of the form Mo6(µ3-Cl)8(OR)6 where R = dendrons with zero to two hyperbranches, respectively; these molecules represent a new type of metal cluster core dendrimer with six arms and high symmetry.}, number={10}, journal={CHEMICAL COMMUNICATIONS}, author={Gorman, CB and Su, WY and Jiang, HW and Watson, CM and Boyle, P}, year={1999}, month={May}, pages={877–878} } @article{littler_miller_hung_wagner_df o'shea_boyle_lindsey_1999, title={Refined synthesis of 5-substituted dipyrromethanes}, volume={64}, ISSN={["0022-3263"]}, DOI={10.1021/jo982015+}, abstractNote={5-Substituted dipyrromethanes are important precursors for the synthesis of meso-substituted porphyrins,1 expanded porphyrins, and porphyrin analogues.2 Several one-flask methods have been reported for the synthesis of 5-substituted dipyrromethanes by the condensation of an aldehyde and pyrrole using various combinations of acids and solvents.3-14 We previously reported a one-flask synthesis of dipyrromethanes in which an aldehyde is dissolved in a 40-fold excess of pyrrole with a catalytic amount of an acid at room temperature in the absence of any other solvent.8 This method has afforded good yields of 5-substituted dipyrromethanes bearing many types of functional groups.15 However, purification of the product is typically achieved by flash column chromatography, restricting application to the small-scale preparation of dipyrromethanes. Our objective in this study was to eliminate the use of chromatography during purification, thereby removing the major bottleneck to the synthesis of multigram quantities of dipyrromethanes. Upon examining the crude reaction mixture from the acid-catalyzed condensation of benzaldehyde in excess pyrrole, we found that the principal reaction products consist of the dipyrromethane (1), the N-confused dipyrromethane (2,3′-dipyrromethane) (2) and the tripyrrane (3) (Scheme 1). The presence of the N-confused dipyrromethane was surprising as this species had not been detected previously.8 Using GC we have examined the distribution of these products as a function of the pyrrole:benzaldehyde ratio and the acid catalyst. We have developed a purification process based on bulb-to-bulb distillation followed by recrystallization that affords analytically pure dipyrromethanes in multigram quantities. The condensation of pyrrole and benzaldehyde upon heating without added acids has also been examined.}, number={4}, journal={JOURNAL OF ORGANIC CHEMISTRY}, author={Littler, BJ and Miller, MA and Hung, CH and Wagner, RW and DF O'Shea and Boyle, PD and Lindsey, JS}, year={1999}, month={Feb}, pages={1391–1396} } @article{keefer_bereman_purrington_knight_boyle_1999, title={The Pt-195 NMR of L2Pt(1,2-dithiolene) complexes}, volume={38}, ISSN={["0020-1669"]}, DOI={10.1021/ic980369d}, abstractNote={The syntheses and characterizations of the novel platinum(II) mono-1,2-dithiolenes (COD)Pt(dddt) (1), (Ph3P)2Pt(dddt) (2), (COD)Pt(edt) (3), (Ph3P)2Pt(edt) (4), (bipy)Pt(edt) (5), and (Ph3P)(CO)Pt(dddt) (6) (COD = 1,5-cyclooctadiene; dddt = 5,6-dihydro-1,4-dithiin-2,3-dithiolate; edt = ethylene-1,2-dithiolate, bipy = 2,2‘-bipyridyl) are reported. 195Pt NMR spectral analysis was performed on the above-mentioned compounds along with the previously reported compounds (COD)Pt(dmid) (7), (Ph3P)2Pt(dmid) (8), (Ph3P)2Pt(dmit) (9), (COD)Pt(mnt) (10), (Ph3P)2Pt(mnt) (11), (COD)Pt(dt) (12), and (Ph3P)2Pt(dt) (13) (dmid = 1,3-dithiole-2-oxo-4,5-dithiolate; dmit = 1,3-dithiole-2-thione-4,5-dithiolate; mnt = maleonitrile-1,2-dithiolate; dt = ethane-1,2-dithiolate). 195Pt NMR results show that, depending on the nature of L2 in L2Pt(1,2-dithiolene) complexes, the 1,2-dithiolene ligands behave as either π donors or acceptors toward the Pt metal center. 195Pt NMR is also sensitive to the relative electron-withdrawing abil...}, number={10}, journal={INORGANIC CHEMISTRY}, author={Keefer, CE and Bereman, RD and Purrington, ST and Knight, BW and Boyle, PD}, year={1999}, month={May}, pages={2294–2302} } @article{keefer_purrington_bereman_boyle_1999, title={The first systematic synthesis of heterobimetallic dithiolene-bridged complexes. Synthesis and characterization of metal complexes of 4-(1 ',2 '-ethylenedithiolate)-1,3-dithiole-2-one and dimeric metal complexes of 1,2,3,4-butadienetetrathiolate}, volume={38}, ISSN={["0020-1669"]}, DOI={10.1021/ic990841a}, abstractNote={The 1,2-dithiolene ligands 4-(1‘,2‘-ethylenedithiolate)-1,3-dithiole-2-one (eddo2-) and 1,2,3,4-butadienetetrathiolate (bdt4-) are synthesized by the controlled single or double 1,3-dithiole-2-one ring opening of 4,4‘-bis(1,3-dithiole-2-one) (bdo) (1). The synthesis and characterization of the novel transition metal complexes (COD)Pt(eddo) (2), (diphos)Ni(eddo) (3), [Bu4N][Ni(eddo)2] (4), (COD)Pt(bdt)Pt(COD) (5), and (diphos)Ni(bdt)Pt(COD) (6) (where COD = 1,5-cyclooctadiene, diphos = 1,2-bis(diphenylphosphino)ethane) are reported. Use of eddo2- and bdt4- results in the selective systematic synthesis of transition metal monomers and dimers. In addition, the synthesis of mixed metal dimeric complexes utilizing a “transition metal 1,2-dithiolene” ligand is demonstrated. The single-crystal X-ray structural analyses of the cocrystallized (COD)Pt(eddo)·1/2bdo, monoclinic, P21/n, a = 13.0388(4) A, b = 9.2048(3) A, c = 16.2384(5) A, β = 97.289(5)°, Z = 4, and [Bu4N][Ni(eddo)2], monoclinic, P21/n, a = 8.3508(2) A...}, number={23}, journal={INORGANIC CHEMISTRY}, author={Keefer, CE and Purrington, ST and Bereman, RD and Boyle, PD}, year={1999}, month={Nov}, pages={5437–5442} } @article{cornman_jantzi_wirgau_stauffer_kampf_boyle_1998, title={Coordination chemistry of a tripodal S2ON ligand: Syntheses, structures, and reactivity of the molybdenum(VI) and nickel(II) complexes of bis(2-mercaptoethyl)-2-amino-4-methylphenol (H(3)btap) and comparison to (VO)-O-v(btap)}, volume={37}, ISSN={["0020-1669"]}, DOI={10.1021/ic9802956}, abstractNote={The tripodal tetradentate ligand H3btap coordinates to VV, MoVI, and NiII via three different bonding modes to yield three complexes with unique ligand-based oxidation chemistry. For VV and MoVI (1), all four of the heteroatom donors are coordinated to the metal ion forming a trigonal bipyramidal complex with the oxovanadium(V) ion, VVO3+, and an octahedral complex with the cis-dioxomolybdenum(VI) ion, [MoO2]2+. Only three of the heteroatom donors of H3btap are used to coordinate to NiII (2), two thiolate sulfurs and the amine nitrogen, yielding a dimeric structure in which each nickel(II) ion has NS3 coordination. The ability of VVO(btap) to form η2-sulfenates, while [MoO2(btap)]- does not form stable η2-sulfenates, has been ascribed to the electron-deficient, π-accepting nature of VVO3+ relative to [MoVIO2]2+. Crystal data for 1 (C11H16NO4S2KMo):  space group Pbcn, a = 6.6596(9) A, b = 13.7446(9) A, c = 32.992(2) A, α = β = γ = 90°, Z = 8. Crystal data for 2 (C24H38N2O4S4Ni):  space group Pbcn, a = 12.0...}, number={22}, journal={INORGANIC CHEMISTRY}, author={Cornman, CR and Jantzi, KL and Wirgau, JI and Stauffer, TC and Kampf, JW and Boyle, PD}, year={1998}, month={Nov}, pages={5851–5855} } @article{comins_lee_boyle_1998, title={Intramolecular photocycloaddition of a tethered bis-2,3-dihydro-4-pyridone: Stereochemistry and reactivity of the cycloadduct}, volume={39}, ISSN={["0040-4039"]}, DOI={10.1016/S0040-4039(97)10520-2}, abstractNote={Abstract The photocycloaddition of enantiopure bis-dihydropyridone 4 provided a single cycloadduct in high yield. Treatment of the cycloadduct with SmI2 effected ring opening and intramolecular aldol addition to give pentacyclic β-hydroxyketone 7.}, number={3-4}, journal={TETRAHEDRON LETTERS}, author={Comins, DL and Lee, YS and Boyle, PD}, year={1998}, month={Jan}, pages={187–190} } @article{ford_ebisuzaki_boyle_1998, title={Methylxanthines. II. Anhydrous theobromine}, volume={54}, ISSN={["0108-2701"]}, DOI={10.1107/S0108270198009469}, abstractNote={Clear crystals of anhydrous theobromine (3,7-dihydro-3,7-dimethyl-1H-purine-2,6-dione, C 7 H 8 N 4 O 2 ) were grown by vacuum sublimation and the structure determined. The melting point of anhydrous theobromine has been determined as 620 (1) K. Two molecules in the asymmetric unit form a pseudo-centrosymmetric dimer and pack to form a layered structure of two-dimensional hydrogen-bonded networks.}, journal={ACTA CRYSTALLOGRAPHICA SECTION C-CRYSTAL STRUCTURE COMMUNICATIONS}, author={Ford, KA and Ebisuzaki, Y and Boyle, PD}, year={1998}, month={Dec}, pages={1980–1983} } @article{summers_roe_boyle_colvin_shaw_1998, title={Structural studies of a borane-modified phosphate diester linkage: Ab initio calculations on the dimethylboranophosphate anion and the single-crystal X-ray structure of its diisopropylammonium salt}, volume={37}, ISSN={["1520-510X"]}, DOI={10.1021/ic971435z}, abstractNote={The dimethylboranophosphate anion was characterized by ab initio calculations and by X-ray crystallography. The 1.905 A P−BH3 bond is longer and less polar than the 1.51 A P−O bond of the analogous normal phosphate diester; the structures are otherwise quite similar. In the crystal, the phosphoryl oxygen forms hydrogen bonds with acidic protons of neighboring cations; the borane does not form hydrogen bonds. The relevance of the structure to boranophosphate nucleotide analogues is discussed.}, number={17}, journal={INORGANIC CHEMISTRY}, author={Summers, JS and Roe, D and Boyle, PD and Colvin, M and Shaw, BR}, year={1998}, month={Aug}, pages={4158–4159} } @article{keefer_purrington_bereman_knight_bedgood_boyle_1998, title={The synthesis and characterization of platinum monodithiolene complexes containing 1,3-dithiole-2-oxo-4,5-dithiolate (dmid(2-)) and 1,3-dithiole-2-thione-4,5-dithiolate (dmit(2-))}, volume={282}, ISSN={["0020-1693"]}, DOI={10.1016/S0020-1693(98)00244-8}, abstractNote={The platinum monodithiolene complexes (COD)Pt(dmid) (1), (COD)Pt(dmit) (2), (Ph3P)2Pt(dmid) (3) and (Ph3P)2Pt(dmit) (4) (where COD = 1.5-cyclooctadiene, dmid = 1,3-dithiole-2-oxo-4,5-dithiolate, dmit = 1,3-dithiole-2-thione-4,5-dithiolate) have been synthesized and characterized. We report here their 1H NMR, 195Pt NMR, UV-Vis spectra and electrochemistry. The X-ray crystal structures are also reported. 2: orthorhombic, Cmcm, a = 10.8535(6), b = 16.8503(7), c = 7.3622(3)Å, Z = 4, 3: triclinic, P1¯, a = 11.2051(4), b = 11.4409(5), c = 15.2212(9)Å, α = 71.882(4), β = 73.125(4), γ = 84.589(6)°, Z = 2. 4·CHCl3: triclinic, P1¯, a = 10.2521(5), b = 13.9387(14), c = 15.6884(18)Å, α = 67.154(9), β = 89.380(15), γ = 83.570(13)°, Z = 2. The crystal structure of 2 exhibits a molecular stacking pattern which has alternating Pt and S(thione) atoms forming a one-dimensional chain which suggests the presence of a weak axial interaction between the Pt and S(thione). A correlation exists between the 195Pt NMR chemical shifts and the one-electron oxidation potentials of these complexes. In addition, it is shown that the carbonyl of the dmid2− ligand is not susceptible to alkoxide attack when coordinated to platinum.}, number={2}, journal={INORGANICA CHIMICA ACTA}, author={Keefer, CE and Purrington, ST and Bereman, RD and Knight, BW and Bedgood, DR and Boyle, PD}, year={1998}, month={Nov}, pages={200–208} } @article{martin_leafblad_sullivan_boyle_1998, title={alpha- and beta-CuAlCl4: Framework construction using corner-shared tetrahedral metal-halide building blocks}, volume={37}, ISSN={["0020-1669"]}, DOI={10.1021/ic971148v}, abstractNote={Rapid quenching of a melt of CuCl and AlCl(3) results in the formation of the metastable framework structure, beta-CuAlCl(4). The structure, determined by single-crystal X-ray crystallography (space group Pna2(1), a = 12.8388(5) Å, b = 7.6455(3) Å, and c = 6.1264(3) Å, Z = 2), can be derived from a distorted hexagonal closest packed anion sublattice. Annealing at temperatures above 100 degrees C results in a phase transition to the more thermodynamically stable alpha-CuAlCl(4). The solid solution CuAlCl(4)(-)(x)()Br(x)() is described for both alpha and beta phases. The structures of alpha-CuAlCl(4) and alpha-CuAlBr(4), determined by single-crystal X-ray diffraction (space group P&fourmacr;2c, a = b 5.4409(1) Å and c = 10.1126(3) Å, V = 299.37(1) Å(3), Z = 2, and a = b = 5.7321(2) Å and c = 10.6046(8) Å, Z = 2, respectively), are derived from a distorted cubic closest packed anion sublattice. The mechanism for this phase transition is described in relation to that previously described for cristobalite-type structures. The structures of both alpha- and beta-CuAlCl(4) reveal large van der Waals channels, which are proposed to be important for the reversible adsorption of carbon monoxide and ethylene by these materials.}, number={6}, journal={INORGANIC CHEMISTRY}, author={Martin, JD and Leafblad, BR and Sullivan, RM and Boyle, PD}, year={1998}, month={Mar}, pages={1341–1346} } @article{ebisuzaki_boyle_smith_1997, title={Methylxanthines .1. Anhydrous theophylline}, volume={53}, ISSN={["0108-2701"]}, DOI={10.1107/S0108270197001960}, abstractNote={Clear crystals of anhydrous theophylline (3,7-dihydro-1,3-dimethyl-1H-purine-2,6-dione), C 7 H 8 N 4 O 2 , were grown by sublimation and the X-ray structure determined. The planar molecules are stacked in columns along [010] and are inclined at an angle of 26.05 (10)° relative to (010). The melting point has been determined to be 548.8 (5) K.}, journal={ACTA CRYSTALLOGRAPHICA SECTION C-CRYSTAL STRUCTURE COMMUNICATIONS}, author={Ebisuzaki, Y and Boyle, PD and Smith, JA}, year={1997}, month={Jun}, pages={777–779} } @article{cornman_stauffer_boyle_1997, title={Oxidation of a vanadium(V)-dithiolate complex to a vanadium(V)-eta(2),eta(2)-disulfenate complex}, volume={119}, ISSN={["0002-7863"]}, DOI={10.1021/ja971007s}, abstractNote={ADVERTISEMENT RETURN TO ISSUEPREVCommunicationNEXTOxidation of a Vanadium(V)−Dithiolate Complex to a Vanadium(V)−η2,η2-Disulfenate ComplexCharles R. Cornman, Thad C. Stauffer, and Paul D. BoyleView Author Information Department of Chemistry North Carolina State University Raleigh, North Carolina 27695-8204 Cite this: J. Am. Chem. Soc. 1997, 119, 25, 5986–5987Publication Date (Web):June 25, 1997Publication History Received31 March 1997Published online25 June 1997Published inissue 1 June 1997https://doi.org/10.1021/ja971007sCopyright © 1997 American Chemical SocietyRIGHTS & PERMISSIONSArticle Views271Altmetric-Citations30LEARN ABOUT THESE METRICSArticle Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated. Share Add toView InAdd Full Text with ReferenceAdd Description ExportRISCitationCitation and abstractCitation and referencesMore Options Share onFacebookTwitterWechatLinked InReddit Read OnlinePDF (106 KB) Get e-AlertsSupporting Info (2)»Supporting Information Supporting Information SUBJECTS:Group theory,Ligands,Oxidation,Peptides and proteins,Thiolates Get e-Alerts}, number={25}, journal={JOURNAL OF THE AMERICAN CHEMICAL SOCIETY}, author={Cornman, CR and Stauffer, TC and Boyle, PD}, year={1997}, month={Jun}, pages={5986–5987} } @article{overby_hanusa_boyle_1997, title={Stabilization of the (pi-indenyl)-lead bond: The first structurally authenticated bis(eta(5)-indenyl) complex of a post-transition element, [Pb{1,3-(SiMe3)(2)C9H5}(2)]}, volume={36}, ISSN={["1521-3773"]}, DOI={10.1002/anie.199723781}, abstractNote={Where are all the bis(π-indenyl) complexes of the “post-transition elements”? The first example of this type of complex with a Group 14 element has now been synthesized and characterized. [Pb{1,3-(SiMe3)2C9H5}2] is unstable in solution, but forms thermally stable crystals that reveal a sandwich geometry with nearly parallel, η5-coordinated C5 rings of the indenyl ligand (see structure on the right).}, number={21}, journal={ANGEWANDTE CHEMIE-INTERNATIONAL EDITION}, author={Overby, JS and Hanusa, TP and Boyle, PD}, year={1997}, month={Nov}, pages={2378–2379} } @article{cornman_geiser-bush_rowley_boyle_1997, title={Structural and electron paramagnetic resonance studies of the square pyramidal to trigonal bipyramidal distortion of vanadyl complexes containing sterically crowded Schiff base ligands}, volume={36}, ISSN={["0020-1669"]}, DOI={10.1021/ic970868z}, abstractNote={[N,N‘-Ethylenebis(o-(tert-butyl-p-methylsalicylaldiminato)]oxovanadium(IV) (1), [N,N‘-propanediylbis(o-(tert-butyl-p-methylsalicylaldiminato)]oxovanadium(IV) (2), bis(N-methylsalicylaldiminato)oxovanadium(IV) (3), bis(N-isopropyl-o-methylsalicylaldiminato)oxovanadium(IV) (4), and, bis(N-methyl-o-(tert-butyl-p-methylsalicylaldiminato)oxovanadium(IV) (5) were prepared and characterized by X-ray crystallography and EPR spectroscopy. Complexes 1 and 2 are best described as square pyramids, while complexes 3−5 are distorted trigonal bipyramids, demonstrating that oxovanadium(IV) complexes can readily adopt a trigonal bipyramidal geometry. All five compounds give nearly the same parallel hyperfine coupling constant (Az) regardless of the fact that the geometry about the vanadium changes from square pyramidal to trigonal bipyramidal. Crystal data for 1:  space group P1, a = 7.9382(3) A, b = 12.6749(7) A, c = 13.8353(7) A, α = 109.608(5)°, β = 96.552(5)°, γ = 96.589(5)°, Z = 2. Crystal data for 2:  space group I...}, number={27}, journal={INORGANIC CHEMISTRY}, author={Cornman, CR and Geiser-Bush, KM and Rowley, SP and Boyle, PD}, year={1997}, month={Dec}, pages={6401–6408} }