Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 17 de 17
Filtrar
Mais filtros










Base de dados
Intervalo de ano de publicação
1.
Phys Chem Chem Phys ; 26(7): 5932-5946, 2024 Feb 14.
Artigo em Inglês | MEDLINE | ID: mdl-38299635

RESUMO

Ionic liquids have drawn great interest as electrolytes for energy storage applications in which they form characteristic electrical double layers at electrode interfaces. For ionic liquids at carbon electrode interfaces, their double layers are subject to nanoscale structuring of the electrode surface, involving altered ion structure and interactions that significantly influence the double layer capacitance. In this regard, we investigate the modulation of ionic liquid double layers by electrode surface roughness and the resulting effects on the ion structure, interaction, and capacitance. We performed fixed voltage molecular dynamics simulations to compute the differential capacitance profiles for the ionic liquids [BMIm+][TFSI-] and [BMIm+][FSI-] at model carbon electrode interfaces with the surface channel width at subnanometer and nanometer scales. We find that both [BMIm+][TFSI-] and [BMIm+][FSI-] exhibit enhanced differential capacitance for the electrode surface with a subnanometer channel width relative to the flat graphene surface, but the most pronounced enhancements for these two ionic liquids unexpectedly appear at different applied potential regimes. For [BMIm+][TFSI-], the nanostructured electrode shows significant enhancement of capacitance at high positive potential. For [BMIm+][FSI-], on the other hand, this enhancement is small at positive polarization but noticeable at low negative potential. We demonstrate that differences in these capacitance trends is due to differences in ion correlation that arise from a steric constraint of nanostructured electrode surface on the voltage-mediated restructuring of ions closest to the electrode interface. For example, the TFSI- and FSI- anions tend to structure with their charged and nonpolar groups in contact with the positive electrode surface when the constraint on these close-contact anions is relaxed. This anion structuring largely retains the cation association near the nanostructured electrode, resulting in only a slight increase in capacitance at positive polarization. Our simulations highlight the sensitive dependence of the innermost ion structure on the electrode surface nanostructure and applied voltage and the resulting influence on ion correlation and capacitance of ionic liquid double layers.

2.
Chemistry ; 29(28): e202300081, 2023 May 16.
Artigo em Inglês | MEDLINE | ID: mdl-36975096

RESUMO

Molecular lanthanide phosphonates [Ln2 (H3 tpmm)2 (H2 O)6 ] ⋅ xH2 O (Ln=Eu, EuP; Ln=Tb, TbP) were synthesized. Single-crystal X-ray diffraction confirmed that EuP has a sandwich-like dinuclear structure, in which the Eu(III) center adopts a {EuO8 } distorted dodecahedral geometry. XRPD patterns prove that TbP and EuP are isomorphous and isostructural. EuP and TbP are highly thermally stable approaching 450 °C and exhibit red- and green-light emissions from the characteristic 4 f-4 f transition of the Eu3+ and Tb3+ , respectively. Interestingly, luminescence modulation is achieved for the chemically mixed Eu/Tb phosphonate analogues, c-Eux Tb2 -x P (x=1.5, 1, 0.5), and physically mixed Eu/Tb phosphonate materials, p-yEuP : zTbP (y : z=3 : 1, 1 : 1, 1 : 3), with varying the excitation wavelength. Of particular note, near-white-light emission is also achieved for c-EuTbP, p-EuP : TbP, and p-EuP : 3TbP when excited at 365 nm. Therefore, these dinuclear molecular lanthanide phosphonates emitting excitation wavelength and Eu3+ : Tb3+ ratio dependent luminescence might be potential candidates for color-tunable luminescence materials and white-light-emitting materials. On the other hand, the bright green-light emission makes TbP to be an excellent reusable luminescence sensor for selective detection of Fe3+ with Stern-Volmer quenching constant (KSV ) of 9.66×103  M-1 and detection limit (DL) of 0.42 µM through absorption competition caused luminescence quenching effect.

3.
Free Radic Biol Med ; 159: 37-43, 2020 11 01.
Artigo em Inglês | MEDLINE | ID: mdl-32738399

RESUMO

Ascorbic acid (vitamin C) is an unusual antioxidant in that it donates a single reducing equivalent, and the radical it forms, monodehydroascorbate, reacts preferentially with radicals instead of with non-radical compounds. This happens because removal of an electron from monodehydroascorbate would create a tricarbonyl structure that is energetically unfavored. Instead of forming this structure, ascorbic acid oxidizes only to monodehydroascorbate, and monodehydroascorbate reacts with other radicals, oxidizing by mechanisms that may circumvent formation of this unfavored structure. Ironically, this tricarbonyl compound, which we suggest be called pseudodehydroascorbate, is commonly and mistakenly cited as the real product of ascorbic acid oxidation. In fact, it has been known for over 40 years that dehydroascorbate has a bicyclic hemiketal structure, and kinetic considerations suggest that it may be produced and reduced without forming pseudodehydroascorbate as an intermediate. This and other significant questions about the chemical basis of the antioxidant properties of ascorbic acid are obscured by this misconception about its oxidation product, dehydroascorbate.


Assuntos
Antioxidantes , Ácido Ascórbico , Cinética , Oxirredução
4.
ACS Catal ; 9(3): 2110-2123, 2019 Mar 01.
Artigo em Inglês | MEDLINE | ID: mdl-31819822

RESUMO

Metal triflates have been utilized to catalytically facilitate numerous glycosylation reactions under mild conditions. In some methods, the metal triflate system provides stereocontrol during the glycosylation, rather than the nature of protecting groups on the substrate. Despite these advances, the true activating nature of metal triflates remains unclear. Our findings indicated that the in situ generation of trace amounts of triflic acid from metal triflates can be the active catalyst species in the glycosylation. This fact has been mentioned previously in metal triflate-catalyzed glycosylation reactions; however, a thorough study on the subject and its implications on stereoselectivity has yet to be performed. Experimental evidence from control reactions and 19F NMR spectroscopy have been obtained to confirm and quantify the triflic acid released from nickel triflate, for which it is of paramount importance in achieving a stereoselective 1,2-cis-2-amino glycosidic bond formation via a transient anomeric triflate. A putative intermediate resembling that of a glycosyl triflate has been detected using variable temperature NMR (1H and 13C) experiments. These observations, together with density functional theory calculations and a kinetic study, corroborate a mechanism involving triflic acid-catalyzed stereoselective glycosylation with N-substituted trifluoromethylbenzylideneamino protected electrophiles. Specifically, triflic acid facilitates formation of a glycosyl triflate intermediate which then undergoes isomerization from the stable α-anomer to the more reactive ß-anomer. Subsequent SN2-like displacement of the reactive anomer by a nucleophile is highly favorable for the production of 1,2-cis-2-aminoglycosides. Although there is a previously reported work regarding glycosyl triflates, none of these reports have been confirmed to come from the counter ion of the metal center. Our work provides supporting evidence for the induction of a glycosyl triflate through the role of triflic acid in metal triflate-catalyzed glycosylation reactions.

5.
Angew Chem Int Ed Engl ; 58(21): 6957-6961, 2019 05 20.
Artigo em Inglês | MEDLINE | ID: mdl-30920099

RESUMO

Carbohydrates are essential moieties of many bioactive molecules in nature. However, efforts to elucidate their modes of action are often impeded by limitations in synthetic access to well-defined oligosaccharides. Most of the current methods rely on the design of specialized coupling partners to control selectivity during the formation of glycosidic bonds. Reported herein is the use of a commercially available phenanthroline to catalyze stereoretentive glycosylation with glycosyl bromides. The method provides efficient access to α-1,2-cis glycosides. This protocol has been performed for the large-scale synthesis of an octasaccharide adjuvant. Density-functional theory calculations, together with kinetic studies, suggest that the reaction proceeds by a double SN 2 mechanism.


Assuntos
Brometos/química , Glicosídeos/química , Fenantrolinas/química , Catálise , Glicosilação , Cinética , Estereoisomerismo
6.
J Chem Phys ; 148(2): 024503, 2018 Jan 14.
Artigo em Inglês | MEDLINE | ID: mdl-29331119

RESUMO

We report a kinetic study of the water exchange on lanthanide ions in water/[1-ethyl-3-methylimidazolium][trifluoromethylsufate] (water/[EMIm][OTf]). The results from 17O-NMR measurements show that the water-exchange rates in water/[EMIm][OTf] increase with decreasing size of the lanthanide ions. This trend for water-exchange is similar to the previously reported trend in water/1-ethyl-3-methylimidazolium ethyl sulfate (water/[EMIm][EtSO4]) but opposite to that in water. To gain atomic-level insight into these water-exchange reactions, molecular dynamics simulations for lanthanide ions in water/[EMIm][OTf] have been performed using the atomic-multipole-optimized-energetics-for-biomolecular-application polarizable force field. Our molecular dynamics simulations reproduce the experimental water-exchange rates in terms of the trend and provide possible explanations for the observed experimental behavior. The smaller lanthanide ions in water/[EMIm][OTf] undergo faster water exchange because the smaller lanthanide ions coordinate to the first shell [OTf]- anions more tightly, resulting in a stronger screening effect for the second-shell water. The screening effect weakens the interaction of the lanthanide ions with the second-shell water molecules, facilitating the dissociation of water from the second-shell and subsequent association of water molecules from the outer solvation shells.

7.
Inorg Chem ; 57(1): 231-240, 2018 Jan 02.
Artigo em Inglês | MEDLINE | ID: mdl-29257679

RESUMO

Photoinduced ligand dissociation of pyridine occurs much more readily in [Ru(tpy)(Me2bpy)(py)]2+ than in [Ru(tpy)(bpy)(py)]2+ (tpy = 2,2':6',2″-terpyridine; bpy = 2,2'-bipyridine, Me2bpy = 6,6'-dimethyl-2,2'-bipyridine; py = pyridine). The S0 ground state and the 3MLCT and 3MC excited states of these complexes have been studied using BP86 density functional theory with the SDD basis set and effective core potential on Ru and the 6-31G(d) basis set for the rest of the atoms. In both complexes, excitation by visible light and intersystem crossing leads to a 3MLCT state in which an electron from a Ru d orbital has been promoted to a π* orbital of terpyridine, followed by pyridine release after internal conversion to a dissociative 3MC state. Interaction between the methyl groups and the other ligands causes significantly more strain in [Ru(tpy)(Me2bpy)(py)]2+ than in [Ru(tpy)(bpy)(py)]2+, in both the S0 and 3MLCT states. Transition to the dissociative 3MC states releases this strain, resulting in lower barriers for ligand dissociation from [Ru(tpy)(Me2bpy)(py)]2+ than from [Ru(tpy)(bpy)(py)]2+. Analysis of the molecular orbitals along relaxed scans for stretching the Ru-N bonds reveals that ligand photodissociation is promoted by orbital mixing between the ligand π* orbital of tpy in the 3MLCT state and the dσ* orbitals that characterize the dissociative 3MC states. Good overlap and strong mixing occur when the Ru-N bond of the leaving ligand is perpendicular to the π* orbital of terpyridine, favoring the release of pyridine positioned in a cis fashion to the terpyridine ligand.

8.
J Chem Phys ; 147(1): 013920, 2017 Jul 07.
Artigo em Inglês | MEDLINE | ID: mdl-28688410

RESUMO

We report the development of a new three-dimensional (3D) momentum-imaging setup based on conventional velocity map imaging to achieve the coincidence measurement of photoelectrons and photo-ions. This setup uses only one imaging detector (microchannel plates (MCP)/phosphor screen) but the voltages on electrodes are pulsed to push both electrons and ions toward the same detector. The ion-electron coincidence is achieved using two cameras to capture images of ions and electrons separately. The time-of-flight of ions and electrons are read out from MCP using a digitizer. We demonstrate this new system by studying the dissociative single and double ionization of PENNA (2-phenylethyl-N,N-dimethylamine). We further show that the camera-based 3D imaging system can operate at 10 kHz repetition rate.

9.
Org Biomol Chem ; 15(20): 4417-4431, 2017 May 23.
Artigo em Inglês | MEDLINE | ID: mdl-28485446

RESUMO

Ascorbic acid is a well-known antioxidant and radical scavenger. It can be oxidized by losing two protons and two electrons, but normally loses only one electron at a time. The reactivity of the ascorbate radical is unusual, in that it can either disproportionate or react with other radicals, but it reacts poorly with non-radical species. To explore the oxidation mechanism of ascorbic acid, the pKa's and reduction potentials have been calculated using the B3LYP/6-31+G(d,p) and CBS-QB3 levels of theory with the SMD implicit solvent model and explicit waters. Calculations show that the most stable form of dehydroascorbic acid in water is the bicyclic hydrated structure, in agreement with NMR studies. The possible oxidation reactions at different pH conditions can be understood by constructing a potential-pH (Pourbaix) diagram from the calculated pKa's and standard reduction potentials. At physiological pH disproportionation of the intermediate radical is thermodynamically favored. The calculations show that disproportionation proceeds via dimerization of ascorbate radical and internal electron transfer, as suggested by Bielski. In the dimer, one of the ascorbate units cyclizes. Then protonation and dissociation yields the fully reduced and bicyclic fully oxidized structures. Calculations show that this mechanism also explains the reaction of the ascorbic acid radical with other radical species such as superoxide. Ascorbate radical combines with the radical, and intramolecular electron transfer followed by cyclization and hydrolysis yields dehydroascorbic acid and converts the radical to its reduced form.

10.
Phys Chem Chem Phys ; 18(44): 30323-30333, 2016 Nov 09.
Artigo em Inglês | MEDLINE | ID: mdl-27828553

RESUMO

The dynamics of ligand exchange on lanthanide ions is important for catalysis and organic reactions. Recent 17O-NMR experiments have shown that water-exchange rates of lanthanide ions in water/1-ethyl-3-methylimidazolium ethyl sulfate (water/[EMIm][EtSO4]) increase as a function of increasing charge density. The trend of water-exchange rates in this solvent is opposite to that observed in water. Since the lanthanide ions and ionic liquids investigated in that work were highly charged, an advanced polarizable potential is desirable for accurate simulations. To this end, we have developed atomic multipole optimized energetics for biomolecular applications (AMOEBA) parameters for all lanthanides and [EMIm][EtSO4], and molecular dynamics simulations with the optimized parameters have been carried out to provide possible explanations for these observed behaviors from the experiments. In water, the association of a water molecule with the first hydration shell can lead to water exchange. Smaller lanthanide ions exhibit slower water-exchange rates than larger ones because they form smaller aqua complexes, preventing the binding of incoming water molecules from the outer hydration shells. By contrast, smaller lanthanide ions undergo faster water exchange in water/[EMIm][EtSO4] because the dissociation of a water molecule is a key step for water-exchange events in this solvent. The first shell [EtSO4]- anions bind closer to the smaller lanthanide ions, resulting in more steric crowding effects on the surrounding water and facilitating the release of water molecules.

11.
Phys Chem Chem Phys ; 18(44): 30877, 2016 11 09.
Artigo em Inglês | MEDLINE | ID: mdl-27711761

RESUMO

Correction for 'Simulations of the water exchange dynamics of lanthanide ions in 1-ethyl-3-methylimidazolium ethyl sulfate ([EMIm][EtSO4]) and water' by Yi-Jung Tu et al., Phys. Chem. Chem. Phys., 2016, DOI: 10.1039/c6cp04957e.

12.
Inorg Chem ; 54(16): 8003-11, 2015 Aug 17.
Artigo em Inglês | MEDLINE | ID: mdl-26244447

RESUMO

Metal complexes that release ligands upon photoexcitation are important tools for biological research and show great potential as highly specific therapeutics. Upon excitation with visible light, [Ru(TQA)(MeCN)2](2+) [TQA = tris(2-quinolinylmethyl)amine] exchanges one of the two acetonitriles (MeCNs), whereas [Ru(DPAbpy)MeCN](2+) [DPAbpy = N-(2,2'-bipyridin-6-yl)-N,N-bis(pyridin-2-ylmethyl)amine] does not release MeCN. Furthermore, [Ru(TQA)(MeCN)2](2+) is highly selective for release of the MeCN that is perpendicular to the plane of the two axial quinolines. Density functional theory calculations provide a clear explanation for the photodissociation behavior of these two complexes. Excitation by visible light and intersystem crossing leads to a six-coordinate (3)MLCT state. Dissociation of acetonitrile can occur after internal conversion to a dissociative (3)MC state, which has an occupied dσ* orbital that interacts in an antibonding fashion with acetonitrile. For [Ru(TQA)(MeCN)2](2+), the dissociative (3)MC state is lower than the (3)MLCT state. In contrast, the (3)MC state of [Ru(DPAbpy)MeCN](2+) that releases acetonitrile has an energy higher than that of the (3)MLCT state, indicating dissociation is unfavorable. These results are consistent with the experimental observations that efficient photodissociation of acetonitrile occurs for [Ru(TQA)(MeCN)2](2+) but not for [Ru(DPAbpy)MeCN](2+). For the release of the MeCN ligand in [Ru(TQA)(MeCN)2](2+) that is perpendicular to the axial quinoline rings, the (3)MLCT state has an occupied quinoline π* orbital that can interact with a dσ* Ru-NCCH3 antibonding orbital as the Ru-NCCH3 bond is stretched and the quinolines bend toward the departing acetonitrile. This reduces the barrier for the formation of the dissociative (3)MC state, leading to the selective photodissociation of this acetonitrile. By contrast, when the acetonitrile is in the plane of the quinolines or bpy, no interaction occurs between the ligand π* orbital and the dσ* Ru-NCCH3 orbital, resulting in high barriers for conversion to the corresponding (3)MC structures and no release of acetonitrile.


Assuntos
Acetonitrilas/química , Compostos Organometálicos/química , Processos Fotoquímicos , Piridinas/química , Teoria Quântica , Rutênio/química , Ligantes , Modelos Moleculares , Conformação Molecular , Propriedades de Superfície
13.
J Inorg Biochem ; 134: 118-33, 2014 May.
Artigo em Inglês | MEDLINE | ID: mdl-24629413

RESUMO

The control over the regio- and/or stereo-selective aliphatic CH oxidation by metalloenzymes is of great interest to scientists. Typically, these enzymes invoke host-guest chemistry to sequester the substrates within the protein pockets, exploiting sizes, shapes and specific interactions such as hydrogen-bonding, electrostatic forces and/or van der Waals interactions to control the substrate specificity, regio-specificity and stereo-selectivity. Over the years, we have developed a series of deuterated and fluorinated variants of these hydrocarbon substrates as probes to gain insights into the controlled CH oxidations of hydrocarbons facilitated by these enzymes. In this review, we illustrate the application of these designed probes in the study of three monooxygenases: (i) the particulate methane monooxygenase (pMMO) from Methylococcus capsulatus (Bath), which oxidizes straight-chain C1-C5 alkanes and alkenes to form their corresponding 2-alcohols and epoxides, respectively; (ii) the recombinant alkane hydroxylase (AlkB) from Pseudomonas putida GPo1, which oxidizes the primary CH bonds of C5-C12 linear alkanes; and (iii) the recombinant cytochrome P450 from Bacillus megaterium, which oxidizes C12-C20 fatty acids at the ω-1, ω-2 or ω-3 CH positions.


Assuntos
Proteínas de Bactérias/química , Citocromo P-450 CYP4A/química , Sistema Enzimático do Citocromo P-450/química , Deutério/química , Hidrocarbonetos Fluorados/química , Oxigenases/química , Bacillus megaterium/química , Bacillus megaterium/enzimologia , Proteínas de Bactérias/isolamento & purificação , Citocromo P-450 CYP4A/isolamento & purificação , Sistema Enzimático do Citocromo P-450/genética , Cinética , Methylococcus capsulatus/química , Methylococcus capsulatus/enzimologia , Modelos Moleculares , Oxirredução , Oxigenases/isolamento & purificação , Pseudomonas putida/química , Pseudomonas putida/enzimologia , Proteínas Recombinantes/química , Proteínas Recombinantes/genética , Especificidade por Substrato , Termodinâmica
14.
Chemistry ; 19(41): 13680-91, 2013 Oct 04.
Artigo em Inglês | MEDLINE | ID: mdl-24092541

RESUMO

We demonstrate herein that wild-type cytochrome P450 BM3 can recognize non-natural substrates, such as fluorinated C12 -C15 chain-length fatty acids, and show better catalysis for their efficient conversion. Although the binding affinities for fluorinated substrates in the P450 BM3 pocket are marginally lower than those for non-fluorinated substrates, spin-shift measurements suggest that fluoro substituents at the ω-position can facilitate rearrangement of the dynamic structure of the bulk-water network within the hydrophobic pocket through a micro desolvation process to expel the water ligand of the heme iron that is present in the resting state. A lowering of the Michaelis-Menten constant (Km ), however, indicates that fluorinated fatty acids are indeed better substrates compared with their non-fluorinated counterparts. An enhancement of the turnover frequencies (kcat ) for electron transfer from NADPH to the heme iron and for CH bond oxidation by compound I (Cpd I) to yield the product suggests that the activation energies associated with going from the enzyme-substrate (ES state) to the corresponding transition state (ES(≠) state) are significantly lowered for both steps in the case of the fluorinated substrates. Delicate control of the regioselectivity by the fluorinated terminal methyl groups of the C12 -C15 fatty acids has been noted. Despite the fact that residues Arg47/Tyr51/Ser72 exert significant control over the hydroxylation of the subterminal carbon atoms toward the hydrocarbon tail, the fluorine substituent(s) at the ω-position affects the regioselective hydroxylation. For substrate hydroxylation, we have found that fluorinated lauric acids probably give a better structural fit for the heme pocket than fluorinated pentadecanoic acid, even though pentadecanoic acid is by far the best substrate among the reported fatty acids. Interestingly, 12-fluorododecanoic acid, with only one fluorine atom at the terminal methyl group, exhibits a comparable turnover frequency to that of pentadecanoic acid. Thus, fluorination of the terminal methyl group introduces additional interactions of the substrate within the hydrophobic pocket, which influence the electron transfers for both dioxygen activation and the controlled oxidation of aliphatics mediated by high-valent oxoferryl species.


Assuntos
Ácidos Graxos/química , Ácidos Láuricos/química , Sistema Enzimático do Citocromo P-450 , Fluorocarbonos , Halogenação , Ligação de Hidrogênio , Hidroxilação , Oxirredução
15.
Org Lett ; 15(15): 3868-71, 2013 Aug 02.
Artigo em Inglês | MEDLINE | ID: mdl-23883270

RESUMO

By increasing the number of amino substituents on triarylamine, the extent of hydrogen bonding between the oxidized form of triarylamine and imidazole could be electrochemically controlled. Three behaviors, depending on the interaction between oxidized amine and imidazole, were obtained in CV patterns. DFT calculation was used to confirm that the electron density of protons of the amino group decreased as the amino moiety increased.

16.
J Phys Chem A ; 116(6): 1632-7, 2012 Feb 16.
Artigo em Inglês | MEDLINE | ID: mdl-22229836

RESUMO

Electrochemical properties of substituted free base meso-tetraphenylporphyrins (H(2)T(o,o'-X)PP, H(2)T(o-X)PP, and H(2)T(p-X)PP, where X = OCH(3), CH(3), H, F, or Cl on the phenyl rings) are examined by cyclic voltammetry. When a substituent is located only at the para position of the meso-phenyl group, the difference between the first and second oxidation potentials (ΔE(ox), i.e., E(2)(ox) - E(1)(ox)), is generally significantly smaller than those of the H(2)TPPs with bulky o,o'-substituents on the phenyl group. This trend is elucidated with density functional theory calculations and attributed mainly to the sterically controlled π-conjugation of the meso-phenyl groups to the central porphyrin ring, rather than the often discussed deformation of porphyrin.

17.
Org Lett ; 13(11): 2826-9, 2011 Jun 03.
Artigo em Inglês | MEDLINE | ID: mdl-21534612

RESUMO

In electrochemical oxidations, the second oxidation potential of phenylenediamines (PD) varies because of hydrogen-bonding formation for PD(+•) with pyridines. A linear relationship was obtained for the potential shift as a function of pK(a) of the protonated pyridines and potential inversion could be observed. The oxidized PD(+•) could also form hydrogen bonding with alcohols and the shift of potential exhibits a different pattern.

SELEÇÃO DE REFERÊNCIAS
DETALHE DA PESQUISA
...