Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 45
Filtrar
1.
Inorg Chem ; 62(21): 8069-8079, 2023 May 29.
Artículo en Inglés | MEDLINE | ID: mdl-37195088

RESUMEN

Extended transition state-natural orbitals for chemical valence (ETS-NOCV)] data were used to rank electron acceptor capacities for several potentially synergistic anionic ligands incorporated into linear d10 [(NH3)Pd(A)]-, square planar d8 [(NN2)Ru(A)]-, and octahedral d6 [(AsN4)Tc(A)]- complexes [A = anionic ligand, NN2 = HN(CH2CH2CH2NH2)2, and AsN4 = [As(CH2CH2CH2NH2)4]-]. It was possible to differentiate between the best acceptors, among them BI2- and B(CF3)2-, and the poorest ones. A sizable fraction of the anionic ligands studied exhibit similar acceptor capacities (backbonding), mostly regardless of d electron count. A number of trends were discerned, including the fact that acceptor capacity decreases down families and across rows but increases down families of the peripheral substituents. The latter appears tied to the ability of the peripheral ligands to compete with the metal in donating electrons to the ligand-binding atom.

2.
Inorg Chem ; 60(16): 12218-12225, 2021 Aug 16.
Artículo en Inglés | MEDLINE | ID: mdl-34353020

RESUMEN

"MoCl3(dme)" (dme = 1,2-dimethoxyethane) is an important precursor for midvalent molybdenum chemistry, particularly for triply Mo-Mo bonded compounds of the type Mo2X6 (X = bulky anionic ligand). However, its exact structural identity has been obscure for more than 50 years. In search of a convenient, large-scale synthesis, we have found that trans-MoCl4(Et2O)2 dissolved in dme can be cleanly reduced with dimethylphenylsilane, Me2PhSiH, to provide khaki Mo2Cl6(dme)2 in ∼90% yield. If the reduction is performed on a small scale, single crystals suitable for X-ray crystallography can be obtained. Two different crystal morphologies were identified, each belonging to the P21/n space group, but with slightly different unit cell constants. The refined structure of each form is an edge-shared bioctahedron with overall Ci symmetry and metal-metal separations on the order of 2.8 Å. The bulk material is diamagnetic as determined by both the Gouy method and SQUID magnetometry. Density functional theory calculations suggest a σ2π2δ*2 ground state for the dimer with the diamagnetism arising from a singlet diradical "broken symmetry" electronic configuration. In addition to a definitive structural assignment for "MoCl3(dme)", this work highlights the utility of organosilanes as easy to handle, alternative reductants for inorganic synthesis.

3.
J Phys Chem A ; 125(24): 5456-5469, 2021 Jun 24.
Artículo en Inglés | MEDLINE | ID: mdl-34110819

RESUMEN

Systematic computational studies of pericyclic Diels-Alder-type reactions between aminoborane (F3C)2B = N(CH3)2, 1, and all permutations of substituted cyclopentadienes c-C5R1R2R3R4R5aR5b (R = H, CH3, CF3, F) allow isolation of substitutional effects on Gibbs free energy barrier heights and reaction Gibbs free energies. The effects appear to be additive in all cases. Substitution at positions 5a and 5b always increases barriers and reaction energies, an effect explained by steric interactions between substituents and the aminoborane moiety. For cases R = CH3, regioselectivities differ from those expected from canonical organic chemistry predictions. Frontier molecular orbital calculations suggest this arises from the extreme polarization of the π interaction in 1. For cases R = CF3, the 2/3-substitution comparison accords with canon, but the 1/4-substitution comparison does not. This appears to arise from a combination of electronic and steric issues. For cases R = F, many of the reactions are exergonic, in contrast to the cases R = CH3, CF3. Additionally, fluorine substitution at positions 2 and 4 has a barrier-lowering effect. Frontier molecular orbital calculations support an orbital-based preference for formation of 2- and 4-substituted "meta" products rather than "ortho/para" products.

4.
Chemistry ; 25(42): 9959-9966, 2019 Jul 25.
Artículo en Inglés | MEDLINE | ID: mdl-31090119

RESUMEN

A series of zinc-based catalysts was evaluated for their efficiency in decomposing formic acid into molecular hydrogen and carbon dioxide in the gas phase using quadrupole ion trap mass spectrometry experiments. The effectiveness of the catalysts in the series [(L)Zn(H)]+ , where L=2,2':6',2''-terpyridine (tpy), 1,10-phenanthroline (phen) or 2,2'-bipyrydine (bpy), was found to depend on the ligand used, which turned out to be fundamental in tuning the catalytic properties of the zinc complex. Specifically, [(tpy)Zn(H)]+ displayed the fastest reaction with formic acid proceeding by dehydrogenation to produce the zinc formate complex [(tpy)Zn(O2 CH)]+ and H2 . The catalysts [(L)Zn(H)]+ are reformed by decarboxylating the zinc formate complexes [(L)Zn(O2 CH)]+ by collision-induced dissociation, which is the only reaction channel for each of the ligands used. The decarboxylation reaction was found to be reversible, since the zinc hydride complexes [(L)Zn(H)]+ react with carbon dioxide yielding the zinc formate complex. This reaction was again substantially faster for L=tpy than L=phen or bpy. The energetics and mechanisms of these processes were modelled using several levels of density functional theory (DFT) calculations. Experimental results are fully supported by the computational predictions.

5.
J Org Chem ; 84(9): 5021-5026, 2019 May 03.
Artículo en Inglés | MEDLINE | ID: mdl-30973732

RESUMEN

The asymmetric direct reductive amination of prochiral ketones with aryl amines using 1-hydrosilatrane with a chiral Brønsted acid catalyst is reported. This is the first known example of chiral Brønsted acid-catalyzed asymmetric reductive amination using a silane as the hydride source. The reaction features a highly practical reducing reagent and proceeds efficiently at room temperature without a specialized reaction setup or equipment to exclude air or moisture. This method provides high conversion and enantiomeric excess up to 84% of the desired chiral secondary amines with minimal side products.

6.
J Phys Chem A ; 123(43): 9413-9419, 2019 Oct 31.
Artículo en Inglés | MEDLINE | ID: mdl-31600076

RESUMEN

The reaction between (F3C)2B═NMe2, 1, and acetonitrile at low temperature in pentane yields a bora-acetonitrile rather than the expected coordination complex. This appears to arise from the two undergoing an ene reaction followed by a rearrangement analogous to a ketenimine-nitrile rearrangement. Computational studies indicate that mechanistic steps suggested for the latter require energies too large for the reaction to take place under the experimental conditions. Instead, a mechanism in which the ene reaction product is attacked by a second molecule of 1, followed by hydrogen transfer and decomposition, exhibits barriers lower than that for the ene reaction. The mechanism implies that the fragment of 1 in the observed product is not the one that underwent the ene reaction. The ene reaction barrier is rate-determining, and it is low enough to conform to the experimental conditions.

7.
J Phys Chem A ; 123(12): 2438-2446, 2019 Mar 28.
Artículo en Inglés | MEDLINE | ID: mdl-30835465

RESUMEN

Computational studies of low spin d6 cis- and trans-[M(en)2X2]+ complexes (M = Co, Rh, Ir) employing multiple model chemistries find that isomer preferences fall into three categories. Complexes where X is largely a σ-donor (H-, CH3-, CF3-) prefer cis geometries, in keeping with predictions associated with the trans influence series. Complexes where this donor characteristic is augmented by π acceptor behavior (B(CF3)2-, BCl2-, SiCl3-) evince even greater preference for cis geometries. QTAIM charge data suggest this is marked by lower positive charge on the metal in cis complexes. In contrast, complexes where X is a π donor and low in the trans influence series (X = OH-, F-, Cl-, I-) prefer trans geometries to varying degrees. QTAIM calculations indicate that this arises because the cis complexes are destabilized by distortions of the electron density in the M-X bonds. This can be viewed conceptually as resulting from repulsions between lone pair electrons on the ligands. Complexes where the X ligands are moderately trans-influencing and can interact conjugatively (CN-, NC-, NO2-, C≡CH-) prefer trans geometries because they combine destabilization of cis geometries with enhanced stabilization of trans geometries resulting from conjugation.

8.
J Phys Chem A ; 120(46): 9315-9323, 2016 Nov 23.
Artículo en Inglés | MEDLINE | ID: mdl-27805402

RESUMEN

exo-Silatranes involve cage structures where the nitrogen lone pair points away from the cage rather than into it. This distinguishes them from the well-established endo-silatranes. exo-Silatranes have not been observed experimentally, consistent with a significant benefit to silicon-nitrogen interactions inside the cages as suggested for endo-silatranes. Identifying examples of exo-silatranes would prove useful in understanding Si-N interactions, as they would represent the "no interaction" extreme of the spectrum. We have found four means by which exo-silatranes might be synthesized: (1) employing smaller cages; (2) employing constrained rings to stiffen the cage backbones; (3) employing steric interactions to enhance preference for the less crowded exo-geometry around nitrogen; (4) modifying the Lewis acidity and basicity of the silicon and nitrogen so significantly as to remove their desire to interact. The preference for exo geometries is established using the parameter Δ, representing the distance between the nitrogen atom and the least-squares plane containing the adjacent carbon atoms. In some cases, Δ values for exo-silatranes are greater than 0.3 Å. In others, they are near zero, indicating a nearly planar nitrogen atom. There are no obvious structural markers besides Δ that distinguish between exo- and endo-silatranes.

9.
J Org Chem ; 80(20): 10234-43, 2015 Oct 16.
Artículo en Inglés | MEDLINE | ID: mdl-26383035

RESUMEN

Estimation of ring strain energies (RSEs) of substituted cyclohexanes c-C6H(x)R(12-x) (R = F, Cl, Me; x = 0, 2, 4, 8, 10, 12) using homodesmotic reaction methods gives implausible results for highly substituted cases, particularly, c-C6R12. Prior work suggests that this stems from poorly canceled interactions between substituents on the acyclic reference molecules. We apply here our semi-homodesmotic approach that minimizes use of acyclic references and ensures cancellation of intramolecular substituent interactions. The approach provides RSEs that are more consistent with chemical intuition, although they are higher than expected for "strain-free" cyclohexanes. The RSE for c-C6Me12 is predicted to be 11.9 kcal mol(-1). RSEs for halogenated rings rise significantly from 8-9 kcal mol(-1) for c-1,1,2,2-C6H8R4 to 44-50 kcal mol(-1) for c-C6R12 (R = F, Cl). The increase, and accompanying observation of larger RSEs for "adjacent CR2" systems, can be tied to increased bond distances in the rings upon progressive substitution. The sizable RSE for perchlorocyclohexane suggests that it may be susceptible to ring-opening reactions, a facet of its chemistry that is currently unexplored.

10.
J Phys Chem A ; 118(31): 6060-7, 2014 Aug 07.
Artículo en Inglés | MEDLINE | ID: mdl-25046030

RESUMEN

A semi-homodesmotic method for estimating of ring strain energies (RSEs) of substituted cyclopropanes is applied to substituted cyclobutanes c-C4HxR8-x (R = F, Cl, Me; x = 0, 2, 4). Whereas (hyper)homodesmotic reaction methods predict implausible results, particularly for c-C4R8, the semi-homodesmotic approach provides RSEs consistent with thermodynamic and independent computational data regardless of the degree of substitution. The method requires employing homodesmotic group equivalent reactions only for disubstituted cyclobutanes, relying solely on absolute energy calculations for more substituted rings. We find that, consistent with QTAIM data, RSEs increase with substitution regardless of the electronic nature of R, although the increase is more dramatic when R is electron-withdrawing. Overall, the semi-homodesmotic method is simpler than hyperhomodesmotic approaches and gives more trustworthy results.


Asunto(s)
Ciclobutanos/química , Modelos Químicos , Termodinámica
11.
J Phys Chem A ; 118(31): 6050-9, 2014 Aug 07.
Artículo en Inglés | MEDLINE | ID: mdl-25046121

RESUMEN

Estimation of ring strain energies (RSEs) of substituted cyclopropanes c-C(3)H(x)R(6-x) (R = F, Cl, Me; x = 0, 2, 4) using homodesmotic reaction methods has been plagued by implausible results. Prior work suggests that this stems from poorly canceled interactions between substituents on the acyclic reference molecules. We report a semi-homodesmotic approach that minimizes use of acyclic references, focusing instead on canceling substituent interactions. The method requires employing homodesmotic group equivalent reactions only for disubstituted cyclopropanes and relies solely on absolute energy calculations for more substituted rings. This provides RSEs consistent with chemical intuition regardless of the degree of substitution. We find that RSEs increase with substitution regardless of the electronic nature of R, although the increase is more dramatic when R is electron-withdrawing. The RSEs determined are consistent with QTAIM data, which show that progressive substitution always increases critical path angles. Overall, the semi-homodesmotic approach is simpler than hyperhomodesmotic reaction methods, and gives more trustworthy results.


Asunto(s)
Ciclopropanos/química , Modelos Químicos , Simulación por Computador , Electrones
12.
Top Curr Chem ; 332: 267-89, 2013.
Artículo en Inglés | MEDLINE | ID: mdl-23097031

RESUMEN

Computational studies that characterize the effects of Lewis acidity/basicity on FLP formation and reactivity are reviewed. Formation of the FLP encounter complex "cage" depends on Lewis acidities and basicities of substituent "external" atoms, and their abilities to interact intramolecularly. Computations indicate that these interactions are worth 9-18 kcal mol⁻¹ for partly fluorinated FLPs such as (F5C6)3B···P(t-Bu)3, and less for less fluorinated species such as (H5C6)3B···P(t-Bu)3. Reactivity within the cage depends on the "classical" Lewis acidities/basicities of the internal atoms. Energetics here fall into the range of 5-50 kcal mol⁻¹; the larger the value, the greater the ability of the FLP to capture or split a substrate. In several cases the computationally predicted reaction barriers differ little with internal Lewis acidity/basicity, indicating that the rate-determining step involves the substrate entering the cage rather than attack by the Lewis acid/base atoms. In others, barriers vary sizably with Lewis acidity/basicity, indicating the opposite. In one case it was shown that these effects cancel, such that the three component barriers are identical for a range of substituted Lewis acid FLP components.

13.
ACS Omega ; 8(15): 14160-14170, 2023 Apr 18.
Artículo en Inglés | MEDLINE | ID: mdl-37091433

RESUMEN

Systematic computational studies of pericyclic Diels-Alder reactions between (H3C)2C=C(CH3)2, 1, and all permutations of substituted cyclopentadienes c-C5R1R2R3R4R5aR5b (R = H, CH3, CF3, F) allowed isolation of substitutional effects on Gibbs free energy barrier heights and reaction Gibbs free energies. "Average Substitution Gibbs Free Energy Correction" ΔG ASC# ‡/ΔG ASC# values for each substituent in each position appeared to be additive. Substituent effects on barriers showed interesting contrasts. Methyl substitution at positions 5a and 5b increased barriers significantly, while substitution at all other positions had essentially no impact. In contrast, fluoro substitution at positions 5a and 5b lowered barriers more than substitution at other positions. Trifluoromethyl substitution mixed these effects, in that substitution at positions 5a and 5b increased barriers, but substitution at other positions lowered them. Despite the variances, ΔG ASC# ‡/ΔG ASC# values allowed reliable prediction of barriers and exergonicities for reactions between 1 and highly substituted cyclopentadienes, and between 1 and cyclopentadienes with random mixtures of CH3/CF3/F substituents. ΔG ASC# ‡/ΔG ASC# values were correlated with steric considerations and quantum theory of atoms in molecules (QTAIM) calculations. Overall, the ASC values provide a resource for predicting which Diels-Alder reactions of this type should occur at rapid rates and/or give stable bicyclic products.

14.
BMC Genomics ; 13: 177, 2012 May 10.
Artículo en Inglés | MEDLINE | ID: mdl-22574620

RESUMEN

BACKGROUND: Second-generation sequencing technologies have revolutionized our ability to recover genetic information from the past, allowing the characterization of the first complete genomes from past individuals and extinct species. Recently, third generation Helicos sequencing platforms, which perform true Single-Molecule DNA Sequencing (tSMS), have shown great potential for sequencing DNA molecules from Pleistocene fossils. Here, we aim at improving even further the performance of tSMS for ancient DNA by testing two novel tSMS template preparation methods for Pleistocene bone fossils, namely oligonucleotide spiking and treatment with DNA phosphatase. RESULTS: We found that a significantly larger fraction of the horse genome could be covered following oligonucleotide spiking however not reproducibly and at the cost of extra post-sequencing filtering procedures and skewed %GC content. In contrast, we showed that treating ancient DNA extracts with DNA phosphatase improved the amount of endogenous sequence information recovered per sequencing channel by up to 3.3-fold, while still providing molecular signatures of endogenous ancient DNA damage, including cytosine deamination and fragmentation by depurination. Additionally, we confirmed the existence of molecular preservation niches in large bone crystals from which DNA could be preferentially extracted. CONCLUSIONS: We propose DNA phosphatase treatment as a mechanism to increase sequence coverage of ancient genomes when using Helicos tSMS as a sequencing platform. Together with mild denaturation temperatures that favor access to endogenous ancient templates over modern DNA contaminants, this simple preparation procedure can improve overall Helicos tSMS performance when damaged DNA templates are targeted.


Asunto(s)
ADN/genética , Fósiles , Caballos/genética , Análisis de Secuencia de ADN/métodos , Análisis de Secuencia de ADN/normas , Animales , Composición de Base/genética , Secuencia de Bases , ADN Mitocondrial/genética , Genoma/genética , Nucleotidasas/metabolismo , Nucleótidos/genética , Filogenia
15.
Dalton Trans ; 51(20): 7856-7863, 2022 May 24.
Artículo en Inglés | MEDLINE | ID: mdl-35531983

RESUMEN

The bis(diethyl ether) and 1,2-dimethoxyethane (dme) adducts of molybdenum(IV) chloride and tungsten(IV) chloride are valuable starting materials for a variety of synthetic inorganic and organometallic reactions. Despite the broad utility and extensive use of these 6-coordinate complexes, their syntheses remain unoptimized, and their characterization incomplete after more than three decades. While exploring the ligand exchange behaviour of trans-MoCl4(OEt2)2, we obtained single crystals of this red-orange complex and subsequently compared its structural parameters with those of the recently reported trans-WCl4(OEt2)2. Significantly improved procedures for both MoCl4(dme) and WCl4(dme) were developed, and X-ray diffraction data were obtained and analysed. The magnetic properties of the dme adducts were probed, both with Gouy and SQUID magnetometry measurements. The magnetic moment of WCl4(dme) was smaller than that of MoCl4(dme), an observation that we attribute to the greater spin-orbit coupling of tungsten. Electronic structure studies were also conducted to probe the preferential trans configuration of the diethyl ether adducts and to assign the UV-Vis spectra of the dme adducts.

16.
J Comput Chem ; 32(8): 1493-9, 2011 Jun.
Artículo en Inglés | MEDLINE | ID: mdl-21328397

RESUMEN

The ability of the composite three-layer ONIOM G2R3 (OG2R3) method to match experimental dissociation energies for group 13-15 donor-acceptor complexes was examined for a database of 34 complexes. The composite approach provides energies that agree reasonably with experiment, performing nearly as well as both the CCSD(T)/aug-CC-pVTZ and CCSD(T)/6-311+G(2df, 2p) models for small molecules and nearly as well as the latter for slightly larger ones. Broadly, all three models exhibit average absolute errors of ∼3 kcal mol(-1) , and root mean square errors of ∼4 kcal mol(-1) . The average signed error suggest that the OG2R3 approach systematically underbinds by ∼2.3 kcal mol(-1) ; if this is used as a general correction, the approach performs as well or better than the pure CCSD(T) models. However, the OG2R3 model can be applied to molecules too large to be studied by the other CCSD(T) methods, as it requires only a fraction of the time and computer resources.

17.
Inorg Chem ; 50(16): 7871-9, 2011 Aug 15.
Artículo en Inglés | MEDLINE | ID: mdl-21774464

RESUMEN

Computational studies of the binding energies of all possible tris(fluoroaryl)boranes B(C(6)H(x)F(5-x))(3) to NMe(3) and PMe(3) show that they (and by extension, the Lewis acidities of the boranes) can be tuned to a sizable range of values through judicious placement of fluorines. This holds despite the fact that the B-X bond distance changes little regardless of substitution, save when the aryl ring is 2,6-disubstituted. Fluorine substitution appears to affect the binding energies additively. Least-squares regression finds substitution at the 2-position to increase the binding energy by ca. 13 kcal·mol(-1), while substitution at the 3- and 5-positions increases it by ca. 3 kcal·mol(-1). Substitution at the 4-position has little to no impact, while substitution at the 6-position decreases the binding energy by ca. 3-6 kcal·mol(-1). The last observation arises from steric congestion because the 6-position can only be substituted in tandem with substitution at the 2-position. Models suggest that the pattern arises from polarization effects that decrease exponentially as the distance between boron and fluorine increases.

18.
Inorg Chem ; 50(1): 336-44, 2011 Jan 03.
Artículo en Inglés | MEDLINE | ID: mdl-21126055

RESUMEN

The phosphinoboranes [R(2)PB(C(6)F(5))(2)](2) (R = Et 1, Ph 2) and R(2)PB(C(6)F(5))(2) (R = tBu 3, Cy 4, Mes 5) were synthesized from the reaction of (C(6)F(5))(2)BCl and the corresponding lithium phosphide. The relationships between B-P distance, P pyramidality, and the extent of BP multiple bonding were further explored computationally. Natural Bond Order (NBO) analyses of 3 and 4 showed that the π-bonding highest occupied molecular orbitals (HOMOs) were highly polarized. In addition the Lewis acid-base adducts, R(2)(H)P·B(H)(C(6)F(5))(2) (R = Et 6; Ph 7; tBu 8; Cy 9; Mes 10) were prepared via the reaction of the phosphines R(2)PH with the borane HB(C(6)F(5))(2). Compounds 1 and 2 showed no signs of reaction with H(2); however, reaction of compounds 3 and 4 with H(2) was observed to give 8 and 9. In a related set of reactions compounds 3 and 4 were reacted with H(3)NBH(3) or Me(2)(H)NBH(3) also led to the generation of 8 and 9, respectively. The reaction profile of the reaction of (CF(3))(2)BPR(2) with H(2) was examined computationally and shown to be exothermic. Efforts to effect the reverse reaction, that is, dehydrogenation of adducts 6-10 were unsuccessful. Compound 4 was also shown to react with 4-tert-butylpyridine to give Cy(2)PB(C(6)F(5))(2)(4-tBuC(5)H(4)N) 11 while reactions of 3 and 4 with the Lewis acid BCl(3) gave the dimers (R(2)PBCl(2))(2) (R = tBu 12, Cy 13) and the byproduct ClB(C(6)F(5))(2).

19.
Inorg Chem ; 48(20): 9910-7, 2009 Oct 19.
Artículo en Inglés | MEDLINE | ID: mdl-19757814

RESUMEN

The addition of diphenyl disulfide (PhSSPh) to tBu(2)P(C(6)F(4))B(C(6)F(5))(2) (1) affords the zwitterionic phosphonium borate [tBu(2)P(SPh)(C(6)F(4))B(SPh)(C(6)F(5))(2)] (2), while the addition of a base or donor solvent to 2 effected the liberation of disulfide and the formation of [tBu(2)P(C(6)F(4))B(donor)(C(6)F(5))(2)]. The reaction of 1 with S(8) gave tBu(2)P(S)(C(6)F(4))B(C(6)F(5))(2) (3). In a similar fashion, the frustrated Lewis pair of tBu(3)P/B(C(6)F(5))(3) reacts with RSSR to give [tBu(3)P(SR)][(RS)B(C(6)F(5))(3)] (R = Ph (4), p-tolyl (5), iPr (6)). In contrast, the corresponding reaction of BnSSBn yields a 1:1:1 mixture of tBu(3)P horizontal lineS, Bn(2)S, and B(C(6)F(5))(3). Species 4 reacts with p-tolylSSp-tolyl to give a mixture of 4, 5, PhSSPh, and p-tolylSS p-tolyl, while treatment of 5 with PhSSPh afforded a similar mixture. To probe this, a crossover experiment between [tBu(3)P(SPh)][B(C(6)F(5))(4)] (7) and [NBu(4)][(p-tolylS)B(C(6)F(5))(3)] (9) was performed. The former species was prepared by a reaction of 4 with [Ph(3)C][B(C(6)F(5)) (4)], while cation exchange of [(Et(2)O)(2)Li( p-tolylS)B(C(6)F(5))(3)] (8) with [NBu(4)]Br gave 9. The reaction of compounds 7 and 9 gave a statistical mixture of the cations [tBu(3)P(SR)](+) and anions [(RS)B(C(6)F(5))(3)](-), R = Ph, Sp-tolyl. The mechanism of this exchange process was probed and is proposed to be an equilibrium involving disulfide and the frustrated Lewis pair. Crystallographic data are reported for compounds 4-8, and the natures of the P-S cations are examined via DFT calculations.

20.
Inorg Chem ; 48(21): 10466-74, 2009 Nov 02.
Artículo en Inglés | MEDLINE | ID: mdl-19810714

RESUMEN

The pyridine adducts of B(C(6)F(5))(3), (4-tBu)C(5)H(4)NB(C(6)F(5))(3) 1, ((2-Me)C(5)H(4)N)B(C(6)F(5))(3) 2, ((2-Et)C(5)H(4)N)B(C(6)F(5))(3) 3, ((2-Ph)C(5)H(4)N)B(C(6)F(5))(3) 4, ((2-C(5)H(4)N)C(5)H(4)N)B(C(6)F(5))(3) 5, (C(9)H(7)N)B(C(6)F(5))(3) 6, and ((2-C(5)H(4)N)NH(2-C(5)H(4)N))B(C(6)F(5))(3) 7, were prepared and characterized. The B-N bond lengths in 2-7 reflect the impact of ortho-substitution, increasing significantly with sterically larger and electron-withdrawing substituents. In the case of 2-amino-6-picoline, reaction with B(C(6)F(5))(3) affords the zwitterionic species (5-Me)C(5)H(3)NH(2-NH)B(C(6)F(5))(3) 8. In contrast, lutidine/B(C(6)F(5))(3) yields an equilibrium mixture containing both the free Lewis acid and base and the adduct (2,6-Me(2)C(5)H(3)N)B(C(6)F(5))(3) 9. This equilibrium has a DeltaH of -42(1) kJ/mol and DeltaS of -131(5) J/mol x K. Addition of H(2) shifts the equilibrium and yields [2,6-Me(2)C(5)H(3)NH][HB(C(6)F(5))(3)] 10. The corresponding reactions of 2,6-diphenylpyridine or 2-tert-butylpyridine with B(C(6)F(5))(3) showed no evidence of adduct formation and upon exposure to H(2) afforded [(2,6-Ph(2))C(5)H(3)NH][HB(C(6)F(5))(3)] 11 and [(2-tBu)C(5)H(4)NH][HB(C(6)F(5))(3)] 12, respectively. The energetics of adduct formation and the reactions with H(2) are probed computationally. Crystallographic data for compounds 1-10 are reported.

SELECCIÓN DE REFERENCIAS
DETALLE DE LA BÚSQUEDA