@article{lilly_ormond_carter_powell_2022, title={Implementation of Recitations in General Chemistry I Laboratory Courses to Increase Student Performance}, volume={99}, ISSN={0021-9584 1938-1328}, url={http://dx.doi.org/10.1021/acs.jchemed.1c00542}, DOI={10.1021/acs.jchemed.1c00542}, number={5}, journal={Journal of Chemical Education}, publisher={American Chemical Society (ACS)}, author={Lilly, Cassandra P. and Ormond, Alexandra B. and Carter, Andrea A. and Powell, Walda J.}, year={2022}, month={Apr}, pages={1838–1846} } @article{brown_lilly_lambic_sommer_ison_2020, title={Synthesis and Reactivity of Re(III) and Re(V) Fischer Carbenes}, volume={39}, ISSN={["1520-6041"]}, DOI={10.1021/acs.organomet.9b00600}, abstractNote={Direct insertion of CO and isocyanides, RNC, into Re–R bonds results in high-oxidation-state acyl and iminoacyl complexes that can be treated with an electrophile to generate rare examples of rheni...}, number={3}, journal={ORGANOMETALLICS}, author={Brown, Caleb A. and Lilly, Cassandra P. and Lambic, Nikola S. and Sommer, Roger D. and Ison, Elon A.}, year={2020}, month={Feb}, pages={388–396} } @article{stromyer_lilly_dillner_knaust_2016, title={Crystal structures of [Ln(NO3)32-bpydo)2], whereLn= Ce, Pr or Nd, and bpydo = 4,4′-bipyridineN,N′-dioxide: layered coordination networks containing 44grids}, volume={72}, ISSN={2056-9890}, url={http://dx.doi.org/10.1107/s205698901502318x}, DOI={10.1107/s205698901502318x}, abstractNote={The crystal structures of poly[bis(μ2-4,4′-bipyridine N,N′-dioxide-κ2 O:O′)trinitratocerium(III)] and its isostructural praseodymium and neodymium analogues feature a 44 grid-like layered structure with interdigitation of layers promoted by C—H⋯O interactions between nitrate anions and 4,4′-bipyridine N,N′-dioxide ligands.}, number={1}, journal={Acta Crystallographica Section E Crystallographic Communications}, publisher={International Union of Crystallography (IUCr)}, author={Stromyer, Michael L. and Lilly, Cassandra P. and Dillner, Adam J. and Knaust, Jacqueline M.}, year={2016}, month={Jan}, pages={25–30} } @article{robbins_lilly_sommer_ison_2016, title={Effect of the Ancillary Ligand on the Mechanism for CO Migratory Insertion in High-Valent Oxorhenium Complexes}, volume={35}, ISSN={["1520-6041"]}, DOI={10.1021/acs.organomet.6b00570}, abstractNote={Several oxorhenium complexes bearing an SSS pincer ligand were isolated and characterized, and their reactivity with carbon monoxide was explored. The corresponding oxorhenium(V) acyl derivatives were also isolated and characterized. Carbonylation reactions required high pressures (400 psi) and temperatures (50 °C). The mechanism for carbonylation was explored with DFT (M06) calculations and revealed that the most likely mechanism for carbonylation involved stepwise formation of CO adducts followed by migration of the carbonyl ligand to the alkyl/aryl groups.}, number={20}, journal={ORGANOMETALLICS}, publisher={American Chemical Society (ACS)}, author={Robbins, Leanna K. and Lilly, Cassandra P. and Sommer, Roger D. and Ison, Elon A.}, year={2016}, month={Oct}, pages={3530–3537} } @article{lambic_lilly_sommer_ison_2016, title={Mechanism for the Reaction of CO with Oxorhenium Hydrides: Migratory Insertion of CO into Rhenium Hydride and Formyl Bonds leads to Migration from Rhenium to the Oxo Ligand}, volume={35}, ISSN={["1520-6041"]}, DOI={10.1021/acs.organomet.6b00591}, abstractNote={Computational studies (M06) have been performed in synergy with experimental studies to show that the thermodynamics for insertion of CO into an oxorhenium–hydride bond to form a formyl ligand is favorable despite conventional wisdom to the contrary. Further, it is shown that insertion of CO into formyl ligands to form α-dicarbonyl ligands is also a viable pathway and results in hydroxy carbonyl or formate complexes, depending on the nature of the ancillary ligand.}, number={17}, journal={ORGANOMETALLICS}, publisher={American Chemical Society (ACS)}, author={Lambic, Nikola S. and Lilly, Cassandra P. and Sommer, Roger D. and Ison, Elon A.}, year={2016}, month={Sep}, pages={3060–3068} } @article{lambic_lilly_robbins_sommer_ison_2016, title={Reductive Carbonylation of Oxorhenium Hydrides Induced by Lewis Acids}, volume={35}, ISSN={["1520-6041"]}, DOI={10.1021/acs.organomet.6b00393}, abstractNote={Several oxorhenium hydride complexes with chelating diamidopyridine (DAP), diamidoamine (DAAm), and 2-mercaptoethyl sulfide (SSS) groups have been isolated and characterized. Adduct formation is observed when the DAP complex 1a is treated with the Lewis acid B(C6F5)3. However, treatment of 1a,b with B(C6F5)3 or BF3·OEt2 in the presence of CO results in reduction of the metal center by four electrons from Re(V) to Re(I).}, number={17}, journal={ORGANOMETALLICS}, publisher={American Chemical Society (ACS)}, author={Lambic, Nikola S. and Lilly, Cassandra P. and Robbins, Leanna K. and Sommer, Roger D. and Ison, Elon A.}, year={2016}, month={Sep}, pages={2822–2829} } @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{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{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{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{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{dillner_lilly_knaust_2010, title={Poly[[tris(μ2-4,4′-bipyridineN,N′-dioxide)hexanitratodieuropium(III)] dichloromethane disolvate]}, volume={66}, ISSN={1600-5368}, url={http://dx.doi.org/10.1107/s1600536810033246}, DOI={10.1107/s1600536810033246}, abstractNote={The title one-dimensional coordination network, {[Eu2(NO3)6(C10H8N2O2)3]·2CH2Cl2}n, is isostructural with the previously reported Tb and Tl coordination networks and to its Gd analog. The EuIII cation is coordinated in a distorted tricapped trigonal-prismatic fashion by nine O atoms from three bridging 4,4′-bipyridine N,N′-dioxide ligands and three chelating nitrate anions. None of the atoms lie on a special position, but there is an inversion center located between the rings of one of the ligands. The network topology is ladder-like, and each ladder interacts with six neighboring ladders through C—H⋯O hydrogen bonds. The packing motif of the ladders allows for the formation of channels that run parallel to the a axis; these channels are filled with CH2Cl2 solvent molecules that interact with the ladders through C—H⋯O hydrogen bonds.}, number={9}, journal={Acta Crystallographica Section E Structure Reports Online}, publisher={International Union of Crystallography (IUCr)}, author={Dillner, Adam J. and Lilly, Cassandra P. and Knaust, Jacqueline M.}, year={2010}, month={Aug}, pages={m1156–m1157} } @article{dillner_lilly_knaust_2010, title={Poly[[tris(μ2-4,4′-bipyridineN,N′-dioxide)hexanitratodigadolinium(III)] dichloromethane disolvate]}, volume={66}, ISSN={1600-5368}, url={http://dx.doi.org/10.1107/s1600536810033258}, DOI={10.1107/s1600536810033258}, abstractNote={The title one-dimensional coordination network, {[Gd2(NO3)6(C10H8N2O2)3]·2CH2Cl2}n, is isostructural with the previously reported Tb and Tl coordination networks and to its Eu analog. The GdIII cation is coordinated in a distorted tricapped trigonal-prismatic fashion by nine O atoms from three bridging 4,4′-bipyridine N,N′-dioxide ligands and three chelating nitrate anions. None of the atoms lie on a special position, but there is an inversion center located between the rings of one of the ligands. The network topology is ladder-like, and each ladder interacts with six neighboring ladders through C—H⋯O hydrogen bonds. The packing motif of the ladders allows for the formation of channels that run parallel to the a axis; these channels are filled with CH2Cl2 solvent molecules that interact with the ladders through C—H⋯O hydrogen bonds}, number={9}, journal={Acta Crystallographica Section E Structure Reports Online}, publisher={International Union of Crystallography (IUCr)}, author={Dillner, Adam J. and Lilly, Cassandra P. and Knaust, Jacqueline M.}, year={2010}, month={Aug}, pages={m1158–m1159} }