57 results on '"Pink M"'
Search Results
2. Influence of the [2.1.1]-(2,6)-pyridinophane macrocycle ring size constant on the structure and reactivity of copper complexes
- Author
-
Vendernikov, A.N., Pink, M., and Caulton, K.G.
- Subjects
Copper compounds -- Structure ,Copper compounds -- Chemical properties ,Oxidation-reduction reaction -- Analysis ,Pyridine -- Chemical properties ,Chemistry - Abstract
The geometric distortions imposed by the pyridinophane macrocycle on Cu(I) and Cu(II) is analyzed and the reactivity of the LCu(super +) fragment toward the dioxygen is reported. It is suggested that O(sub 2) oxidizes the Cu(II) species by an inner sphere mechanism through coordinated O(sub 2) and the impact of macrocycle is not much in reactant destabilization but it is occurs when copper reaches higher oxidation states.
- Published
- 2004
3. Influence of the [2.1.1]-(2,6)-Pyridinophane Macrocycle Ring Size Constraint on the Structure and Reactivity of Copper Complexes
- Author
-
Vedernikov, A. N., primary, Pink, M., additional, and Caulton, K. G., additional
- Published
- 2004
- Full Text
- View/download PDF
4. Powder Structure Solutions of the Compounds Potassium Phenoxide−Phenol: C6H5OK·xC6H5OH (x= 2, 3)
- Author
-
Dinnebier, R. E., primary, Pink, M., additional, Sieler, J., additional, Norby, P., additional, and Stephens, P. W., additional
- Published
- 1998
- Full Text
- View/download PDF
5. Powder Structure Solutions of the Compounds Potassium Phenoxide-Phenol: C6H5OK.xC6H5OH(x = 2,3).
- Author
-
Dinnebier, R.E. and Pink, M.
- Subjects
- *
X-ray diffraction , *POTASSIUM - Abstract
Focuses on the use of the high-resolution powder x-ray diffraction in reporting the ab initio structure solutions, relating to the solvent which contains potassium phenoxides. Industrial importance of the carboxylation of alkali phenolates, known as Kolbe-Schmitt synthesis; Sublimation of phenol; Similarity of coordination of the potassium ions for both compounds; Information on these structure solutions.
- Published
- 1998
- Full Text
- View/download PDF
6. Trimethylsilyldiazomethane Disassembly at a Three-Fold Symmetric Iron Site.
- Author
-
Chivington AD, Squire S, Yamamoto N, Pink M, Griffith MD, Fletcher J, Gao Y, Zadrozny JM, and Smith JM
- Abstract
The reaction of equimolar trimethylsilyldiazomethyllithium (LiTMSD) with high spin ( S = 2) PhB(AdIm)
3 FeCl (PhB(AdIm)3 - = tris(3-adamantylimidazol-2-ylidene)phenylborate) affords the corresponding N- nitrilimido complex PhB(AdIm)3 Fe-N═N═C(SiMe3 ). This complex can be converted to the thermodynamically more favorable C -isocyanoamido isomer PhB(AdIm)3 Fe-C═N═N(SiMe3 ) by reaction with an additional equivalent of LiTMSD. While the iron(II) complexes are four-coordinate, the diazomethane is bound side-on in the iron(I) congener PhB(AdIm)3 Fe( N , N '-κ2 -N2 C(H)Si(CH3 )3 ). The latter complex adopts high spin ( S = 3/2) ground state and features an unusually weak C-H bond. Photolysis of the iron(II) complexes induces N═N bond cleavage, with the iron(II) cyanide PhB(AdIm)3 Fe-C≡N and iron(IV) nitride PhB(AdIm)3 Fe≡N complexes being the major products of the reaction. The same products are obtained when the iron(I) complex is photolyzed or treated with a fluoride source. The trimethylsilyldiazomethane-derived ligand disassembly reactions are contrasted with those observed for related tris(carbene)amine complexes.- Published
- 2024
- Full Text
- View/download PDF
7. Redox-Neutral Transformations of Carbon Dioxide Using Coordinatively Unsaturated Late Metal Silyl Amide Complexes.
- Author
-
Huerfano IJ, Laskowski CA, Pink M, Carta V, Hillhouse GL, Caulton KG, and Smith JM
- Abstract
Two-coordinate silylamido complexes of nickel and copper rapidly react with CO
2 to selectively form a new cyanate ligand along with hexamethyldisiloxane byproducts. Mechanistic insight into these reactions was obtained from the synthesis of proposed intermediates, several silyl- and phenyl- substituted amido analogues, and their subsequent reactivity with CO2 . These studies suggest that a unique intramolecular double silyl transfer step facilitates CO2 deoxygenation, which likely contributes to the rapid rates of reaction. The deoxygenation reactions create a platform for a synthetic cycle in which copper amido complexes convert CO2 to organic silylcarbamates.- Published
- 2022
- Full Text
- View/download PDF
8. Facile Addition of B-H and B-B Bonds to an Iron(IV) Nitride Complex.
- Author
-
Tran BG, Carta V, Pink M, Caulton KG, and Smith JM
- Subjects
- Ligands, Kinetics, Iron chemistry, Boranes
- Abstract
The nitride ligand in the iron(IV) complex PhB(
i Pr2 Im)3 Fe≡N reacts with boron hydrides to afford PhB(i Pr2 Im)3 FeN( B )H ( B = 9-BBN ( 1 ), Bpin ( 2 )) and with (Bpin)2 to afford PhB(i Pr2 Im)3 FeN(Bpin)2 ( 3 ). The iron(II) borylamido products have all been structurally and spectroscopically characterized, demonstrating facile insertion into B-H and B-B bonds by PhB(i Pr2 Im)3 Fe≡N. Density functional theory (DFT) calculations reveal that the quintet state ( S = 2) is significantly lower in energy than the singlet ( S = 0) and triplet ( S = 1) states for all products. Stoichiometric reaction with (Bpin)2 does not produce the mono-borylated iron imido species PhB(i Pr2 Im)3 FeN(Bpin). DFT calculations suggest that this is because PhB(i Pr2 Im)3 FeN(Bpin) is unstable toward disproportionation to the starting iron(IV) nitride and PhB(i Pr2 Im)3 FeN(Bpin)2 . Attempts at B-C bond insertion using phenyl- and benzyl-pinacol borane were unsuccessful, which we attribute to unfavorable kinetics.- Published
- 2022
- Full Text
- View/download PDF
9. An Integrated View of Nitrogen Oxyanion Deoxygenation in Solution Chemistry and Electrospray Ion Production.
- Author
-
Beagan DM, Cabelof AC, Pepin R, Pink M, Carta V, and Caulton KG
- Abstract
There has been an increasing interest in chemistry involving nitrogen oxyanions, largely due to the environmental hazards associated with increased concentrations of these anions leading to eutrophication and aquatic "dead zones". Herein, we report the synthesis and characterization of a suite of MNO
x complexes (M = Co, Zn: x = 2, 3). Reductive deoxygenation of cobalt bis(nitrite) complexes with bis(boryl)pyrazine is faster for cobalt than previously reported nickel, and pendant O-bound nitrito ligand is still readily deoxygenated, despite potential implication of an isonitrosyl primary product. Deoxygenation of zinc oxyanion complexes is also facile, despite zinc being unable to stabilize a nitrosyl ligand, with liberation of nitric oxide and nitrous oxide, indicating N-N bond formation. X-ray photoelectron spectroscopy is effective for discriminating the types of nitrogen in these molecules. ESI mass spectrometry of a suite of M(NOx )y ( x = 2, 3 and y = 1, 2) shows that the primary form of ionization is loss of an oxyanion ligand, which can be alleviated via the addition of tetrabutylammonium (TBA) as a nonintuitive cation pair for the neutral oxyanion complexes. We have shown these complexes to be subject to deoxygenation, and there is evidence for nitrogen oxyanion reduction in several cases in the ESI plume. The attractive force between cation and neutral is explored experimentally and computationally and attributed to hydrogen bonding of the nitrogen oxyanion ligands with ammonium α-CH2 protons. One example of ESI-induced reductive dimerization is mimicked by bulk solution synthesis, and that product is characterized by X-ray diffraction to contain two Co(NO)2 + groups linked by a highly conjugated diazapolyene.- Published
- 2021
- Full Text
- View/download PDF
10. Iron(II) Complexes of an Anionic Bis(ylide)diphenylborate Ligand.
- Author
-
Gao Y, Pink M, and Smith JM
- Abstract
Double deprotonation of the salt [Ph
2 B(PMe3 )2 ][OTf] ( 1 ) provides access to a bis(ylide)diphenylborate ligand that is readily transferred in situ to iron(II). Depending on the reaction stoichiometry, both the "ate" complex [Ph2 B(Me2 PCH2 )2 Fe(μ-Cl)2 Li(THF)2 ] ( 2 ) and the homoleptic complex [Ph2 B(Me2 PCH2 )2 ]2 Fe( 3 ) can be prepared from FeCl2 (THF)1.5 . Further reaction of 3 with FeCl2 (THF)1.5 produces the chloride-bridged dimer [Ph2 B(Me2 PCH2 )2 Fe(μ-Cl)2 Fe(CH2 PMe2 )2 BPh2 ]( 4 ). Attempts to reduce or alkylate 4 provide 3 as the only isolable product, likely a consequence of the low steric hindrance of the bis(ylide)diphenylborate ligand. On the other hand, reaction of 4 with the strong field ligand CNt Bu provides the six-coordinate, diamagnetic complex [Ph2 B(Me2 PCH2 )2 Fe(CNt Bu)4 ][Cl]( 5 ). Electronic structure calculations for the bis(ylide)diphenylborate ligand and homoleptic complex 3 suggest that the C(ylide) atoms are strong σ-donors with little π-bonding character. These initial results suggest the potential for this bis(ylide)diphenylborate ligand in coordination chemistry.- Published
- 2020
- Full Text
- View/download PDF
11. Anion Control of Lanthanoenediyne Cyclization.
- Author
-
Kirschner KM, Ratvasky SC, Pink M, and Zaleski JM
- Abstract
A suite of lanthanoenediyne complexes of the form Ln(macrocycle)X
3 (Ln = La3+ , Ce3+ , Eu3+ , Gd3+ , Tb3+ , Lu3+ ; X = NO3 - , Cl- , OTf- ) was prepared by utilizing an enediyne-containing [2 + 2] hexaaza-macrocycle ( 2 ). The solid-state Bergman cyclization temperatures, measured via DSC, decrease with the denticity of X (bidentate NO3 - , T = 267-292 °C; monodentate Cl- , T = 238-262 °C; noncoordinating OTf- , T = 170-183 °C).13 C NMR characterization shows that the chemical shifts of the acetylenic carbon atoms also rely on the anion identity. The alkyne carbon closest to the metal binding site, CA , exhibits a Δδ > 3 ppm downfield shift, while the more distal alkyne carbon, CB , displays a concomitant Δδ ≤ 2.5 ppm upfield shift, reflecting a depolarization of the alkyne on metal inclusion. For all metals studied, the degree of perturbation follows the trend 2 < NO3 - < Cl- < OTf- . This belies a greater degree of electronic rearrangement in the coordinated macrocycle as the denticity of X and its accompanying shielding of the metal's Lewis acidity decrease. Computationally modeled structures of LnX3 show a systematic increase in the lanthanide- 2 coordination number (CNLa-mc = 2 (NO3 - ), 4 (Cl- ), 5 (H2 O, model for OTf- )) and a decrease in the mean Ln-N bond length (La-Naverage = 2.91 Å (NO3 - ), 2.78 Å (Cl- ), 2.68 Å (H2 O)), further suggesting that a decrease in the anion coordination number correlates with an increase in the metal-macrocycle interaction. Taken together, these data illustrate a Bergman cyclization landscape that is influenced by the bonding of metal to an enediyne ligand but whose reaction barrier is ultimately dominated by the coordinating ability of the accompanying anion.- Published
- 2019
- Full Text
- View/download PDF
12. Seeking Redox Activity in a Tetrazinyl Pincer Ligand: Installing Zerovalent Cr and Mo.
- Author
-
Maciulis NA, Schaugaard RN, Losovyj Y, Chen CH, Pink M, and Caulton KG
- Abstract
Reaction of the readily reduced pincer ligand bis-tetrazinylpyridine, btzp, with the zerovalent metal source M(CO)
3 (MeCN)3 yields M(btzp)2 for M = Cr, Mo. These diamagnetic molecules show intrapincer bond lengths consistent with major charge transfer from metal to ligand, a result which is further supported by X-ray photoelectron spectroscopy. These molecules show up to five reversible outer-sphere electron transfers by cyclic voltammetry. The electronic structure of neutral M(btzp)2 is analyzed by DFT and CASSCF calculations, which reveal the degree of back-donation from the metal into pincer π* orbitals and also subtle differences in metal-ligand interaction for Mo vs Cr. Near-IR absorptions exhibited by both M(btzp)2 species originate from charge transfer among differently reduced tetrazine rings, which thus further support pincer reduction in these species.- Published
- 2018
- Full Text
- View/download PDF
13. Electron and Oxygen Atom Transfer Chemistry of Co(II) in a Proton Responsive, Redox Active Ligand Environment.
- Author
-
Cook BJ, Pink M, Pal K, and Caulton KG
- Abstract
The bis-pyrazolato pyridine complex LCo(PEt
3 )2 serves as a masked form of three-coordinate CoII and shows diverse reactivity in its reaction with several potential outer sphere oxidants and oxygen atom transfer reagents. N-Methylmorpholine N-oxide (NMO) oxidizes coordinated PEt3 from LCo(PEt3 )2 , but the final cobalt product is still divalent cobalt, in LCo(NMO)2 . The thermodynamics of a variety of oxygen atom transfer reagents, including NMO, are calculated by density functional theory, to rank their oxidizing power. Oxidation of LCo(PEt3 )2 with AgOTf in the presence of LiCl as a trapping nucleophile forms the unusual aggregate [LCo(PEt3 )2 Cl(LiOTf)2 ]2 held together by Li+ binding to very nucleophilic chloride on Co(III) and triflate binding to those Li+ . In contrast, Cp2 Fe+ effects oxidation to trivalent cobalt, to form (HL)Co(PEt3 )2 Cl+ ; proton and the chloride originate from solvent in a rare example of CH2 Cl2 dehydrochlorination. An unexpected noncomplementary redox reaction is reported involving attack by 2e reductant PEt3 nucleophile on carbon of the 1e oxidant radical Cp2 Fe+ , forming a P-C bond and H+ ; this reaction competes in the reaction of LCo(PEt3 )2 with Cp2 Fe+ .- Published
- 2018
- Full Text
- View/download PDF
14. Ligand Substituent Effects in Manganese Pyridinophane Complexes: Implications for Oxygen-Evolving Catalysis.
- Author
-
Xu S, Bucinsky L, Breza M, Krzystek J, Chen CH, Pink M, Telser J, and Smith JM
- Abstract
A series of Mn(II) complexes of differently substituted pyridinophane ligands, (Py
2 NR2 )MnCl2 (R =i Pr, Cy) and [(Py2 NR2 )MnF2 ](PF6 ) (R =i Pr, Cy,t Bu) are synthesized and characterized. The electrochemical properties of these complexes are investigated by cyclic voltammetry, along with those of previously reported (Py2 NMe2 )MnCl2 and the Mn(III) complex [(Py2 NMe2 )MnF2 ](PF6 ). The electronic structure of this and other Mn(III) complexes is probed experimentally and theoretically, via high-frequency and -field electron paramagnetic resonance (HFEPR) spectroscopy ab initio quantum chemical theory (QCT), respectively. These studies show that the complexes contain relatively typical six-coordinate Mn(III). The catalytic activity of these complexes toward both H2 O2 disproportionation and H2 O oxidation has also been investigated. The rate of H2 O2 disproportionation decreases with increasing substituent size. Some of these complexes are active for electrocatalytic H2 O oxidation; however this activity cannot be rationalized in terms of simple electronic or steric effects.- Published
- 2017
- Full Text
- View/download PDF
15. Partial nitrogen atom transfer: a new synthetic tool to design single-molecule magnets.
- Author
-
Ding M, Rouzières M, Losovyj Y, Pink M, Clérac R, and Smith JM
- Abstract
Incomplete nitrogen atom transfer from the iron(IV) nitride complex PhB(MesIm)3Fe≡N to the vanadium(III) complex V(Mes)3(THF) quantitatively provides the bimetallic complex PhB(MesIm)3Fe-N═V(Mes)3. Structural and spectroscopic characterizations reveal that the nitride ligand forms a linear bridge between V(V) and high-spin Fe(II) metal ions, confirming that atom transfer is accompanied by electron transfer. In the presence of an applied dc field, the complex displays slow relaxation of the magnetization, revealing its single-molecule magnet properties with an estimation of the energy barrier at about 10 K. This complex establishes a synthetic principle for the assembly of paramagnetic complexes bridged by nitride ligands.
- Published
- 2015
- Full Text
- View/download PDF
16. Addition of Si-H and B-H bonds and redox reactivity involving low-coordinate nitrido-vanadium complexes.
- Author
-
Thompson R, Tran BL, Ghosh S, Chen CH, Pink M, Gao X, Carroll PJ, Baik MH, and Mindiola DJ
- Abstract
In this study we enumerate the reactivity for two molecular vanadium nitrido complexes of [(nacnac)V≡N(X)] formulation [nacnac = (Ar)NC(Me)CHC(Me)(Ar)(-), Ar = 2,6-(CHMe2)2C6H3); X(-) = OAr (1) and N(4-Me-C6H4)2 (Ntolyl2) (2)]. Density functional theory calculations and reactivity studies indicate the nitride motif to have nucleophilic character, but where the nitrogen atom can serve as a conduit for electron transfer, thus allowing the reduction of the vanadium(V) metal ion with concurrent oxidation of the incoming substrate. Silane, H2SiPh2, readily converts the nitride ligand in 1 into a primary silyl-amide functionality with concomitant two-electron reduction at the vanadium center to form the complex [(nacnac)V{N(H)SiHPh2}(OAr)] (3). Likewise, addition of the B-H bond in pinacolborane to the nitride moiety in 2 results in formation of the boryl-amide complex [(nacnac)V{N(H)B(pinacol)}(Ntolyl2)] (4). In addition to spectroscopic data, complexes 3 and 4 were also elucidated structurally by single-crystal X-ray diffraction analysis. One-electron reduction of 1 with 0.5% Na/Hg on a preparative scale allowed for the isolation and structural determination of an asymmetric bimolecular nitride radical anion complex having formula [Na]2[(nacnac)V(N)(OAr)]2 (5), in addition to room-temperature solution X-band electron paramagnetic resonance spectroscopic studies.
- Published
- 2015
- Full Text
- View/download PDF
17. Probing the steric and electronic characteristics of a new bis-pyrrolide pincer ligand.
- Author
-
Komine N, Buell RW, Chen CH, Hui AK, Pink M, and Caulton KG
- Abstract
A new pincer ligand is synthesized to be dianionic, with the potential to be redox active. It has pyrrrole rings attached to both ortho sites of a pyridine, as the linking element. This H2L can be doubly deprotonated and then used to replace two chloride ligands in MCl2(NCPh)2, to form LM(NCPh) for M = Pd, Pt. The acid form H2L reacts with ZnEt2 with elimination of only 1 mol of ethane to yield (HL)ZnEt, a three-coordinate species with one pendant pyrrole NH functionality. This molecule binds the Lewis base p-dimethylaminopyridine (DMAP) to give first a simple 1:1 adduct that eliminates ethane on heating to form four-coordinate LZn(DMAP), which has an unusual structure due to the strong preference of the pincer ligand to bind in a mer (planar) geometry. A molecule with two HL(-) ligands each bonded in a bidentate manner to FeCl2 is synthesized and shown to contain four-coordinate iron with a flattened-tetrahedral structure. The electrochemistry of LM(NCPh) and (L)Zn(DMAP) shows three oxidation processes, which is interpreted to involve at least two oxidations of the pyrrolide arms.
- Published
- 2014
- Full Text
- View/download PDF
18. 3d early transition metal complexes supported by a new sterically demanding aryloxide ligand.
- Author
-
Searles K, Tran BL, Pink M, Chen CH, and Mindiola DJ
- Abstract
The bulky aryloxide 2,6-bis(diphenylmethyl)-4-tert-butylphenol [HOAr(tBu)] (1) can be synthesized from 4-tert-butylphenol and benzhydrol in solvent-free conditions and obtained pure in 91% yield. Deprotonation of HOAr(tBu) is accomplished with M(N(SiMe3)2) (M = Na, Li), yielding the corresponding salts of the aryloxide [MOAr(tBu)] (M(+) = Na (2), Li(3)) in 83% and 73% yield, respectively. Facile salt formation of the aryloxide ligand allows for transmetalation to a variety of metal halides. Through transmetalation reactions involving two aryloxides, mononuclear complexes of the type [M'(OAr(tBu))2Cl(THF)2] (M' = Sc (4), V (5), Cr (6), Ti (7)) can be prepared from the corresponding metal halide precursor MCl3(THF)3. Additionally, two aryloxides can be coordinated to Ti(IV) via a protonolysis route of Ti(NMe2)2Cl2 and 2 equiv of HOAr(tBu) to yield [Ti(OAr(tBu))2Cl2(NHMe2)] (8) in 72% isolated yield. Single-crystal X-ray diffraction studies of 1, 2, and the 3d metal complexes 5-8 clearly show the steric demand of the bulky ligand, whereas in transition metal complexes we do not observe the formation of mononuclear tris-aryloxide complexes.
- Published
- 2013
- Full Text
- View/download PDF
19. Synthesis and oxidative reactivity of 2,2'-pyridylpyrrolide complexes of Ni(II).
- Author
-
Tsvetkov NP, Chen CH, Andino JG, Lord RL, Pink M, Buell RW, and Caulton KG
- Subjects
- Models, Molecular, Molecular Structure, Oxidation-Reduction, Quantum Theory, Nickel chemistry, Organometallic Compounds chemical synthesis, Organometallic Compounds chemistry, Pyrroles chemistry
- Abstract
Synthesis and characterization of divalent nickel complexed by 2-pyridylpyrrolide bidentate ligands are reported, as possible precursors to complexes with redox active ligands. Varied substituents on the pyrrolide, two CF3 (L(2)), two (t)Bu (L(0)), and one of each type (L(1)) are employed and the resulting Ni(L(n))2 complexes show different Lewis acidity toward CO, H2O, tetrahydrofuran (THF), or MeCN, the L(2) case being the most acidic. Density functional theory calculations show that the frontier orbitals of all three Ni(L(n))2 species are localized at the pyrrolide groups of both ligands and Ni(L(n))2(+) can be detected by mass spectrometry and in cyclic voltammograms (CVs). Following cyclic voltammetry studies, which show electroactivity primarily in the oxidative direction, reactions with pyridine N-oxide or Br2 are reported. The former yield simple bis adducts, Ni(L(2))2(pyNO)2 and the latter effects electrophilic aromatic substitution of the one pyrrolide ring hydrogen for both chelates, leaving it brominated.
- Published
- 2013
- Full Text
- View/download PDF
20. 2,2'-Pyridylpyrrolide ligand redistribution following reduction.
- Author
-
Searles K, Das AK, Buell RW, Pink M, Chen CH, Pal K, Morgan DG, Mindiola DJ, and Caulton KG
- Abstract
The potential redox activity of the 2,2'-pyridylpyrrolide ligand carrying two CF3 substituents (L(2)) is investigated. Synthesis and characterization of d(6) and d(7) species M(L(2))2 for M = Fe and Co are described (both are nonplanar, but not tetrahedral), as are the Lewis acidity of each. In spite of CV evidence for quasireversible reductions to form M(L(2))2(q-) where q = 1 and 2, chemical reductants instead yield divalent metal complexes KM(L(2))3, which show attractive interactions of K(+) to pyrrolide, to F, and to lattice toluene π cloud. The collected evidence on these products indicates that pyridylpyrrolide is a weak field ligand here, but CO can force spin pairing in Fe(L(2))2(CO)2. Evidence is presented that the overall reductive reaction yields 33 mol % of bulk metal, which is the fate of the reducing equivalents, and a mechanism for this ligand redistribution is proposed. Analogous ligand redistribution behavior is also seen for nickel and for trimeric monovalent copper analogues; reduction of Cu(L(2))2 simply forms Cu(L(2))2(-).
- Published
- 2013
- Full Text
- View/download PDF
21. Assessment of the electronic structure of 2,2'-pyridylpyrrolides as ligands.
- Author
-
Flores JA, Andino JG, Tsvetkov NP, Pink M, Wolfe RJ, Head AR, Lichtenberger DL, Massa J, and Caulton KG
- Subjects
- Electrons, Ligands, Models, Molecular, Molecular Structure, Organometallic Compounds chemical synthesis, Oxidation-Reduction, Pyrroles chemical synthesis, Quantum Theory, Organometallic Compounds chemistry, Pyrroles chemistry
- Abstract
The ligand class 2,2'-pyridylpyrrolide is surveyed, both for its structural features and its electronic structure, when attached to monovalent K, Cu, Ag, Au, and Rh. The influence of pyrrolide ring substituents is studied, as well as the question of push/pull interaction between the pyridyl and pyrrolide halves. The π donor ability of the pyrrolide is found to be less than that of an analogous phenyl. However, in contrast to the phenyl analog, the HOMO is pyrrolide π in character for pyridylpyrrolide complexes of copper and rhodium, while it is conventionally metal localized for planar, d(8) rhodium pyridylphenyl. Monovalent three-coordinate copper complexes show great deviations from Y-shaped toward T-shaped structures, including cases where the pyridyl ligand bonds only weakly.
- Published
- 2011
- Full Text
- View/download PDF
22. Modulating the light switch by (3)MLCT-(3)ππ* state interconversion.
- Author
-
Spencer BR, Kraft BJ, Hughes CG, Pink M, and Zaleski JM
- Subjects
- Crystallography, X-Ray, Luminescent Measurements methods, Magnetic Resonance Spectroscopy, Quantum Theory, Spectrophotometry, Ultraviolet, Spectroscopy, Fourier Transform Infrared, Spectrum Analysis, Raman, Coordination Complexes chemistry, DNA chemistry, Intercalating Agents chemistry, Pyrazines chemistry, Ruthenium chemistry
- Abstract
The spectroscopic, electronic, and DNA-binding characteristics of two novel ruthenium complexes based on the dialkynyl ligands 2,3-bis(phenylethynyl)-1,4,8,9-tetraaza-triphenylene (bptt, 1) and 2,3-bis(4-tert-butyl-phenylethynyl)-1,4,8,9-tetraaza-triphenylene (tbptt, 2) have been investigated. Electronic structure calculations of bptt reveal that the frontier molecular orbitals are localized on the pyrazine-dialkynyl portion of the free ligand, a property that is reflected in a red shift of the lowest energy electronic transition (1: λ(max) = 393 nm) upon substitution at the terminal phenyl groups (2: λ(max) = 398 nm). Upon coordination to ruthenium, the low-energy ligand-centered transitions of 1 and 2 are retained, and metal-to-ligand charge transfer transitions (MLCT) centered at λ(max) = 450 nm are observed for [Ru(phen)(2)bptt](2+)(3) and [Ru(phen)(2)tbptt](2+)(4). The photophysical characteristics of 3 and 4 in ethanol closely parallel those observed for [Ru(bpy)(3)](2+) and [Ru(phen)(3)](2+), indicating that the MLCT excited state is primarily localized within the [Ru(phen)(3)](2+) manifold of 3 and 4, and is only sparingly affected by the extended conjugation of the bptt framework. In an aqueous environment, 3 and 4 possess notably small luminescence quantum yields (3: ϕ(H(2)O) = 0.005, 4: ϕ(H(2)O) = 0.011) and biexponential decay kinetics (3: τ(1) = 40 ns, τ(2) = 230 ns; 4: τ(1) ∼ 26 ns, τ(2) = 150 ns). Addition of CT-DNA to an aqueous solution of 3 causes a significant increase in the luminescence quantum yield (ϕ(DNA) = 0.045), while the quantum yield of 4 is relatively unaffected (ϕ(DNA) = 0.013). The differential behavior demonstrates that tert-butyl substitution on the terminal phenyl groups inhibits the ability of 4 to intercalate with DNA. Such changes in intrinsic luminescence demonstrate that 3 binds to DNA via intercalation (K(b) = 3.3 × 10(4) M(-1)). The origin of this light switch behavior involves two competing (3)MLCT states similar to that of the extensively studied light switch molecule [Ru(phen)(2)dppz](2+). The solvent- and temperature-dependence of the luminescence of 3 reveal that the extended ligand aromaticity lowers the energy of the (3)ππ* excited state into competition with the emitting (3)MLCT state. Interconversion between these two states plays a significant role in the observed photophysics and is responsible for the dual emission in aqueous environments.
- Published
- 2010
- Full Text
- View/download PDF
23. Structure modulated electronic contributions to metalloenediyne reactivity: synthesis and thermal Bergman cyclization of MLX2 compounds.
- Author
-
Bhattacharyya S, Clark AE, Pink M, and Zaleski JM
- Subjects
- Crystallography, X-Ray, Cyclization, Electrons, Enediynes chemical synthesis, Molecular Conformation, Temperature, Copper chemistry, Enediynes chemistry, Palladium chemistry, Zinc chemistry
- Abstract
The synthesis of novel metalloendiyne complexes MLRX(2) (where L = 1,4-dibenzyl/diethyl-1,4-diaza-cyclododec-8-ene-6,10-diyne, X = halogen) are reported with their X-ray crystal structures and thermal Bergman cyclization temperatures. Two distinct types of constructs are obtained; the Zn(II) compounds are tetrahedral, while the Cu(II) and the Pd(II) compounds are all distorted- or square-planar. Each possesses structurally similar enediyne conformations and critical distances (3.75-3.88 A). The tetragonal Cu(II) species all exhibit Bergman cyclization temperatures between 140 and 150 degrees C in the solid state, while the square-planar Pd(II) analogues possess similar critical distances but cyclize at significantly higher temperatures (205-220 degrees C). In contrast, the Zn(II) derivatives show a marked halogen dependence, with X = Cl having the highest Bergman cyclization temperature, which is comparable to the Pd(II) square-planar set, while the ZnLX(2) compound with X = I shows the lowest Bergman cyclization temperature (144 degrees C), similar to the Cu(II) derivatives. Moreover, for the planar constructs, the R group has little influence on the cyclization temperatures; however, for the tetrahedral ZnLX(2) compounds, the steric influence of the R group plays a more significant role in the cyclization reaction coordinate by influencing the stability of the precyclized intermediate. This complex set of results is best interpreted by a combination of steric contributions and electronic interactions between the halogen through space (in the case of Zn(II)) and through bonds (in the case of Pd(II)) and the pi orbitals of the endiyne fragment. In contrast, for Cu(II) systems, the distorted square-planar geometry permits neither direct through space nor symmetry-allowed through bond communication between the orbital partners, and thus little variation in Bergman cyclization reactivity is observed.
- Published
- 2009
- Full Text
- View/download PDF
24. Influence of the metal orbital occupancy and principal quantum number on organoazide (RN3) conversion to transition-metal imide complexes.
- Author
-
Walstrom AN, Fullmer BC, Fan H, Pink M, Buschhorn DT, and Caulton KG
- Abstract
The reaction of phenyl azide with (PNP)Ni, where PNP = ( (t)Bu 2PCH 2SiMe 2) 2N (-), promptly evolves N 2 and forms a P=N bond in the product (PNP=NPh)Ni (I). A similar reaction with (PNP)FeCl proceeds to form a P=N bond but without N 2 evolution, to furnish (PNP=N-N=NPh)FeCl. An analogous reaction with (PNP)RuCl occurs with a more dramatic redox change at the metal (and N 2 evolution), to give the salt composed of (PNP)Ru(NPh) (+) and (PNP)RuCl 3 (-), together with equimolar (PNP)Ru(NPh). The contrast among these results is used to deduce what conditions favor N 2 loss and oxidative incorporation of the NPh fragment from PhN 3 into a metal complex.
- Published
- 2008
- Full Text
- View/download PDF
25. Nitrogen-ligated iron complexes: photolytic approach to the FeN+ moiety.
- Author
-
Buschhorn D, Pink M, Fan H, and Caulton KG
- Abstract
The synthesis of (PNP)FeCl, (PNP)Fe[NH(xylyl)], and (PNP)FeN3 are reported(PNP = (tBu2PCH2SiMe2)2N-). While the azide is thermally stable, it is photosensitive to lose N2 and form [(PNPN)Fe]2,in which the nitride ligand has formed a double bond to one phosphorus, and this N bridges to a second iron to form a 2-fold symmetric dimer. The reaction energy to form the (undetected) monomeric [eta3- tBu2PCH2SiMe2NSiMe2CH2PtBu2N]Fe is -15.9 kcal/mol, so this PIII --> PV oxidation is favorable. The eta2 version of this same species is less stable by 23.7 kcal/mol, which shows that the loss of one P--> Fe bond is caused by dimerization, and therefore, it does not precede and cause dimerization. A comparison is made to Ru analogs.
- Published
- 2008
- Full Text
- View/download PDF
26. P=N bond formation via incomplete N-atom transfer from a ferrous amide precursor.
- Author
-
Adhikari D, Basuli F, Fan H, Huffman JC, Pink M, and Mindiola DJ
- Subjects
- Amides chemistry, Ferrous Compounds chemistry, Nitrogen chemistry, Phosphorus chemistry
- Abstract
Incomplete N-atom transfer from Fe to P is observed when the ferrous amide complex (PNP)Fe(dbabh) (PNP-=N[2-P(iPr)2-4-methylphenyl]2, dbabh=2,3:5,6-dibenzo-7-azabicyclo[2.2.1]hepta-2,5-diene), prepared from salt metathesis of (PNP)FeCl and Li(dbabh), is thermolyzed at 70 degrees C over 48 h in C6D6. Several plausible reaction pathways resulting from the transformation of (PNP)Fe(dbabh) are discussed, including the possibility of an Fe(IV) nitride as an intermediate.
- Published
- 2008
- Full Text
- View/download PDF
27. The effect of one valence electron: contrasting (PNP)Ni(CO) with (PNP)Ni(NO) to understand the half-bent NiNO unit.
- Author
-
Fullmer BC, Pink M, Fan H, Yang X, Baik MH, and Caulton KG
- Abstract
Reaction of a (PNP)Ni radical with NO finishes in the time of mixing to form a 1:1 adduct with a NO stretching frequency of 1654 cm (-1). NMR data of this diamagnetic product indicate C 2 v symmetry, which is contradicted by the X-ray structure, which shows it to be nonplanar at Ni, with a geometry intermediate between planar and tetrahedral; the planar geometry is thus the transition state for fluxionality giving time-averaged C 2 v symmetry. The X-ray structure, together with DFT calculations, reveals that the "half-bent" NiNO unit and the intermediate coordination geometry result from a Ni --> NO charge transfer, which has a nonintegral value, resulting in a continuum between NO (+) (hence Ni (0)) and NO (-) (hence Ni (II)). This is related to the nonaxially symmetric character of the Ni --> NO back-donation caused by the (PNP) environment on Ni. Steric effects of ( t )Bu and even chelate constraints are ruled out as the cause of the unusual electronic and structural features.
- Published
- 2008
- Full Text
- View/download PDF
28. O/C bond cleavage of CO2 by NiI.
- Author
-
Fullmer BC, Fan H, Pink M, and Caulton KG
- Abstract
Reaction of (PNP)Ni, where PNP is [((t)Bu2PCH2SiMe2)2N](-1), with CO2 occurs rapidly even at -60 degrees C to form exclusively the product of transposition of the amide N and one CO2 oxygen: [((t)Bu2PCH2SiMe2)2O]Ni(NCO). DFT(B3LYP) evaluation of several candidate intermediates for breaking two Si/N and one C/O bond and forming two Si/O and one N/C bond reveal species at and below the energy of the separated particles, and establish the location of the spin densities in each.
- Published
- 2008
- Full Text
- View/download PDF
29. Influence of the d-Electron Count on CO binding by three-coordinate [(tBu2PCH2SiMe2)2N]Fe, -Co, and -Ni.
- Author
-
Ingleson MJ, Fullmer BC, Buschhorn DT, Fan H, Pink M, Huffman JC, and Caulton KG
- Abstract
Reduction of (PNP)MCl [PNP = ((t)Bu(2)PCH(2)SiMe(2))(2)N] with Mg gives three-coordinate, T-shaped (PNP)M for M = Fe(S = 3/2) and Ni. Their reactivity was tested toward CO; Ni binds one CO, but only reversibly (i.e., CO is completely lost in vacuum), and has a CO stretching frequency showing effective back-donation by NiI. The structure of (PNP)Ni(CO) is intermediate between planar and tetrahedral, in contrast to the planar d8 analogue, (PNP)Co(CO). This structural reorganization on carbonylation changes the singly occupied molecular orbital from having negligible phosphorus character [no P hyperfine structure in the electron paramagnetic resonance (EPR) spectrum of (PNP)Ni] to having enough P character to have a triplet structure in the EPR spectrum of the CO. The presence of one fewer electron in (PNP)Fe (vs the Co analogue) leads to binding of two CO, and (PNP)Fe(CO)(2) is characterized as a spin doublet with square-pyramidal structure. Density functional theory calculations strengthen the understanding of the structural and spectroscopic changes along this dn series (n = 7-9).
- Published
- 2008
- Full Text
- View/download PDF
30. Exploring the reactivity of four-coordinate PNPCoX with access to three-coordinate spin triplet PNPCo.
- Author
-
Ingleson MJ, Pink M, Fan H, and Caulton KG
- Abstract
The compounds (PNP)CoX, where PNP is (tBu2PCH2SiMe2)2N- and X is Cl, I, N3, OAr, OSO2CF3 and N(H)Ar, are reported. Some of these show magnetic susceptibility, color, and 1H NMR evidence of being in equilibrium between a blue, tetrahedral S=3/2 state and a red, planar S=1/2 state; the equilibrium populations are influenced by subtle solvent effects (e.g., benzene and cyclohexane are different), as well as by temperature. Attempted oxidation to Co(III) with O2 occurs instead at phosphorus, giving [P(O)NP(O)]CoX species. The single O-atom transfer reagent PhI=O likewise oxidizes P. Even I2 oxidizes P to give the pendant phosphonium species (tBu2P(I)CH2SiMe2NSiMe2CH2PtBu2)CoI2 with a tetrahedral S=3/2 cobalt; the solid-state structure shows intermolecular PI...ICo interactions. Attempted alkyl metathesis of PNPCoX inevitably results in reduction, forming PNPCo, which is a spin triplet with planar T-shaped coordination geometry with no agostic interaction. Triplet PNPCo binds N2(weakly) and CO (whose low CO stretching frequency indicates strong PNP-->Co donor power), but not ethene or MeCCMe.
- Published
- 2007
- Full Text
- View/download PDF
31. Radical (NO) and nonradical (N(2)O) reagents convert a ruthenium(IV) nitride to the same nitrosyl complex.
- Author
-
Walstrom A, Pink M, Fan H, Tomaszewski J, and Caulton KG
- Abstract
The ruthenium(IV) nitride complex (PNP)RuN (PNP = (tBu2PCH2-SiMe2)2N-) reacts rapidly with 2NO to form (PNP)Ru(NO) and N2O, via no detectable intermediate. The linear nitrosyl complex has a planar structure. In a slower reaction, (PNP)RuN reacts with N2O by O-atom transfer (established by 15N labeling) to give the same nitrosyl complex and N2. Density functional theory (B3LYP) calculations show both reactions to be very thermodynamically favorable. Analysis of possible intermediates in each reaction shows that radical (PNP)RuN(NO) has much spin density on nitride N (hence, N2-), while one 2 + 3 metallacycle, (PNP)RuN3O, has the wrong connectivity to form a product. Instead, an intermediate with a doubly bent N2O (hence, a two-electron reduced N-nitrosoimide form) brings the O atom in proximity to the nitride N on the path to a product.
- Published
- 2007
- Full Text
- View/download PDF
32. Influence of chelate substituents on the structure and spin state of unsaturated [N(SiMe2CH2PtBu2)2]Ru-X.
- Author
-
Yang X, Walstrom A, Tsvetkov N, Pink M, and Caulton KG
- Abstract
Density functional theory calculations on the conformational preferences in the two fused five-membered rings of anionic N(SiR2CH2PR'2)2 chelated to RuX+ are compared to several experimental structures (X=halide). The calculations consider the structures of both singlet and triplet states and reveal that both the four tBu groups and the crowded juncture (N(SiMe2)2) of the two rings must be included computationally to understand the observed structures. Computational experiments with different substituents R and R' show the reality of N-->Ru pi donation. The cases where X=H and CH3 are also studied.
- Published
- 2007
- Full Text
- View/download PDF
33. Si-N bond hydrolysis furnishes a planar 4-coordinate 14-electron Ru(II) complex with a triplet ground state.
- Author
-
Walstrom A, Pink M, and Caulton KG
- Abstract
Reaction of stoichiometric (2:1) water with [(tBu2PCH2SiMe2)2N]Ru(OSO2CF3) produces planar, 14-valence-electron spin triplet trans-Ru(tBu2PCH2SiMe2O)2. A possible mechanism for this hydrolysis is discussed. This molecule reacts rapidly with CO to give a monocarbonyl, then a cis-dicarbonyl. Reaction with HCCR (R = H or Ph) yields the vinylidene (tBu2PCH2SiMe2O)2Ru=C=CHR.
- Published
- 2006
- Full Text
- View/download PDF
34. Coordination chemistry and insulin-enhancing behavior of vanadium complexes with maltol C6H6O3 structural isomers.
- Author
-
Saatchi K, Thompson KH, Patrick BO, Pink M, Yuen VG, McNeill JH, and Orvig C
- Subjects
- Animals, Blood Glucose analysis, Blood Glucose metabolism, Crystallography, X-Ray, Diabetes Mellitus, Experimental drug therapy, Diabetes Mellitus, Experimental metabolism, Hydrogen-Ion Concentration, Hypoglycemic Agents therapeutic use, Isomerism, Ligands, Pyrones pharmacology, Pyrones therapeutic use, Rats, Time Factors, Vanadates pharmacology, Vanadates therapeutic use, Vanadium Compounds pharmacology, Vanadium Compounds therapeutic use, Hypoglycemic Agents pharmacology, Insulin metabolism, Pyrones chemistry, Vanadium Compounds chemistry
- Abstract
Syntheses of vanadium complexes using the naturally occurring ligands isomaltol (Hima) and allomaltol (Hama), as well as a newly synthesized, potentially tetradentate diaminodipyrone [H(2)(en(ama)(2)], are reported. Complete characterization of the resulting compounds [trans-VO(ima)(2)(H(2)O), VO(ama)(2), V(ima)(3), V(ama)(3) and VO(en(ama)(2))], including X-ray crystallography analyses for trans-VO(ima)(2)(H(2)O) and V(ima)(3), are presented herein. Potentiometric titrations (25 degrees C, I = 0.16 M NaCl) were used to measure stability constants in the V(IV)-Hima system; these data were compared to previous data collected on the V(IV)-L (L = Hma, Hama) systems. The in vivo efficacy of these compounds to lower the blood glucose levels of STZ-diabetic rats was tested; all but VO(en(ama)(2)) produced significant decreases in plasma glucose levels. The results were compared to those of the benchmark compound BMOV [VO(ma)(2), bis(maltolato)oxovanadium(IV)], a known insulin-enhancing agent.
- Published
- 2005
- Full Text
- View/download PDF
35. Large spin differences in structurally related Fe6 molecular clusters and their magnetostructural explanation.
- Author
-
Cañada-Vilalta C, O'Brien TA, Brechin EK, Pink M, Davidson ER, and Christou G
- Subjects
- Crystallography, X-Ray, Models, Chemical, Temperature, Iron chemistry, Magnetics, Organometallic Compounds chemical synthesis, Organometallic Compounds chemistry
- Abstract
The syntheses, crystal structures, and magnetic characterizations of three new hexanuclear iron(III) compounds are reported. Known [Fe(6)O(2)(OH)(2)(O(2)CBu(t))(10)(hep)(2)] (1) is converted to new [Fe(6)O(2)(OH)(O(2)CBu(t))(9)(hep)(4)] (3) when treated with an excess of 2-(2-hydroxyethyl)-pyridine (hepH). Similarly, the new compound [Fe(6)O(2)(OH)(2)(O(2)CPh)(10)(hep)(2)] (2), obtained from the reaction of [Fe(3)O(O(2)CPh)(6)(H(2)O)(3)] with hepH, is converted to [Fe(6)O(2)(OH)(O(2)CPh)(9)(hep)(4)] (4) when treated with an excess of hepH. This can be reversed by recrystallization from MeCN. The cores of the four Fe(6) complexes all comprise two triangular [Fe(3)(mu(3)-O)(O(2)CR)(3)(hep)](+3) units connected at two of their apices by two sets of bridging ligands. However, 1 and 2 differ slightly from 3 and 4 in the precise way the two Fe(3) units are linked together. In 1 and 2, the two sets of bridging ligands are identical, consisting of one mu-hydroxo and two mu-carboxylate groups bridging each Fe(2) pair, i.e., a (mu-OH(-))(mu-O(2)CR(-))(2) set. In contrast, 3 and 4 have two different sets of bridging ligands, a (mu-OH(-))(mu-O(2)CR(-))(2) set as in 1 and 2, and a (mu-OR(-))(2)(mu-O(2)CR(-)) set, where RO(-) refers to the alkoxide arm of the hep(-) chelate. Variable-field and -temperature dc magnetization measurements establish that 1 and 2 have S = 5 ground states and significant and positive zero-field splitting parameters (D), whereas 3 and 4 have S = 0 ground states. This dramatic difference of 10 unpaired electrons in the ground state S values for near-isomeric compounds demonstrates an acute sensitivity of the magnetic properties to small structural changes. The factors leading to this have been quantitatively analyzed. The semiempirical method ZILSH, based on unrestricted molecular orbital calculations, was used to obtain initial estimates of the Fe(2) pairwise exchange interaction constants (J). These calculated values were then improved by fitting the experimental susceptibility versus T data, using a genetic algorithm approach. The final J values were then employed to rationalize the observed magnetic properties as a function of the core topologies and the presence of spin frustration effects. The large difference in ground state spin value was identified as resulting from a single structural difference between the two types of complexes, the different relative dispositions (cis vs trans) of two frustrated exchange pathways. In addition, use of the structural information and corresponding J values allowed a magnetostructural correlation to be established between the J values and both the Fe-O bond distances and the Fe-O-Fe angles at the bridging ligands.
- Published
- 2004
- Full Text
- View/download PDF
36. Hydrocarbyl ligand "tuning" of the PtII/IV redox potential.
- Author
-
Vedernikov AN, Pink M, and Caulton KG
- Abstract
The potentially tridentate macrocycle [2.1.1]-(2,6)-pyridinophane (L) enables the transient LPt(II)(CH(3))(+) to cleave the C-H bond of two molecules of C(6)F(5)H. The resulting product has two aryl groups on Pt but, in contrast to nonfluorinated analogue, varies in its location of the cleaved H, as is evident from the two products (HL(+))Pt(II)R(2) and (eta(3)-L)Pt(IV)H(R)(2)(+). At equilibrium, the related example where R = CH(3) is purely the Pt(IV) redox isomer, which with R = C(6)H(5) shows detectable populations of both isomers, and with R = C(6)F(5) is purely the pyridine-protonated (HL(+))Pt(II) redox isomer. All species show a hydrogen bond from the pyridinium proton to Pt(II). Consistent with the idea that electron-withdrawing R makes platinum(II) more resistant to oxidation (i.e., a proton on Pt), and thus less Brønsted basic, the (1)J(PtH) coupling constant falls in the series R = Me (90 Hz), R = C(6)H(5) (86 Hz), and R = C(6)F(5) (63 Hz).
- Published
- 2004
- Full Text
- View/download PDF
37. Reactivity of the hydrido/nitrosyl radical MHCl(NO)(CO)(P(i)Pr(3))(2), M = Ru, Os.
- Author
-
Marchenko AV, Vedernikov AN, Dye DF, Pink M, Zaleski JM, and Caulton KG
- Abstract
The reaction of equimolar NO with the 16 electron molecule RuHCl(CO)L(2) (L = P(i)Pr(3)) proceeds, via a radical adduct RuHCl(CO)(NO) L(2), onward to form RuCl(NO)(CO)L(2) (X-ray structure determination) and RuHCl(HNO)(CO)L(2), in a 1:1 mole ratio. The HNO ligand, bound by N and trans to hydride, is rapidly degraded by excess NO. The osmium complex behaves analogously, but the adduct has a higher formation constant, permitting determination of its IR spectrum; both MHCl(CO)(NO)L(2) radicals are characterized by EPR spectroscopy, and DFT calculations on the Ru system show it to have a "half-bent" Ru-N-O unit with the spin density mainly on nitrogen. DFT (PBE) energies rule out certain possible mechanistic steps for forming the two products. A survey of the literature leads to the hypothesis that NO should generally be considered as a (neutral) Lewis base (2-electron donor) when it binds to a 16 electron complex which is resistant to oxidation or reduction, and that the resulting N-centered radical has a M-N-O angle of approximately 140 degrees, which distinguishes it from NO(-) (bent at <140 degrees ) and from NO(+) (>170 degrees ).
- Published
- 2004
- Full Text
- View/download PDF
38. Methanolysis and phenolysis routes to Fe6, Fe8, and Fe1) complexes and their magnetic properties: a new type of Fe8 ferric wheel.
- Author
-
Cañada-Vilalta C, O'Brien TA, Pink M, Davidson ER, and Christou G
- Abstract
Alcoholysis of preformed tetranuclear and hexanuclear iron(III) clusters has been employed for the synthesis of four higher-nuclearity clusters. Treatment of [Fe(4)O(2)(O(2)CMe)(7)(bpy)(2)](ClO(4)) with phenol affords the hexanuclear cluster [Fe(6)O(3)(O(2)CMe)(9)(OPh)(2)(bpy)(2)](ClO(4)) (1). Reaction of [Fe(6)O(2)(OH)(2)(O(2)CR)(10)(hep)(2)] (R = Bu(t) or Ph) with PhOH affords the new "ferric wheel" complexes [Fe(8)(OH)(4)(OPh)(8)(O(2)CR)(12)] [R = Bu(t) (2) or Ph (3)]. Complexes 2 and 3 exhibit the same structure, which is an unprecedented type for Fe(III). In contrast, treatment of [Fe(6)O(2)(OH)(2)(O(2)CBu(t))(10)(hep)(2)] with MeOH leads to the formation of [Fe(10)(OMe)(20)(O(2)CBu(t))(10)] (4), which exhibits the more common type of ferric wheel seen in analogous complexes with other carboxylate groups. Solid-state variable-temperature magnetic susceptibility measurements indicate spin-singlet ground states for complexes 2 and 4. The recently developed semiempirical method ZILSH was used to estimate the pairwise exchange parameters (J(AB)) and the average spin couplings S(A)[empty set].S(B)[empty set] between the Fe(III) centers, providing a clear depiction of the overall magnetic behavior of the molecules. All exchange interactions between adjacent Fe(III) atoms are antiferromagnetic.
- Published
- 2003
- Full Text
- View/download PDF
39. Dlf complexes with uniform coordination geometry: structural and magnetic properties of an LnNi2 core supported by a heptadentate amine phenol ligand.
- Author
-
Bayly SR, Xu Z, Patrick BO, Rettig SJ, Pink M, Thompson RC, and Orvig C
- Abstract
The synthesis and physical characterization of a series of lanthanide (Ln(III)) and nickel (Ni(II)) mixed trimetallic complexes with the heptadentate (N(4)O(3)) amine phenol ligand H(3)trn [tris(2'-hydroxybenzylaminoethyl)amine] has been accomplished in order to extend our understanding of how amine phenol ligands can be used to coaggregate d- and f-block metal ions and to investigate further the magnetic interaction between these ions. The one-pot reaction in methanol of stoichiometric amounts of H(3)trn with NiX(2).6H(2)O (X = ClO(4), NO(3)) followed by addition of the corresponding LnX(3).6H(2)O salt, and then base, produces complexes of the general formula [LnNi(2)(trn)(2)]X.nH(2)O. The complexes were characterized by a variety of analytical techniques. Crystals of five of the complexes were grown from methanol solutions and their structures were determined by X-ray analysis: [PrNi(2)(trn)(2)(CH(3)OH)]ClO(4).4CH(3)OH.H(2)O, [SmNi(2)(trn)(2)(CH(3)OH)]NO(3).4CH(3)OH.2H(2)O, [TbNi(2)(trn)(2)(CH(3)OH)]NO(3).4CH(3)OH.3H(2)O, [ErNi(2)(trn)(2)(CH(3)OH)]NO(3).6CH(3)OH, and [LuNi(2)(trn)(2)(CH(3)OH)]NO(3).4.5CH(3)OH.1.5H(2)O. The [LnNi(2)(trn)(2)(CH(3)OH)](+) complex cation consists of two octahedral Ni(II) ions, each of which is encapsulated by the ligand trn(3)(-) in an N(4)O(2) coordination sphere with one phenolate O atom not bound to Ni(II). Each [Ni(trn)](-) unit acts as a tridentate ligand toward the Ln(III) ion via two bridging and one nonbridging phenolate donors. Remarkably, in all of the structurally characterized complexes, Ln(III) is seven-coordinate and has a flattened pentagonal bipyramidal geometry. Such uniform coordination behavior along the whole lanthanide series is rare and can perhaps be attributed to a mismatch between the geometric requirements of the bridging and nonbridging phenolate donors. Magnetic studies indicate that ferromagnetic exchange occurs in the Ni(II)/Ln(II) complexes where Ln = Gd, Tb, Dy, Ho, or Er.
- Published
- 2003
- Full Text
- View/download PDF
40. An electron-excessive nitrosyl complex: reactivity of a ligand-centered radical leading to coordinated HNO.
- Author
-
Marchenko AV, Vedernikov AN, Dye DF, Pink M, Zaleski JM, and Caulton KG
- Abstract
The reaction of RuHCl(CO)L(2) (L = P(i)Pr(3)) with NO initially forms a 1:1 adduct, shown by DFT calculations and EPR spectroscopy (including the RuD isotopomer) to contain a bent ( 90 degree angle Ru-N-O = 143.9 degrees ) nitrosyl where the majority of the spin density is on the nitrosyl nitrogen. This radical adduct transforms further to give equimolar RuCl(NO)(CO)L(2) and RuHCl(HNO)(CO)L(2), the latter with hydride trans to the nitroxyl ligand HN=O. This is the first observation of the synthesis of coordinated HNO from NO itself. DFT calculations lead to the proposal that this H-atom transfer is effected by free NO, and the lifetime of RuHCl(HNO)(CO)L(2) is indeed qualitatively dependent on the presence of free NO.
- Published
- 2002
- Full Text
- View/download PDF
41. Linear trimer analogues of calixarene as chiral coordinating ligands: X-ray crystallographic and NMR spectroscopic characterization of chiral and achiral trisphenolates complexed to titanium(IV) and aluminum(III).
- Author
-
Appiah WO, DeGreeff AD, Razidlo GL, Spessard SJ, Pink M, Young VG Jr, and Hofmeister GE
- Abstract
Achiral and chiral linear trisphenol analogues of calixarene (HOArCH(2)Ar'(OH)C(R)HArOH, Ar = 4,6-di-tert-butylphenyl; Ar' = 4-tert-butylphenyl; R = H (achiral), Me (chiral)) were prepared in anticipation of their adoption of a chiral conformation upon coordination to Lewis acidic metal centers. The trisphenols react with 1 equiv of Ti(OR')(4) (R' = i-Pr or t-Bu) to yield complexes with molecular formula Ti(2)(OArCH(2)Ar'(O)C(R)HArO)(2)(OR')(2) (R = H, Me; R' = i-Pr or t-Bu). An X-ray crystal structure of the titanium complex of the achiral trisphenol (R = H; R' = t-Bu) reveals that the trisphenolate ligand adopts an unsymmetrical (and therefore chiral) conformation, with eta(2)-coordination to one metal center and eta(1)-coordination to the second metal center. The chiral trisphenol, which contains a stereogenic center (indicated as C in the shorthand notation used above), coordinates titanium in an analogous fashion to produce only one diastereomer (out of four possible); therefore, the configuration of the stereogenic center controls the conformation adopted by the bound ligand. The reaction of achiral trisphenol with AlMe(3) produces a compound with molecular formula Al(2)(OArCH(2)Ar'(O)CH(2)ArO)(2). (1)H NMR spectroscopy and X-ray crystallography reveal that the trisphenolate ligand adopts an asymmetric, C(2) conformation in this complex, where the central phenolate oxygen bridges the aluminum centers and the terminal phenolate oxygens each coordinate a separate aluminum center. Because these trisphenolate ligands adopt chiral conformations when coordinated to metal centers, they may be useful for developing diastereo- or enantioselective catalysts and reagents.
- Published
- 2002
- Full Text
- View/download PDF
42. Tristhiolatomolybdenum nitrides, (RS)(3)Mo[triple bond]N where R = (i)Pr and (t)Bu, preparation, characterization and comparisons with related trialkoxymolybdenumnitrides.
- Author
-
Chisholm MH, Davidson ER, Pink M, and Quinlan KB
- Abstract
The addition of thiols to ((t)BuO)(3)Mo[triple bond]N in toluene leads to the formation of (RS)(3)Mo[triple bond]N compounds as yellow, air-sensitive compounds, where R = (i)Pr and (t)Bu. The single-crystal structure of ((t)BuS)(3)Mo[triple bond]N reveals a weakly associated dimeric structure where two ((t)BuS)(3)Mo[triple bond]N units (Mo-N = 1.61 A, Mo-S = 2.31 A (av)) are linked via thiolate sulfur bridges with long 3.03 A (av) Mo-S interactions. Density functional theory calculations employing Gaussian 98 B3LYP (LANL2DZ for Mo and 6-31G* for N, O, S, and H) have been carried out for model compounds (HE)(3)Mo[triple bond]N and (HE)(3)MoNO, where E = O and S. A comparison of the structure and bonding within the related series ((t)BuE)(3)Mo[triple bond]N and ((t)BuE)(3)MoNO is made for E = O and S. In the thiolate compounds, the highest energy orbitals are sulfur lone-pair combinations. In the alkoxides, the HOMO is the N 2p lone-pair which has M-N sigma and M-O pi* character for the nitride. As a result of greater O p pi to Mo pi interactions, the M-N pi orbitals of the Mo-N triple bond are destabilized with respect to their thiolate counterpart. For the nitrosyl compounds, the greater O p pi to Mo d pi interaction favors greater back-bonding to the nitrosyl pi* orbitals for the alkoxides relative to the thiolates. The results of the calculations are correlated with the observed structural features and spectroscopic properties of the related alkoxide and thiolate compounds.
- Published
- 2002
- Full Text
- View/download PDF
43. Three-coordinate copper(II)-phenolate complexes.
- Author
-
Jazdzewski BA, Holland PL, Pink M, Young VG Jr, Spencer DJ, and Tolman WB
- Subjects
- Catalysis, Chemical Phenomena, Chemistry, Physical, Crystallography, X-Ray, Electrochemistry, Electron Spin Resonance Spectroscopy, Ligands, Lithium chemistry, Magnetic Resonance Spectroscopy, Molecular Conformation, Molecular Structure, Organometallic Compounds chemistry, Spectrophotometry, Ultraviolet, Spectrum Analysis, Raman, Copper chemistry, Organometallic Compounds chemical synthesis, Phenols chemistry
- Abstract
The reactions of LCuCl (L = 2,4-bis((2,6-diisopropylphenyl)imido)pentane (L(iPr)), 2,4-bis((2,6-diisopropylphenyl)imido)-3-chloropentane (L(CliPr))) with the phenolates TlOAr (Ar = C(6)H(3)Me(2), C(6)H(4)OMe, C(6)H(4)tBu) and NaOC(6)H(3)(tBu)(2) were explored. Novel three-coordinate Cu(II)-phenolates, LCuOAr, were isolated from the reactions with the thallium phenolates and were characterized by X-ray crystallography and spectroscopy (UV-vis, EPR). The complexes feature short Cu-O(phenolate) distances (average Cu-O = 1.81 A) and, with one exception, irregular N-Cu-O(phenolate) angles that differ within each compound (15 degrees < Delta < 28 degrees, where Delta = angleN(1)-Cu-O - angleN(2)-Cu-O). The exception is L(iPr)Cu(OC(6)H(4)tBu), for which X-ray structures at -100 and 25 degrees C differed due to an unusual reversible phase change with nonmerohedral twinning (2:1 ratio) in the low-temperature form. The high-temperature form has local C(2)(v) symmetry (Delta = 0 degrees ), and upon cooling below the phase transition temperature (-8 +/- 5 degrees C) lateral movement of the phenolate ligand (Delta = 17.6 degrees ) and rotation of the phenolate plane by 10.7 degrees occurs. Resonance Raman spectroscopic data acquired for L(iPr)Cu(OC(6)H(4)tBu) corroborated assignment of phenolate --> Cu(II) LMCT character in the UV-vis spectra. Cyclic voltammetry experiments (THF, 0.5 M NBu(4)PF(6)) revealed negative E(1/2) values for the Cu(II)/Cu(I) couples relative to NHE, consistent with enhanced stabilization of the Cu(II) state by both the strongly electron donating beta-diketiminate ligand and the phenolates. Although thermally stable, the Cu(II)-phenolates are unusually reactive with dioxygen, albeit to give product(s) that have yet to be identified. In the reaction of L(iPr)CuCl with NaOC(6)H(3)(tBu)(2) no Cu(II)-phenolate was observed. Instead, a Cu(I) complex was generated quantitatively by trapping with added isocyanide, [L(iPr)CuNC(C(6)H(3)Me(2))], along with 3,3',5,5'-tetra-tert-butyl-4,4'-dibenzoquinone and 2,6-di-tert-butylphenol in 27 +/- 3% and 46 +/- 6% yields, respectively, corresponding to the overall reaction 4L(iPr)Cu(II)Cl + 4NaOAr --> 4L(iPr)Cu(I) + 4NaCl + dibenzoquinone + 2(phenol).
- Published
- 2001
- Full Text
- View/download PDF
44. Macrocyclic metalloenediynes of Cu(II) and Zn(II): a thermal reactivity comparison.
- Author
-
Chandra T, Pink M, and Zaleski JM
- Subjects
- Chemical Phenomena, Chemistry, Physical, Crystallography, X-Ray, Cyclization, Diamines chemistry, Heterocyclic Compounds chemical synthesis, Magnetic Resonance Spectroscopy, Models, Chemical, Molecular Conformation, Molecular Structure, Organometallic Compounds chemical synthesis, Spectrometry, Mass, Electrospray Ionization, Spectroscopy, Fourier Transform Infrared, Copper chemistry, Heterocyclic Compounds chemistry, Organometallic Compounds chemistry, Peroxides chemistry, Zinc chemistry
- Abstract
The syntheses of tetradentate enediyne macrocycles with 24 (tact1:1)-, 26 (tact1:2)-, and 28 (tact2:2)-membered rings are described, along with their thermal reactivities and those of the corresponding Cu(II) (Cu(tact1:1), Cu(tact1:2)) and Zn(II) (Zn(tact1:1), Zn(tact1:2)) complexes. These enediyne macrocyclic ligands are not benzannulated and thus exhibit thermal Bergman cyclization temperatures near 200 degrees C by differential scanning calorimetry (DSC). Moreover, the synthetic route allows incorporation of additional carbon atoms into the macrocycles which increases their conformational flexibilities and lowers their Bergman cyclization temperatures. Specifically, as the size of the macrocycle increases, the temperatures at which these compounds undergo Bergman cyclization decrease by approximately 5 degrees C per additional carbon atom, leading to an overall decrease across the series of 19 degrees C. Incorporation of Cu(II) and Zn(II) into these macrocycles further reduces their cyclization temperatures relative to those of the free ligands. More uniquely, for Cu(tact1:1) and Zn(tact1:1), the observed cyclization temperatures vary by 27 degrees C with the Zn(II) complex lying to higher temperature (Cu(tact1:1) = 121 degrees C, (Zn(tact1:1) = 148 degrees C). As the macrocycle size is increased, the decrease in the Bergman cyclization temperatures observed for the free ligands does not systematically hold for the Cu(II) and Zn(II) derivatives. Rather, the Cu(II) complex exhibits the expected 9 degrees C decrease in the cyclization temperature (Cu(tact1:2) = 112 degrees C), whereas the temperature for the Zn(II) analogue increases by 15 degrees C (Zn(tact1:2) = 163 degrees C). From the X-ray crystal structure of the free ligand and the geometric structural preferences of the electronic configurations of Cu(II) and Zn(II), the higher cyclization temperatures for the Zn(II) complex with the larger ring size can be explained by a distortion of the macrocycle toward a more tetrahedral metal center geometry.
- Published
- 2001
- Full Text
- View/download PDF
45. Synthesis, isolation, and characterization of trisodium tricarbonyliridate (3-), Na3[Ir(CO)3]. Initial studies on its derivative chemistry and structural characterizations of trans-[Ir(CO)3(EPh3)2](-), E = Ge, Sn, and trans-[Co(CO)3(SnPh3)2](-1).
- Author
-
Allen JM, Brennessel WW, Buss CE, Ellis JE, Minyaev ME, Pink M, Warnock GF, Winzenburg ML, and Young VG Jr
- Abstract
Reduction of Na[Ir(CO)4] by sodium metal in (Me2N)3PO, followed by treatment with liquid ammonia, provided high yields (ca. 90%) of unsolvated Na3[Ir(CO)3], a thermally stable, pyrophoric orange solid. This substance contains iridium in its lowest known formal oxidation state of -3 and has been characterized by IR spectroscopy, elemental analyses, and derivative chemistry, i.e., by its conversion to the triphenylgermyl and triphenylstannyl complexes, trans-[Ir(CO)3(EPh3)2](-), E = Ge, Sn. Single-crystal X-ray structures of the tetraethylammonium salts of these species, as well as [Co(CO)3(SnPh3)2](-), confirm the trigonal bipyramidal nature of the anions, originally predicted on the basis of their IR spectra in the carbonyl stretching frequency region. These structural characterizations provide important additional evidence for the presence of metal tricarbonyl units in Na3[M(CO)3], M = Co, Ir.
- Published
- 2001
- Full Text
- View/download PDF
46. Benzo[2,1-c:3,4-c']bis(1,2,3-thiaselenazole) (BSe) and its charge trasfer chemistry. Crystal and electronic structure of [BSe]3[ClO4]2.
- Author
-
Beer L, Britten JF, Cordes AW, Clements OP, Oakley RT, Pink M, and Reed RW
- Abstract
The S-Se-N-based heterocycle benzo[2,1-c:3,4-c']bis(1,2,3-thiaselenazole) (BSe) can be prepared by the condensation of 1,4-diaminobenzene-2,3-dithiol with selenium tetrachloride. Crystals of this compound are not isomorphous with the related benzo[2,1-c:3,4-c']bis(1,2,3-dithiazole) (BT); a structure is adopted that allows for more extensive intermolecular Se- - -Se contacts. Electro-oxidation of BSe in the presence of [n-Bu4N][ClO4] affords metallic green needles of the charge transfer salt [BSe]3[ClO4]2, which exhibit a pressed pellet conductivity sigma(RT) = 10(-1) S cm(-1). The crystal structure of [BSe]3[ClO4]2 consists of slipped pi-stacks based on the triple-decker closed shell [BSe]3(2+) building block. The packing is analogous to that found for the charge transfer salt [BT]3[FSO3]2, for which sigma(RT) = 10(-2) S cm(-1). Extended Hückel band structure calculations on these two (sulfur- and selenium-based) 3:2 salts reveal more extensive intermolecular interactions in the selenium compound. As a result, the latter has a more two-dimensional electronic structure. Crystal data for Se2S2N2C6H2, a = 4.103(2) A, b = 12.159(2) A, c = 16.171(2) A, orthorhombic, Pbnm, Z = 4. Crystal data for Se6S6N6C18H6Cl2O4, a =17.00(1) A, b = 18.36(1) A, c = 10.679(4) A, 110.27(3), monoclinic, C2/c, Z = 4.
- Published
- 2001
- Full Text
- View/download PDF
47. Variable-temperature x-ray structural investigation of [Fe[HC(3,5-Me2pz)3]2](BF4)2 (pz= pyrazolyl ring): observation of a thermally induced spin state change from all high spin to an equal high spin-low spin mixture, concomitant with the onset of nonmerohedral twinning.
- Author
-
Reger DL, Little CA, Young VG Jr, and Pink M
- Abstract
The complex [Fe[HC(3,5-Me(2)pz)(3)](2)](BF(4))(2) (pz = pyrazolyl ring) undergoes a phase transition that occurs concomitantly with a thermally induced spin conversion between the high-spin (HS, S = 2) and low-spin (LS, S = 0) states. Above 204 K the compound is completely HS with the structure in the C2/c space group with Z = 4. A crystal structure determination of this phase was performed at 220 K yielding the cell constants a = 20.338(2) A, b = 10.332(1) A, c = 19.644(2) A, beta = 111.097(2) degrees, and V = 3851.5(6) A(3). There is one unique iron(II) site at this temperature. Below 206 K the compound converts to a 50:50 mixture of HS and LS. The radical change in the coordination sphere for half of the iron(II) sites, most notably a shortening of the Fe-N bond distances by ca. 0.2 A, that accompanies this magnetic transition causes a phase transition. The crystal system changes from C-centered monoclinic to primitive triclinic with Z = 2 with two half-molecules on independent inversion centers. A crystal structure determination was performed at 173 K in space group P1 with a = 10.287(2) A, b = 11.355(3) A, c = 18.949(4) A, alpha = 90.852(4) degrees, beta = 105.245(4) degrees, gamma = 116.304(4) degrees, and V = 1892.3(8) A(3). All specimens investigated below the phase transition temperature were determined to be nonmerohedral twins. Temperature cycling between these two forms does not appear to degrade crystal quality. Previous magnetic susceptibility measurements indicate a second, irreversible increase in the magnetic moment the first time the crystals are cooled below 85 K. A crystal structure determination at 220 K of a specimen precooled to 78 K was not significantly different from those not cooled below 220 K.
- Published
- 2001
- Full Text
- View/download PDF
48. Gallane complexes with amido-amine ligands.
- Author
-
Luo B, Pink M, and Gladfelter WL
- Abstract
Gallane complexes bearing amido-amine ligands -N(R)CH2CMe2CH2NMe2 [R = H or SiMe3 (TMS)], (H2Ga[N(H)CH2CMe2CH2NMe2])2, 1, H2Ga[N(TMS)CH2CMe2CH2NMe2], 2, (H(Cl)Ga[N(H)CH2CMe2CH2NMe2])2, 3, ([(TMS)2N](H)Ga[N(H)CH2CMe2CH2NMe2])2, 4, and HGa[N(TMS)CH2CMe2CH2NMe2]2, 5, were synthesized from the reactions of the quinuclidine adducts of mono- and dichlorogallane with the corresponding lithium amides. Structural determinations of compounds 1, 3, and 4 showed all were dimeric with bridging amido groups. Rather than bond to gallium the tertiary amine groups in 1 and 4 were hydrogen-bonded to the amino N-H. In the structure of compound 3 the amine group occupied an axial position in the trigonal bipyramidal geometry of the five-coordinate gallium. The results were rationalized in terms of the steric and electronic properties of gallium ligands.
- Published
- 2001
- Full Text
- View/download PDF
49. Seven-coordinate [ReVON4X2]+ complexes (X = O and Cl).
- Author
-
Xu L, Setyawati IA, Pierreroy J, Pink M, Young VG, Patrick BO, Rettig SJ, and Orvig C
- Subjects
- Crystallography, X-Ray, Models, Molecular, Molecular Conformation, Molecular Structure, Organometallic Compounds chemical synthesis, Radioisotopes, Radiopharmaceuticals, Structure-Activity Relationship, Organometallic Compounds chemistry, Rhenium
- Abstract
The oxorhenium(V) complexes with ligands containing N4 (H2pmen) and N4O2 (H2bbpen, H2Clbbpen, and H2bped) donor atom sets have been synthesized. X-ray crystallographic analyses of the [ReO(H2pmen)Cl2]+, [ReO(bbpen)]+, and [ReO(bped)]+ complexes showed that all three cations share a rare seven-coordinate structure with a distorted pentagonal bipyramidal geometry, which represents a novel and potentially general structural motif in ReV = O complexes. 1H NMR spectroscopy shows that the structures of the complexes are retained in the solution.
- Published
- 2000
- Full Text
- View/download PDF
50. The first noncoordinated phosphonium diylide, [Me2P(C13H8)2]-, and its ylidic and cationic counterparts: synthesis, structural characterization, and interaction with the heavy group 2 metals.
- Author
-
Brady ED, Hanusa TP, Pink M, and Young VG
- Abstract
Treatment of potassium or lithium fluorenide with MePCl2 generates the organophosphine MeP(C13H9)2, which on reaction with methyl iodide produces the phosphonium species [Me2P(C13H9)2]I in 74% yield. In the solid state, H...I contacts of < 3.3 A help generate a layered structure in which the fluorenyl rings are nearly parallel. On subsequent reaction of [Me2P(C13H9)2]I with either KH or K[N(SiMe3)2], the corresponding neutral phosphoylide, Me2P(C13H9)(C13H8), forms in 67% yield and was structurally characterized. The phosphonium iodide [Me2P(C13H9)2]I was allowed to react with Ae[N(SiMe3)2]2 (Ae = Ca, Ba), and the product from the reaction with the calcium complex was structurally identified as the salt [CaI(thf)5][Me2P(C13H8)2]. The anion, which is outside the coordination sphere of the calcium, represents the first structurally authenticated example of a free phosphonium diylide. The P-C(ylidic) bond length of 1.748(4) A reflects some partial multiple bond character. 1H and 31P NMR spectra suggest that the barium analogue is similar. Density functional theory calculations were performed on representative phosphonium diylides as an aid to interpreting the bonding in this class of compounds. Despite the strong electrostatic attraction that usually drives metal-ligand binding in highly ionic systems, calcium and barium prefer to coordinate to a single iodide ion and several neutral oxygen donors rather than to the charged diylide.
- Published
- 2000
- Full Text
- View/download PDF
Catalog
Discovery Service for Jio Institute Digital Library
For full access to our library's resources, please sign in.