Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 96
Filtrar
Mais filtros

Base de dados
Tipo de documento
Intervalo de ano de publicação
1.
Inorg Chem ; 62(30): 11751-11760, 2023 Jul 31.
Artigo em Inglês | MEDLINE | ID: mdl-37471624

RESUMO

Six new solvent-free, homoleptic paramagnetic tris(alkyl)lanthanides Ln{C(SiHMe2)3}3 (1Ln) and Ln{C(SiHMe2)2Ph}3 (2Ln) (Ln = Gd, Dy, and Er) were synthesized to investigate the magnetic properties of 4f organometallic compounds stabilized by secondary Ln↼H-Si and benzylic interactions. The unit cell of 1Gd contains one independent molecule (Z = 2), while 1Dy and 1Er crystallize with four independent isostructural molecules per unit cell (Z = 16). In all molecules, as in other 1Ln compounds, the three tris(dimethylsilyl)methyl ligands form a trigonal planar LnC3 core, and six secondary interactions involving Ln↼H-Si bonding in Ln{C(SiHMe2)3}3 form above and below the equatorial plane. Two and five crystallographically independent molecules of each 2Ln (2Gd, Z = 8; 2Dy, Z = 20) form with three π-coordinated phenyl groups in addition to either one or two secondary Ln↼H-Si interactions per molecule. The packing of these midseries organolanthanide compounds contrasts the single crystallographically unique molecules in previously reported La{C(SiHMe2)3}3 (1La, Z = 2, Z' = 1) and La{C(SiHMe2)2Ph}3 (2La, Z = 2, Z' = 1/3). 2La doped with 2Dy can adopt the crystallographic structure of 2La, which promotes magnetic properties, namely a higher χmT value at low temperatures as well as stronger magnetic anisotropy. The ac susceptibility data for 10% 2Dy doped into 2La suggests slow relaxation at low temperatures with a relaxation barrier of ∼45 K. The computed saturated magnetization of 1Er (M ≈ 4.5 µB) and 1Dy (M ≈ 6 µB) matches the experimental values, while the computed value for 2Dy better matches the value measured for 2Dy diluted in 2La (M ≈ 5 µB). Gas-phase calculations predict that the ground-state and first excited-state multiplet separations are larger for 1Er than 2Er, while the ordering for dysprosium is 1Dy > 2Dy.

2.
J Org Chem ; 87(2): 1507-1511, 2022 Jan 21.
Artigo em Inglês | MEDLINE | ID: mdl-34985897

RESUMO

Metastable radicals exist in a steady-state equilibrium in solution with dimers, which can be either σ dimers or π dimers. Here, we show that steric hindrance at the para position causes julolidine-derived dicyanomethyl radicals to form σ dimers rather than π dimers, the opposite behavior as seen in other carbon-centered radicals, where steric hindrance typically favors pimerization. The change in dimerization mode can be attributed to weaker London dispersion forces and a decreased orbital overlap in the sterically hindered dicyanomethyl radical π dimers, while the bulky groups exert relatively little effect on the energy of the σ dimer.

3.
Chemistry ; 27(40): 10428-10436, 2021 Jul 16.
Artigo em Inglês | MEDLINE | ID: mdl-33876468

RESUMO

Three-coordinate Ph BOX Me 2 ZnR (Ph BOX Me 2 =phenyl-(4,4-dimethyl-oxazolinato; R=Me: 2 a, Et: 2 b) catalyzes the dehydrocoupling of primary or secondary silanes and alcohols to give silyl ethers and hydrogen, with high turnover numbers (TON; up to 107 ) under solvent-free conditions. Primary and secondary silanes react with small, medium, and large alcohols to give various degrees of substitution, from mono- to tri-alkoxylation, whereas tri-substituted silanes do not react with MeOH under these conditions. The effect of coordinative unsaturation on the behavior of the Zn catalyst is revealed through a dramatic variation of both rate law and experimental rate constants, which depend on the concentrations of both the alcohol and hydrosilane reactants. That is, the catalyst adapts its mechanism to access the most facile and efficient conversion. In particular, either alcohol or hydrosilane binds to the open coordination site on the Ph BOX Me 2 ZnOR catalyst to form a Ph BOX Me 2 ZnOR(HOR) complex under one set of conditions or an unprecedented σ-adduct Ph BOX Me 2 ZnOR(H-SiR'3 ) under other conditions. Saturation kinetics provide evidence for the latter species, in support of the hypothesis that σ-bond metathesis reactions involving four-centered electrocyclic 2σ-2σ transition states are preceded by σ-adducts.

4.
Angew Chem Int Ed Engl ; 60(47): 25082-25088, 2021 11 15.
Artigo em Inglês | MEDLINE | ID: mdl-34528379

RESUMO

A new strategy to stabilize free radicals electronically is described by conjugating formally antiaromatic substituents to the free radical. With an antiaromatic substituent, the radical acts as an electron sink to allow configuration mixing of a low-energy zwitterionic state that provides antiaromaticity relief to the substituent. A combination of X-ray crystallography, VT-EPR and VT-UV/Vis spectroscopy, as well as computational analysis, was used to investigate this phenomenon. We find that this strategy of antiaromaticity relief is successful at stabilizing radicals, but only if the antiaromatic substituent is constrained to be planar by synthetically imposed conformational restraints that enable state mixing. This work leads to the counterintuitive finding that increasing the antiaromaticity of the radical substituent leads to greater radical stability, providing proof of concept for a new stereoelectronic approach for stabilizing free radicals.

5.
J Am Chem Soc ; 142(11): 5304-5313, 2020 03 18.
Artigo em Inglês | MEDLINE | ID: mdl-32088961

RESUMO

Some free radicals are stable enough to be isolated, but most are either unstable transient species or exist as metastable species in equilibrium with a dimeric form, usually a spin-paired sigma dimer or a pi dimer (pimer). To gain insight into the different modes of dimerization, we synthesized and evaluated a library of 15 aryl dicyanomethyl radicals in order to probe what structural and molecular parameters lead to σ- versus π-dimerization. We evaluated the divergent dimerization behavior by measuring the strength of each radical association by variable-temperature electron paramagnetic resonance spectroscopy, determining the mode of dimerization (σ- or π-dimer) by UV-vis spectroscopy and X-ray crystallography, and performing computational analysis. We evaluated three different hypotheses to explain the difference in the dimerization behavior: (1) that the dimerization behavior is dictated by radical spin densities; (2) that it is dictated by radical polarizability; (3) that it is dictated by London dispersion stabilization of the pimer. However, no single parameter model in itself was predictive. Two-parameter models incorporating either the computed degree of spin delocalization or the radical polarizability as well as computed estimates for the attractive London dispersion forces in the π-dimers lead to improved forecasts of σ- vs π-dimerization mode, and suggest that a balance of spin delocalization of the isolated radical as well as attractive forces between the stacked radicals, govern the formation of diradical pimers.

6.
Inorg Chem ; 58(6): 3815-3824, 2019 Mar 18.
Artigo em Inglês | MEDLINE | ID: mdl-30821972

RESUMO

The rhodium dicarbonyl {PhB(OxMe2)2ImMes}Rh(CO)2 (1) and primary silanes react by oxidative addition of a nonpolar Si-H bond and, uniquely, a thermal dissociation of CO. These reactions are reversible, and kinetic measurements model the approach to equilibrium. Thus, 1 and RSiH3 react by oxidative addition at room temperature in the dark, even in CO-saturated solutions. The oxidative addition reaction is first-order in both 1 and RSiH3, with rate constants for oxidative addition of PhSiH3 and PhSiD3 revealing kH/ kD ∼ 1. The reverse reaction, reductive elimination of Si-H from {PhB(OxMe2)2ImMes}RhH(SiH2R)CO (2), is also first-order in [2] and depends on [CO]. The equilibrium concentrations, determined over a 30 °C temperature range, provide Δ H ° = -5.5 ± 0.2 kcal/mol and Δ S ° = -16 ± 1 cal·mol-1K-1 (for 1 ⇄ 2). The rate laws and activation parameters for oxidative addition (Δ H⧧ = 11 ± 1 kcal·mol-1 and Δ S⧧ = -26 ± 3 cal·mol-1·K-1) and reductive elimination (Δ H⧧ = 17 ± 1 kcal·mol-1 and Δ S⧧ = -10 ± 3 cal·mol-1K-1), particularly the negative activation entropy for both forward and reverse reactions, suggest the transition state of the rate-determining step contains {PhB(OxMe2)2ImMes}Rh(CO)2 and RSiH3. Comparison of a series of primary silanes reveals that oxidative addition of arylsilanes is ca. 5× faster than alkylsilanes, whereas reductive elimination of Rh-Si/Rh-H from alkylsilyl and arylsilyl rhodium(III) occurs with similar rate constants. Thus, the equilibrium constant Ke for oxidative addition of arylsilanes is >1, whereas reductive elimination is favored for alkylsilanes.

7.
Inorg Chem ; 58(9): 6044-6051, 2019 May 06.
Artigo em Inglês | MEDLINE | ID: mdl-30995014

RESUMO

The reaction of FeBr2 and 1 equiv of thallium tris(4,4-dimethyl-2-oxazolinyl)phenylborate (TlToM) in THF provides ToMFeBr (1), whereas FeBr2 and 2 equiv of TlToM react to give (ToM)2Fe (2). Two νCN bands at 1604 and 1548 cm-1 indicated bidentate coordination of ToM to iron in 2. Homoleptic 2 and FeBr2 react in THF overnight through an unusual ligand exchange process to give compound 1, which is apparently the thermodynamic product. The salt metathesis reaction of 1 and KCH2Ph affords ToMFeCH2Ph (3). The effective magnetic moments of compounds 1-3 range from 4.9 to 5.4 µB, and these values are consistent with high-spin iron(II) ( S = 2). A single 1H NMR signal assigned to the methyl groups of the ToM ligand suggested tridentate coordination of the ToM ligand to iron in 1 and 3. X-ray crystallography studies of 1-3 establish their structure as four-coordinated tetrahedral iron complexes. ToMFeBn and CO (1 atm) react to afford isolable ToMFe{C(═O)Bn}(CO)2 (4) as a yellow solid. Complex 4 is diamagnetic ( S = 0), and the three distinct methyl signals in the 1H NMR spectrum are consistent with a six-coordinate, C s-symmetric species. This assignment is supported by its IR spectrum, which revealed intense bands at 2004 and 1935 cm-1 (symmetric and asymmetric νCO), at 1680 and 1662 cm-1 (acyl rotamers, νCO), and at 1593 and 1553 cm-1 (νCN) and is confirmed by a single-crystal X-ray diffraction study.

8.
J Am Chem Soc ; 140(43): 14308-14313, 2018 10 31.
Artigo em Inglês | MEDLINE | ID: mdl-30277768

RESUMO

Stable organic radicals with switchable spin states have attracted attention for a variety of applications, but a fundamental understanding of how radical structure effects the weak bonding interactions between organic radicals is limited. To evaluate the effect of chemical structure on the strength and nature of such spin interactions, a series of 14 tethered aryl dicyanomethyl diradicals were synthesized, and the structure and thermodynamic properties of the diradicals were investigated. These studies indicate that the nature of the dimer and the equilibrium thermodynamic parameters of the diradical-dimer equilibria are highly sensitive to the attachment point of the linker, the length of the linker, and the substituents on the radical itself. Values of the intramolecular Ka vary from as small as 5 to as high as 105 depending on these variables. An X-ray crystal structure for a linked ortho-substituted diradical shows that the diradical forms an intramolecular sigma dimer in the crystalline state with an elongated C-C bond (1.637 Å). Subtle changes to the radical structure influences the nature of the spin interactions, as fixing the dimethylamino substituent on the radical into a ring to make a julolidine-derived diradical leads to the weakest bonding interaction observed (Δ Gbonding = 1 kcal mol-1) and changes the spin-paired species from a sigma dimer to a diradical pimer. This work has implications for the design of stimuli-responsive materials that can reversibly switch between the dramatically different properties of closed-shell species and the unique properties of diradicals.

9.
J Org Chem ; 83(17): 9850-9857, 2018 09 07.
Artigo em Inglês | MEDLINE | ID: mdl-30024169

RESUMO

The delocalization of electron density upon tautomerization of a proton across a conjugated bridge can alter the strength of hydrogen bonds. This effect has been dubbed resonance-assisted hydrogen bonding (RAHB) and plays a major role in the energetics of the tautomeric equilibrium. The goal of this work was to investigate the role that π-delocalization plays in the stability of RAHBs by engaging other isomerization processes. Similarly, acid-base chemistry has received little experimental attention in studies of RAHB, and we address the role that acid-base effects play in the tautomeric equilibrium. We find that π-delocalization and the disruption of adjacent aromatic rings is the dominant effect in determining the stability of a RAHB.

10.
J Am Chem Soc ; 139(46): 16862-16874, 2017 11 22.
Artigo em Inglês | MEDLINE | ID: mdl-28991458

RESUMO

Homoleptic tris(alkyl) rare earth complexes Ln{C(SiHMe2)3}3 (Ln = La, 1a; Ce, 1b; Pr, 1c; Nd, 1d) are synthesized in high yield from LnI3THFn and 3 equiv of KC(SiHMe2)3. X-ray diffraction studies reveal 1a-d are isostructural, pseudo-C3-symmetric molecules that contain two secondary Ln↼HSi interactions per alkyl ligand (six total). Spectroscopic assignments are supported by comparison with Ln{C(SiDMe2)3}3 and DFT calculations. The Ln↼HSi and terminal SiH exchange rapidly on the NMR time scale at room temperature, but the two motifs are resolved at low temperature. Variable-temperature NMR studies provide activation parameters for the exchange process in 1a (ΔH⧧ = 8.2(4) kcal·mol-1; ΔS⧧ = -1(2) cal·mol-1K-1) and 1a-d9 (ΔH⧧ = 7.7(3) kcal·mol-1; ΔS⧧ = -4(2) cal·mol-1K-1). Comparisons of lineshapes, rate constants (kH/kD), and slopes of ln(k/T) vs 1/T plots for 1a and 1a-d9 reveal that an inverse isotope effect dominates at low temperature. DFT calculations identify four low-energy intermediates containing five ß-Si-H⇀Ln and one γ-C-H⇀Ln. The calculations also suggest the pathway for Ln↼HSi/SiH exchange involves rotation of a single C(SiHMe2)3 ligand that is coordinated to the Ln center through the Ln-C bond and one secondary interaction. These robust organometallic compounds persist in solution and in the solid state up to 80 °C, providing potential for their use in a range of synthetic applications. For example, reactions of Ln{C(SiHMe2)3}3 and ancillary proligands, such as bis-1,1-(4,4-dimethyl-2-oxazolinyl)ethane (HMeC(OxMe2)2) give {MeC(OxMe2)2}Ln{C(SiHMe2)3}2, and reactions with disilazanes provide solvent-free lanthanoid tris(disilazides).

11.
Angew Chem Int Ed Engl ; 56(2): 628-631, 2017 01 09.
Artigo em Inglês | MEDLINE | ID: mdl-27900844

RESUMO

The homoleptic organocerium complex Ce{C(SiHMe2 )3 }3 (1) reacts with B(C6 F5 )3 to produce the zwitterionic bis(alkyl) hydridoborato Ce{C(SiHMe2 )3 }2 HB(C6 F5 )3 (2). NMR and IR spectroscopy and X-ray crystallography indicate that each alkyl ligand contains two bridging Ce↼H-Si interactions in both 1 and 2. Compound 2 serves as a precatalyst for the hydrosilylation of acrylates to give α-silyl esters at room temperature with a turnover number of 2200.

12.
Chemistry ; 22(44): 15619-15623, 2016 Oct 24.
Artigo em Inglês | MEDLINE | ID: mdl-27572933

RESUMO

We report the first examples of transition metal-catalyzed enantioselective alkene hydroacylations with 1,1,2-trisubstituted alkenes. DFT and mechanistic studies are consistent with a reaction pathway for these rhodium-catalyzed processes including intramolecular alkene hydroacylation and α-epimerization to generate highly enantioenriched, polycyclic architectures. This reaction sequence enables the hydroacylation of 2-(cyclohex-1-en-1-yl)benzaldehydes to form hexahydro-9H-fluoren-9-ones in moderate to high yields (68-91 %) with high enantioselectivities (up to 99 % ee) and diastereoselectivities (typically >20:1).

13.
Chemistry ; 21(6): 2387-97, 2015 Feb 02.
Artigo em Inglês | MEDLINE | ID: mdl-25403795

RESUMO

Mixed-valence polyoxovanadates(IV/V) have emerged as one of the most intricate class of supramolecular all-inorganic host species, able to encapsulate a wide variety of smaller guest templates during their self-assembly formation process. As showcased herein, the incorporation of guests, though governed solely by ultra-weak electrostatic and van der Waals interactions, can cause drastic effects on the electronic and magnetic characteristics of the shell complex of the polyoxovanadate. We address the question of methodology for the magnetochemical analysis of virtually isostructural {V(IV/V) 22 O54 }-type polyoxoanions of D2d symmetry enclosing diamagnetic VO2 F2 (-) (C2v ), SCN(-) (C∞v ), or ClO4 (-) (Td ) template anions. These induce different polarization effects related to differences in their geometric structures, symmetry, ion radii, and valence shells, eventually resulting in a supramolecular modulation of magnetic exchange between the V(3d) electrons that are partly delocalized over the {V22 O54 } shells. We also include the synthesis and characterization of the novel [V(V) O2 F2 @HV(IV) 8 V(V) 14 O54 ](6-) system that comprises the rarely encountered discrete difluorovanadate anion as a quasi-isolated guest species.

14.
Inorg Chem ; 54(14): 6938-46, 2015 Jul 20.
Artigo em Inglês | MEDLINE | ID: mdl-26110352

RESUMO

{Bo(M)Cp(tet)}Lu(CH2Ph)2 (1; Bo(M)Cp(tet) = MeC(Ox(Me2))2C5Me4; Ox(Me2) = 4,4-dimethyl-2-oxazoline) was prepared in 95% yield from the reaction of Bo(M)Cp(tet)H and Lu(CH2Ph)3THF3. Compound 1 reacts with 1 or 2 equiv of H2NCH2R (R = C6H5, 1-C10H7) to give the corresponding imido complexes [{Bo(M)Cp(tet)}LuNCH2R]2 (R = C6H5 (2a), 1-C10H7 (2b)) or amido complexes {Bo(M)Cp(tet)}Lu(NHCH2R)2 (R = C6H5 (3a), 1-C10H7 (3b)). Once isolated, the imido species are insoluble in nonprotic organic solvents. Crystallographic characterization reveals dimeric [{Bo(M)Cp(tet)}LuNCH2(1-C10H7)]2 in the solid state. The reaction of 1 and NH3B(C6F5)3 affords crystallographically characterized {Bo(M)Cp(tet)}Lu{NHB(C6F5)2}C6F5. This species is proposed to form via a transient lutetium imido, which undergoes C6F5 migration to the lutetium center.

15.
Org Biomol Chem ; 13(36): 9462-70, 2015 Sep 28.
Artigo em Inglês | MEDLINE | ID: mdl-26248770

RESUMO

Bithiophene dicarboxaldehydes are promising electron-rich building blocks for the development of arylene vinylene-based organic semiconductors, but their use has been limited due to their synthetic inaccessibility. To facilitate the facile synthesis of these compounds we have prepared a novel functional bithiophene, namely 2,2'-(3,3'-dibromo-[2,2'-bithiophene]-5,5'-diyl)bis(5,5-dimethyl-1,3-dioxane) in two high yielding steps from 3,3',5,5'-tetrabromo-2,2'-bithiophene. This synthon is readily transformed into variety of bithiophene-based dicarboxaldehydes, also in high yields. The use of these functional molecules in the synthesis of arylene vinylene-linked donor-acceptor copolymers is demonstrated by the synthesis of two copolymers with electron deficient benzobisazoles.

16.
J Am Chem Soc ; 135(19): 7235-50, 2013 May 15.
Artigo em Inglês | MEDLINE | ID: mdl-23631736

RESUMO

Aminoalkenes are catalytically cyclized in the presence of cyclopentadienylbis(oxazolinyl)borato group 4 complexes {PhB(C5H4)(Ox(R))2}M(NMe2)2 (M = Ti, Zr, Hf; Ox(R) = 4,4-dimethyl-2-oxazoline, 4S-isopropyl-5,5-dimethyl-2-oxazoline, 4S-tert-butyl-2-oxazoline) at room temperature and below, affording five-, six-, and seven-membered N-heterocyclic amines with enantiomeric excesses of >90% in many cases and up to 99%. Mechanistic investigations of this highly selective system employed synthetic tests, kinetics, and stereochemistry. Secondary aminopentene cyclizations require a primary amine (1-2 equiv vs catalyst). Aminoalkenes are unchanged in the presence of a zirconium monoamido complex {PhB(C5H4)(Ox(4S-iPr,Me2))2}Zr(NMe2)Cl or a cyclopentadienylmono(oxazolinyl)borato zirconium diamide {Ph2B(C5H4)(Ox(4S-iPr,Me2))}Zr(NMe2)2. Plots of initial rate versus [substrate] show a rate dependence that evolves from first-order at low concentration to zero-order at high concentration, and this is consistent with a reversible substrate-catalyst interaction preceding an irreversible step. Primary kinetic isotope effects from substrate conversion measurements (k'obs((H))/k'obs((D)) = 3.3 ± 0.3) and from initial rate analysis (k2((H))/k2((D)) = 2.3 ± 0.4) indicate that a N-H bond is broken in the turnover-limiting and irreversible step of the catalytic cycle. Asymmetric hydroamination/cyclization of N-deutero-aminoalkenes provides products with higher optical purities than obtained with N-proteo-aminoalkenes. Transition state theory, applied to the rate constant k2 that characterizes the irreversible step, provides activation parameters consistent with a highly organized transition state (ΔS(++) = -43(7) cal·mol(-1) K(-1)) and a remarkably low enthalpic barrier (ΔH(++) = 6.7(2) kcal·mol(-1)). A six-centered, concerted transition state for C-N and C-H bond formation and N-H bond cleavage involving two amidoalkene ligands is proposed as most consistent with the current data.


Assuntos
Alcenos/química , Aminas/química , Zircônio/química , Aminação , Boranos/química , Boratos/química , Catálise , Ciclização , Modelos Moleculares , Oxazóis/química , Estereoisomerismo
17.
J Am Chem Soc ; 135(40): 15225-37, 2013 Oct 09.
Artigo em Inglês | MEDLINE | ID: mdl-24067161

RESUMO

The reactivity of a series of disilazido zirconocene complexes is dominated by the migration of anionic groups (hydrogen, alkyl, halide, OTf) between the zirconium and silicon centers. The direction of these migrations is controlled by the addition of two-electron donors (Lewis bases) or two-electron acceptors (Lewis acids). The cationic nonclassical [Cp2ZrN(SiHMe2)2](+) ([2](+)) is prepared from Cp2Zr{N(SiHMe2)2}H (1) and B(C6F5)3 or [Ph3C][B(C6F5)4], while reactions of B(C6F5)3 and Cp2Zr{N(SiHMe2)2}R (R = Me (3), Et (5), n-C3H7 (7), CH═CHSiMe3 (9)) provide a mixture of [2](+) and [Cp2ZrN(SiHMe2)(SiRMe2)](+). The latter products are formed through B(C6F5)3 abstraction of a ß-H and R group migration from Zr to the ß-Si center. Related ß-hydrogen abstraction and X group migration reactions are observed for Cp2Zr{N(SiHMe2)2}X (X = OTf (11), Cl (13), OMe (15), O-i-C3H7 (16)). Alternatively, addition of DMAP (DMAP = 4-(dimethylamino)pyridine) to [2](+) results in coordination to a Si center and hydrogen migration to zirconium, giving the cationic complex [Cp2Zr{N(SiHMe2)(SiMe2DMAP)}H](+) ([19](+)). Related hydrogen migration occurs from [Cp2ZrN(SiHMe2)(SiMe2OCHMe2)](+) ([18](+)) to give [Cp2Zr{N(SiMe2DMAP)(SiMe2OCHMe2)}H](+) ([22](+)), whereas X group migration is observed upon addition of DMAP to [Cp2ZrN(SiHMe2)(SiMe2X)](+) (X = OTf ([12](+)), Cl ([14](+))) to give [Cp2Zr{N(SiHMe2)(SiMe2DMAP)}X](+) (X = OTf ([26](+)), Cl ([20](+))). The species involved in these transformations are described by resonance structures that suggest ß-elimination. Notably, such pathways are previously unknown in early metal amide chemistry. Finally, these migrations facilitate direct Si-H addition to carbonyls, which is proposed to occur through a pathway that previously had been reserved for later transition metal compounds.

18.
Inorg Chem ; 52(8): 4154-6, 2013 Apr 15.
Artigo em Inglês | MEDLINE | ID: mdl-23534822

RESUMO

One-dimensional chain coordination polymers based on hexanuclear iron(III) pivalate building blocks and 1,4-dioxane (diox) or 4,4'-bipyridine (4,4'-bpy) bridging ligands, [Fe6O2(O2CH2)(O2CCMe3)12(diox)]n (1) and [Fe6O2(O2CH2)(O2CCMe3)12(4,4'-bpy)]n (2), showcase the utility of the angular overlap model, implemented in the program wxJFinder, in the predictive identification of the relative role of intra- and intercluster coupling.

19.
Acta Crystallogr C ; 69(Pt 10): 1144-6, 2013 Oct.
Artigo em Inglês | MEDLINE | ID: mdl-24096504

RESUMO

The structure of the title compound, {[Mg(C4H7O2)2(H2O)3]·H2O}n, features one-dimensional ···(µ2-ib)Mg(µ2-ib)Mg··· zigzag chains (ib is isobutyrate) parallel to the c axis. The octahedral Mg environment is completed by three fac-oriented terminal water ligands, as well as one further monodentate end-on coordinated ib ligand. In the crystal structure, the hydrophobic ib groups are all oriented within one half of the coordination perimeter of each chain, whereas the water ligands, together with hydrogen-bonded noncoordinated solvent water molecules, define the other half. Along the a axis, neighbouring strands are oriented so that both the hydrophilic and hydrophobic sides are adjacent to each other. This results in an extensive hydrogen-bonding network within the hydrophilic areas, also involving an additional solvent water molecule per formula unit. There are van der Waals contacts between the aliphatic isopropyl groups of the hydrophobic areas.

20.
J Am Chem Soc ; 134(31): 13018-26, 2012 Aug 08.
Artigo em Inglês | MEDLINE | ID: mdl-22839354

RESUMO

Metal alkylperoxides are remarkable, highly effective, yet often thermally unstable, oxidants that may react through a number of possible pathways including O-O homolytic cleavage, M-O homolytic cleavage, nucleophilic O-atom transfer, and electrophilic O-atom transfer. Here we describe a series of zinc alkyl compounds of the type To(M)ZnR (To(M) = tris(4,4-dimethyl-2-oxazolinyl)phenylborate; R = Et, n-C3H7, i-C3H7, t-Bu) that react with O2 at 25 °C to form isolable monomeric alkylperoxides To(M)ZnOOR in quantitative yield. The series of zinc alkylperoxides is crystallographically characterized, and the structures show systematic variations in the Zn-O-O angle and O-O distances. The observed rate law for the reaction of To(M)ZnEt (2) and O2 is consistent with a radical chain mechanism, where the rate-limiting SH2 step involves the interaction of (•)OOR and To(M)ZnR. In contrast, To(M)ZnH and To(M)ZnMe are unchanged even to 120 °C under 100 psi of O2 and in the presence of active radical chains (e.g., (•)OOEt). This class of zinc alkylperoxides is unusually thermally robust, in that the compounds are unchanged after heating at 120 °C in solution for several days. Yet, these compounds are reactive as oxidants with phosphines. Additionally, an unusual alkylperoxy group transfer to organosilanes affords To(M)ZnH and ROOSiR3'.

SELEÇÃO DE REFERÊNCIAS
DETALHE DA PESQUISA