Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 64
Filtrar
1.
Phys Chem Chem Phys ; 25(28): 18844-18853, 2023 Jul 19.
Artigo em Inglês | MEDLINE | ID: mdl-37403595

RESUMO

Laser-induced reduction of metal ions is attracting increasing attention as a sustainable route to ligand-free metal nanoparticles. In this work, we investigate the photochemical reactions involved in reduction of Ag+ and [AuCl4]- upon interaction with lasers with nanosecond and femtosecond pulse duration, using strong-field ionization mass spectrometry and spectroscopic assays to identify stable molecular byproducts. Whereas Ag+ in aqueous isopropyl alcohol (IPA) is reduced through plasma-mediated mechanisms upon femtosecond laser excitation, low-fluence nanosecond laser excitation induces electron transfer from IPA to Ag+. Both nanosecond and femtosecond laser excitation of aqueous [AuCl4]- produce reactive chlorine species by Au-Cl bond homolysis. Formation of numerous volatile products by IPA decomposition during both femtosecond and nanosecond laser excitation of [AuCl4]- is attributed to enhanced optical breakdown by the Au nanoparticle products of [AuCl4]- reduction. These mechanistic insights can inform the design of laser synthesis procedures to improve control over metal nanoparticle properties and enhance byproduct yields.

2.
J Am Chem Soc ; 144(22): 9684-9694, 2022 06 08.
Artigo em Inglês | MEDLINE | ID: mdl-35609235

RESUMO

The stepwise hydration of the benzonitrile•+ radical cation with one-seven H2O molecules was investigated experimentally and computationally with density functional theory in C6H5CN•+(H2O)n clusters. The stepwise binding energies (ΔHn-1,n°) were determined by equilibrium measurements for C6H5CN•+(H2O) and for •C6H4CNH+(H2O)n with n = 5, 6, and 7 to be 8.8 and 11.3, 11.0, and 10.0 kcal/mol, respectively. The populations of n = 2 and 3 of the C6H5CN•+(H2O)n clusters were observed only in trace abundance due to fast depletion processes leading to the formation of the hydrated distonic cations •C6H4CNH+(H2O)n with n = 4-7. The observed transition occurs between conventional radical cations hydrated on the ring in C6H5CN•+(H2O)n clusters with n = 1-3 and the protonated radical •C6H4CNH+ (distonic ion) formed by a proton transfer to the CN nitrogen and ionic hydrogen bonding to water molecules in •C6H4CNH+(H2O)n clusters with n = 4-7. The measured binding energy of the hydrated ion C6H5CN•+(H2O) (8.8 kcal/mol) is similar to that of the hydrated benzene radical cation (8.5 kcal/mol) that involves a relatively weak CHδ+···O hydrogen bonding interaction. Also, the measured binding energies of the •C6H4CNH+(H2O)n clusters with n = 5-7 are similar to those of the protonated benzonitrile (methanol)n clusters [C6H5CNH+(CH3OH)n, n = 5-7] that involve CNH+···O ionic hydrogen bonds. The proton shift from the para-•C ring carbon to the nitrogen of the benzonitrile radical cation is endothermic without solvent but thermoneutral for n = 1 and exothermic for n = 2-4 in C6H5CN•+(H2O)n clusters to form the distonic •C6H4CN···H+(OH2)n clusters. The distonic clusters •C6H4CN···H+(OH2)n constitute a new class of structures in radical ion/solvent clusters.


Assuntos
Prótons , Água , Cátions/química , Radicais Livres/química , Hidrogênio , Nitrilas , Nitrogênio , Solventes , Água/química
3.
J Phys Chem A ; 126(31): 5160-5166, 2022 Aug 11.
Artigo em Inglês | MEDLINE | ID: mdl-35917155

RESUMO

The gas-phase sequential association of methanol onto protonated benzonitrile (C6H5CNH+) and the proton-bound dimer (C6H5CN)2H+ have been examined experimentally by equilibrium thermochemical measurements and computationally by density functional theory (DFT). The bonding enthalpy (ΔH°) for the association of methanol with protonated benzonitrile (25.2 kcal mol-1) reflects the strong electrostatic interaction provided by the formation of an ionic hydrogen bond in the C6H5CNH+OHCH3 cluster in excellent agreement with a DFT-calculated binding energy of 24.9 kcal mol-1. The sequential bonding enthalpy within the (C6H5CN)H+(OHCH3)n clusters decreases from 25.2 to 10.6 kcal mol-1 for the eighth solvation step (n = 8), which remains more than 25% above the enthalpy of vaporization of liquid methanol (8.4 kcal mol-1). The nonbulk convergence of ΔH°n-1,n with eight solvent molecules is attributed to the external solvation of a benzonitrile molecule by an extended hydrogen bonding network of protonated methanol clusters H+(CH3OH)n. In the external solvation of protonated benzonitrile by methanol, the proton resides on the methanol subcluster and the neutral benzonitrile molecule remains outside and bonded to the surface of the protonated methanol cluster. The bonding enthalpy of methanol to the proton-bound benzonitrile dimer (C6H5CN)H+(NCC6H5) is measured to be 18.0 kcal mol-1, in good agreement with a DFT-calculated value of 17.1 kcal mol-1, which reflects the association of the proton with the lower proton affinity methanol molecule, thus forming a highly stable structure of protonated methanol terminated by two ionic hydrogen bonds to the two benzonitrile molecules. The external solvation of benzonitrile by methanol ices in space allows benzonitrile to remain on the ice grain surface rather than being isolated inside the ice. This could provide accessibility for reactions with incoming ions and molecules or for photochemical processes by UV irradiation, leading to the formation of complex organics on the surface of ice grains.


Assuntos
Metanol , Prótons , Hidrogênio , Ligação de Hidrogênio , Metanol/química , Nitrilas , Termodinâmica
4.
J Environ Manage ; 312: 114956, 2022 Jun 15.
Artigo em Inglês | MEDLINE | ID: mdl-35398697

RESUMO

The presence of high concentrations of selenium ions in wastewater is considered an environmental problem. However, the mechanism of selenium ions (Se (IV)) removal by the adsorption process has not been investigated in-depth so far. Also, the recovery and conversion of the industrial waste materials into valuable materials is a vital issue. Therefore, in this study, zinc ferrite nanopowders are economically synthesized from steel-making wastes by co-precipitation method for investigating as adsorbents of selenium species. The produced nanopowders were annealed at 150, 300, 500, and 850 °C for 5 h to scrutinize the impact of annealing temperature on their crystallite size. The compositional, optical, and magnetic features of the nanopowders were defined by X-ray diffraction (XRD), field emission scanning electron microscopy (FESEM) and transmission electron microscopy (TEM), UV-Vis. spectrophotometer along with vibrating sample magnetometer (VSM). Optical absorbance spectra were found characteristic due to the electronic structure of Fe3+ (3d5) considering the C3v local symmetry of Fe3+ ions. The prepared nanopowders demonstrated good adsorption capacity toward selenium ions (43.67 mg/g at pH 2.5) from an aqueous medium. Adsorption data were found fitting to Freundlich isotherm model. Thus, ZnFe2O4 can be recommended to effectively eliminate selenium ions from aqueous solutions.


Assuntos
Nanopartículas , Selênio , Poluentes Químicos da Água , Adsorção , Compostos Férricos , Resíduos Industriais , Íons , Águas Residuárias/análise , Água , Poluentes Químicos da Água/química , Zinco
5.
J Am Chem Soc ; 143(10): 4064-4073, 2021 Mar 17.
Artigo em Inglês | MEDLINE | ID: mdl-33661615

RESUMO

Platinum (Pt)-based-nanomaterials are currently the most successful catalysts for the oxygen reduction reaction (ORR) in electrochemical energy conversion devices such as fuel cells and metal-air batteries. Nonetheless, Pt catalysts have serious drawbacks, including low abundance in nature, sluggish kinetics, and very high costs, which limit their practical applications. Herein, we report the first rationally designed nonprecious Co-Cu bimetallic metal-organic framework (MOF) using a low-temperature hydrothermal method that outperforms the electrocatalytic activity of Pt/C for ORR in alkaline environments. The MOF catalyst surpassed the ORR performance of Pt/C, exhibiting an onset potential of 1.06 V vs RHE, a half-wave potential of 0.95 V vs RHE, and a higher electrochemical stability (ΔE1/2 = 30 mV) after 1000 ORR cycles in 0.1 M NaOH. Additionally, it outperformed Pt/C in terms of power density and cyclability in zinc-air batteries. This outstanding behavior was attributed to the unique electronic synergy of the Co-Cu bimetallic centers in the MOF network, which was revealed by XPS and PDOS.

6.
Phys Chem Chem Phys ; 22(33): 18294-18303, 2020 Sep 07.
Artigo em Inglês | MEDLINE | ID: mdl-32785346

RESUMO

Laser photoreduction of metal ions onto graphene oxide (GO) is a facile, environmentally friendly method to produce functional metal-GO nanocomposites for a variety of applications. This work compares Au-GO nanocomposites prepared by photoreduction of [AuCl4]- in aqueous GO solution using laser pulses of nanosecond (ns) and femtosecond (fs) duration. The presence of GO significantly accelerates the [AuCl4]- photoreduction rate, with a more pronounced effect using ns laser pulses. This difference is rationalized in terms of the stronger interaction of the 532 nm laser wavelength and long pulse duration with the GO. Both the ns and fs lasers produce significant yields of sub-4 nm Au nanoparticles attached to GO, albeit with different size distributions: a broad 5.8 ± 1.9 nm distribution for the ns laser and two distinct distributions of 3.5 ± 0.8 and 10.1 ± 1.4 nm for the fs laser. Despite these differences, both Au-GO nanocomposites had the same high catalytic activity towards p-nitrophenol reduction as compared to unsupported 4-5 nm Au nanoparticles. These results point to the key role of GO photoexcitation in catalyzing metal ion reduction and indicate that both ns and fs lasers are suitable for producing functional metal-GO nanocomposites.

7.
J Phys Chem A ; 123(7): 1363-1371, 2019 Feb 21.
Artigo em Inglês | MEDLINE | ID: mdl-30452257

RESUMO

We report on the gas phase association of the small polar and aprotic solvent molecules acetonitrile (CH3CN) and acetone (CH3COCH3) with the halogenated benzene radical cations (C6H5X•+, X = F, Cl, Br, and I) using the mass-selected ion mobility technique and density functional theory calculations. The association energies (-Δ H°) of CH3CN (CH3COCH3) with C6H5F•+ and C6H5I•+ are similar [13.0 (13.3) and 13.2 (14.1) kcal/mol, respectively] but higher than those of CH3CN (CH3COCH3) with C6H5Cl•+ and C6H5Br•+ [10.5 (11.5) and 10.9 (10.6) kcal/mol, respectively]. However, the electrostatic potentials of the lowest energy structures of C6H5Br•+(CH3CN) and C6H5Br•+(CH3COCH3) or C6H5I•+(CH3CN) and C6H5I•+(CH3COCH3) complexes clearly show the formation of the ionic halogen bonds (IXBs) C-Brδ+- -NCCH3 and C-Brδ+- -OC(CH3)2 or C-Iδ+- -NCCH3 and C-Iδ+- -OC(CH3)2 driven by positively charged σ-holes on the external sides of the C-Br and C-I bond axes of the bromobenzene and iodobenzene radical cations, respectively. For the C6H5F•+(CH3CN) complex, the dominant interaction involves a T-shaped structure between the N atom of CH3CN and the C atom of the C-F bond of C6H5F•+. The structure of the C6H5Cl•+(CH3CN) complex shows the formation of unconventional ionic hydrogen bonds (uIHBs) between the N atom of CH3CN and the C-H bonds of the C6H5Cl•+ cation. Similar results are obtained for the association of acetone with the halogenated benzene radical cations. The formation of IXBs of the iodobenzene cation with acetonitrile or acetone involves a significant entropy loss (-Δ S° = 25-27 cal /(mol K)) resulting from the formation of more ordered and highly directional structures between the nitrogen or oxygen lone pair of electrons of acetonitrile or acetone, respectively, and the electropositive region around the iodine atom of the iodobenzene cation. In comparison, for the association of acetonitrile or acetone with the fluorobenzene, chlorobenzene, and bromobenzene cations, -Δ S° = 16-23 cal/(mol K), consistent with the formation of less ordered structures and loose interactions. The lowest energy structures of the C6H5Br•+(CH3COCH3)2 and C6H5I•+(CH3COCH3)2 clusters show a novel combination of ionic halogen bonding and hydrogen bonding where the oxygen atom of one acetone molecule forms the halogen bond while the oxygen atom of the second acetone molecule becomes the hydrogen acceptor from the methyl group of the first acetone molecule.

8.
J Chem Phys ; 150(12): 124303, 2019 Mar 28.
Artigo em Inglês | MEDLINE | ID: mdl-30927904

RESUMO

The recent discovery of benzonitrile (C6H5CN), one of the simplest nitrogen-bearing polar aromatic molecules, in the interstellar medium motivates structural characterization of the benzonitrile-containing molecular ions as potential precursors for nitrogen-containing complex organics in space. Herein, we present mass-selected ion mobility measurements combined with density functional theory (DFT) calculations to reveal, for the first time, the structures of the benzonitrile dimer radical cation, the protonated dimer, and the protonated hydrated small clusters in the gas phase. The measured collision cross sections of the investigated ions in helium are in excellent agreement with the calculated values of the lowest energy DFT structures. Unlike the dimer radical cations of nonpolar aromatic molecules which adopt parallel sandwich configurations, the (C6H5CN)2 ·+ displays a symmetrically planar geometry with a double hydrogen bond formed between the nitrogen and hydrogen atoms. The protonated dimer has the structure of a proton-bound dimer (C6H5CNH+NCC6H5) where the bridging proton connects the nitrogen atoms in the two benzonitrile molecules resulting in a calculated collision cross section of 101.1 Å2 in excellent agreement with the measured value of 103.3 Å2. The structure of the hydrated protonated trimer consists of a hydronium ion core solvated by three benzonitrile molecules. By locating the proton on the lower proton affinity water molecule, the resulting hydronium ion can be fully solvated by forming three ionic hydrogen bonds with the benzonitrile molecules. These unique structural motifs could be useful for the molecular design and recognition involving charged aromatic systems and also for the search of nitrogen-containing complex organics in space.

9.
Phys Chem Chem Phys ; 20(45): 28465-28475, 2018 Nov 21.
Artigo em Inglês | MEDLINE | ID: mdl-30411753

RESUMO

Irradiation of aqueous [AuCl4]- with 532 nm nanosecond (ns) laser pulses produces monodisperse (PDI = 0.04) 5 nm Au nanoparticles (AuNPs) without any additives or capping agents via a plasmon-enhanced photothermal autocatalytic mechanism. Compared with 800 nm femtosecond (fs) laser pulses, the AuNP growth kinetics under ns laser irradiation follow the same autocatalytic rate law, but with a significantly lower sensitivity to laser pulse energy. The results are explained using a simple model for simulating heat transfer in liquid water and at the interface with AuNPs. While the extent of water superheating with the ns laser is smaller compared to the fs laser, its significantly longer duration can provide sufficient energy to dissociate a small fraction of the [AuCl4]- present, resulting in the formation of AuNPs by coalescence of the resulting Au atoms. Irradiation of initially formed AuNPs at 532 nm results in plasmon-enhanced superheating of water, which greatly accelerates the rate of thermal dissociation of [AuCl4]- and accounts for the observed autocatalytic kinetics. The plasmon-enhanced heating under ns laser irradiation fragments the AuNPs and results in nearly uniform 5 nm particles, while the lack of particles' heating under fs laser irradiation results in the growth of the particles as large as 40 nm.

10.
J Am Chem Soc ; 139(34): 11923-11932, 2017 08 30.
Artigo em Inglês | MEDLINE | ID: mdl-28759221

RESUMO

Radical organic ions can be stabilized by complexation with neutral organics via interactions that can resemble chemical bonds, but with much diminished bond energies. Those interactions are a key factor in cluster growth and polymerization reactions in ionizing environments such as regions of the interstellar medium and solar nebulae. Such radical cation complexes between naphthalene (Naph) and pyridine (Pyr) are characterized using mass-selected ion mobility experiments. The measured enthalpy of binding of the Naph+•(Pyr) heterodimer (20.9 kcal/mol) exceeds that of the Naph+•(Naph) homodimer (17.8 kcal/mol). The addition of 1-3 more pyridine molecules to the Naph+•(Pyr) heterodimer gives 10-11 kcal/mol increments in binding enthalpy. A rich array of Naph+•(Pyr) isomers are characterized by electronic structure calculations. The calculated Boltzmann distribution at 400 K yields an enthalpy of binding in reasonable agreement with experiment. The global minimum is a distonic cation formed by Pyr attack on Naph+• at the α-carbon, changing its hybridization from sp2 to distorted sp3. The measured collision cross section in helium for the Naph+•(Pyr) heterodimer of 84.9 ± 2.5 Å2 at 302 K agrees well with calculated angle-averaged cross sections (83.9-85.1 Å2 at 302 K) of the lowest energy distonic structures. A remarkable 16 kcal/mol increase in the binding energy between Naph+•(Pyr) and Bz+•(Pyr) (Bz is benzene) is understood by energy decomposition analysis. A similar increase in binding from Naph+•(NH3) to Naph+•(Pyr) (as well as between Bz+•(NH3) and Bz+•(Pyr)) is likewise rationalized.

11.
Phys Chem Chem Phys ; 19(28): 18603-18611, 2017 Jul 19.
Artigo em Inglês | MEDLINE | ID: mdl-28691724

RESUMO

Halogen bonding (XB) non-covalent interactions can be observed in compounds containing chlorine, bromine, or iodine which can form directed close contacts of the type R1-XY-R2, where the halogen X acts as a Lewis acid and Y can be any electron donor moiety including electron lone pairs on hetero atoms such as O and N, or π electrons in olefin double bonds and aromatic conjugated systems. In this work, we present the first evidence for the formation of ionic halogen bonds (IXBs) in the hydration of bromobenzene and iodobenzene radical cations in the gas phase. We present a combined thermochemical investigation using the mass-selected ion mobility (MSIM) technique and density functional theory (DFT) calculations of the stepwise hydration of the fluoro, chloro, bromo, and iodobenzene radical cations. The binding energy associated with the formation of an IXB in the hydration of the iodobenzene cation (11.2 kcal mol-1) is about 20% higher than the typical unconventional ionic hydrogen bond (IHB) of the CHδ+OH2 interaction. The formation of an IXB in the hydration of the iodobenzene cation involves a significant entropy loss (29 cal mol-1 K-1) resulting from the formation of a more ordered structure and a highly directional interaction between the oxygen lone pair of electrons of water and the electropositive region around the iodine atom of the iodobenzene cation. In comparison, the hydration of the fluorobenzene and chlorobenzene cations where IHBs are formed, -ΔS° = 18-21 cal mol-1 K-1 consistent with the formation of less ordered structures and loose interactions. The electrostatic potentials on the lowest energy structures of the hydrated halogenated benzene radical cations show clearly that the formation of an IXB is driven by a positively charged σ-hole on the external side of the halogen atom X along the C-X bond axis. The size of the σ-hole increases significantly in bromobenzene and iodobenzene radical cations which results in strong interaction potentials with the electron lone pairs of the oxygen atom of the water molecules and thus IXBs provide the most stable hydrated structures of the bromobenzene and iodobenzene radical cations. The results clearly distinguish the hydration behaviors resulting from the ionic hydrogen and halogen bonding interactions of fluorobenzene and iodobenzene cations, respectively, and establish the different bonding and structural features of the two interactions.

12.
Phys Chem Chem Phys ; 18(41): 28606-28616, 2016 Oct 19.
Artigo em Inglês | MEDLINE | ID: mdl-27711723

RESUMO

We report a detailed experimental and theoretical study of the gas phase reactivity of Au+ with CO, O2, N2 and their mixtures in the presence of a trace amount of water impurity. The gold cation is found to strongly interact with CO and H2O molecules via successive addition reactions until reaching saturation. The stoichiometry of the formed complex is determined by the strength of the binding energy of the neutral molecule to the gold cation. CO binds the strongest to Au+, followed by H2O, N2 and then O2. We found that the gold cation (Au+) can activate the O2 molecule within the Au+(CO)2(O2) complex which could react with another CO molecule to form Au+(CO)(CO2) + CO2. The product Au+(CO)(CO2) is observed experimentally with a small intensity at room temperature. However, the presence of water leads to the formation of Au+(CO)(H2O)(O2) instead of Au+(CO)2(O2) due to the strong interaction between Au+ and water. The current experiments and calculations might lead to a molecular level understanding of the interactions between the active sites, reactants and impurities which could pave the way for the design of efficient nanocatalysts.

13.
J Chem Phys ; 142(19): 191102, 2015 May 21.
Artigo em Inglês | MEDLINE | ID: mdl-26001439

RESUMO

Dimer radical cations of aromatic and polycyclic aromatic molecules are good model systems for a fundamental understanding of photoconductivity and ferromagnetism in organic materials which depend on the degree of charge delocalization. The structures of the dimer radical cations are difficult to determine theoretically since the potential energy surface is often very flat with multiple shallow minima representing two major classes of isomers adopting the stacked parallel or the T-shape structure. We present experimental results, based on mass-selected ion mobility measurements, on the gas phase structures of the naphthalene(+⋅) ⋅ naphthalene homodimer and the naphthalene(+⋅) ⋅ benzene heterodimer radical cations at different temperatures. Ion mobility studies reveal a persistence of the stacked parallel structure of the naphthalene(+⋅) ⋅ naphthalene homodimer in the temperature range 230-300 K. On the other hand, the results reveal that the naphthalene(+⋅) ⋅ benzene heterodimer is able to exhibit both the stacked parallel and T-shape structural isomers depending on the experimental conditions. Exploitation of the unique structural motifs among charged homo- and heteroaromatic-aromatic interactions may lead to new opportunities for molecular design and recognition involving charged aromatic systems.

14.
Langmuir ; 30(8): 2230-40, 2014 Mar 04.
Artigo em Inglês | MEDLINE | ID: mdl-24555702

RESUMO

We report herein the development of a facile microwave irradiation (MWI) method for the synthesis of high-quality N-doped ZnO nanostructures with controlled morphology and doping level. We present two different approaches for the MWI-assisted synthesis of N-doped ZnO nanostructures. In the first approach, N-doping of Zn-poor ZnO prepared using zinc peroxide (ZnO2) as a precursor is carried out under MWI in the presence of urea as a nitrogen source and oleylamine (OAm) as a capping agent for the shape control of the resulting N-doped ZnO nanostructures. Our approach utilizes the MWI process for the decomposition of ZnO2, where the rapid transfer of energy directly to ZnO2 can cause an instantaneous internal temperature rise and, thus, the activation energy for the ZnO2 decomposition is essentially decreased as compared to the decomposition under conductive heating. In the second synthesis method, a one-step synthesis of N-doped ZnO nanostructures is achieved by the rapid decomposition of zinc acetate in a mixture of urea and OAm under MWI. We demonstrate, for the first time, that MWI decomposition of zinc acetate in a mixture of OAm and urea results in the formation of N-doped nanostructures with controlled shape and N-doping level. We report a direct correlation between the intensity of the Raman scattering bands in N-doped ZnO and the concentration of urea used in the synthesis. Electrochemical measurements demonstrate the successful synthesis of stable p-type N-doped ZnO nanostructures using the one-step MWI synthesis and, therefore, allow us to investigate, for the first time, the relationship between the doping level and morphology of the ZnO nanostructures. The results provide strong evidence for the control of the electrical behavior and the nanostructured shapes of ZnO nanoparticles using the facile MWI synthesis method developed in this work.

15.
J Phys Chem A ; 118(37): 8251-63, 2014 Sep 18.
Artigo em Inglês | MEDLINE | ID: mdl-24689826

RESUMO

Here we report a detailed study aimed at elucidating the mechanism of intracluster ionic polymerization following the electron impact ionization of van der Waals clusters of ethynylbenzene (C8H6)n generated by a supersonic beam expansion. The structures of the C16H12, C24H18, C32H24, C40H30, and C48H36 radical cations resulting from the intracluster ion-molecule addition reactions have been investigated using a combination of mass-selected ion dissociation and ion mobility measurements coupled with theoretical calculations. Noncovalent structures can be totally excluded primarily because the measured fragmentations cannot result from noncovalent structures, and partially because of the large difference between the measured collision cross sections and the calculated values corresponding to noncovalent ion-neutral complexes. All the mass-selected cluster ions show characteristic fragmentations of covalently bonded molecular ions by the loss of stable neutral fragments such as CH3, C2H, C6H5, and C7H7. The population of the C16H12 dimer ions is dominated by structural isomers of the type (C6H5)-C≡C-CH(•+)CH-(C6H5), which can grow by the sequential addition of ethynylbenzene molecules, in addition to some contributions from cyclic isomers such as the 1,3- or 1,4-diphenyl cyclobutadiene ions. Similarly, two major covalent isomers have been identified for the C24H18 trimer ions: one that has a blocked cyclic structure assigned to 1,2,4- or 1,3,5-triphenylbenzene cation, and a second isomer of the type (C6H5)-C≡C-C(C6H5)═CH-CH(•+)CH-(C6H5) where the covalent addition of further ethynylbenzene molecules can occur. For the larger ions such as C32H24, C40H30, and C48H36, the major isomers present involve the growing oligomer sequence (C6H5)-C≡C-[C(C6H5)═CH]n-CH(•+)CH-(C6H5) with different locations and orientations of the phenyl groups along the chain. In addition, the larger ions contain another family of structures consisting of neutral ethynylbenzene molecules associated with the blocked cyclic isomer ions such as the diphenylcyclobutadiene and triphenylbenzene cations. Low-energy dissociation channels corresponding to evaporation of ethynylbenzene molecules weakly associated with the covalent ions are observed in the large clusters in addition to the high-energy channels corresponding to fragmentation of the covalently bonded ions. However, in small clusters only high-energy dissociation channels are observed corresponding to the characteristic fragmentation of the molecular ions, thus providing structural signatures to identify the product ions and establish the mechanism of intracluster ionic polymerization.

16.
J Chem Phys ; 141(5): 054710, 2014 Aug 07.
Artigo em Inglês | MEDLINE | ID: mdl-25106603

RESUMO

This work reports the direct observation and separation of size-selected aluminum nanoparticles acting as heterogeneous nuclei for the condensation of supersaturated vapors of both polar and nonpolar molecules. In the experiment, we study the condensation of supersaturated acetonitrile and n-hexane vapors on charged and neutral Al nanoparticles by activation of the metal nanoparticles to act as heterogeneous nuclei for the condensation of the organic vapor. Aluminum seed nanoparticles with diameters of 1 and 2 nm are capable of acting as heterogeneous nuclei for the condensation of supersaturated acetonitrile and hexane vapors. The comparison between the Kelvin and Fletcher diameters indicates that for the heterogeneous nucleation of both acetonitrile and hexane vapors, particles are activated at significantly smaller sizes than predicted by the Kelvin equation. The activation of the Al nanoparticles occurs at nearly 40% and 65% of the onset of homogeneous nucleation of acetonitrile and hexane supersaturated vapors, respectively. The lower activation of the charged Al nanoparticles in acetonitrile vapor is due to the charge-dipole interaction which results in rapid condensation of the highly polar acetonitrile molecules on the charged Al nanoparticles. The charge-dipole interaction decreases with increasing the size of the Al nanoparticles and therefore at low supersaturations, most of the heterogeneous nucleation events are occurring on neutral nanoparticles. No sign effect has been observed for the condensation of the organic vapors on the positively and negatively charged Al nanoparticles. The present approach of generating metal nanoparticles by pulsed laser vaporization within a supersaturated organic vapor allows for efficient separation between nucleation and growth of the metal nanoparticles and, consequently controls the average particle size, particle density, and particle size distribution within the liquid droplets of the condensing vapor. Strong correlation is found between the seed nanoparticle's size and the degree of the supersaturation of the condensing vapor. This result and the agreement among the calculated Kelvin diameters and the size of the nucleating Al nanoparticles determined by transmission electron microscopy provide strong proof for the development of a new approach for the separation and characterization of heterogeneous nuclei formed in organic vapors. These processes can take place in the atmosphere by a combination of several organic species including polar compounds which could be very efficient in activating charged nanoparticles and cluster ions of atmospheric relevance.

17.
J Chem Phys ; 141(5): 054305, 2014 Aug 07.
Artigo em Inglês | MEDLINE | ID: mdl-25106585

RESUMO

Equilibrium thermochemical measurements using the ion mobility drift cell technique have been utilized to investigate the binding energies and entropy changes for the stepwise association of HCN molecules with the pyridine and pyrimidine radical cations forming the C5H5N(+·)(HCN)n and C4H4N2 (+·)(HCN)n clusters, respectively, with n = 1-4. For comparison, the binding of 1-4 HCN molecules to the protonated pyridine C5H5NH(+)(HCN)n has also been investigated. The binding energies of HCN to the pyridine and pyrimidine radical cations are nearly equal (11.4 and 12.0 kcal/mol, respectively) but weaker than the HCN binding to the protonated pyridine (14.0 kcal/mol). The pyridine and pyrimidine radical cations form unconventional carbon-based ionic hydrogen bonds with HCN (CH(δ+)⋯NCH). Protonated pyridine forms a stronger ionic hydrogen bond with HCN (NH(+)⋯NCH) which can be extended to a linear chain with the clustering of additional HCN molecules (NH(+)⋯NCH··NCH⋯NCH) leading to a rapid decrease in the bond strength as the length of the chain increases. The lowest energy structures of the pyridine and pyrimidine radical cation clusters containing 3-4 HCN molecules show a strong tendency for the internal solvation of the radical cation by the HCN molecules where bifurcated structures involving multiple hydrogen bonding sites with the ring hydrogen atoms are formed. The unconventional H-bonds (CH(δ+)⋯NCH) formed between the pyridine or the pyrimidine radical cations and HCN molecules (11-12 kcal/mol) are stronger than the similar (CH(δ+)⋯NCH) bonds formed between the benzene radical cation and HCN molecules (9 kcal/mol) indicating that the CH(δ+) centers in the pyridine and pyrimidine radical cations have more effective charges than in the benzene radical cation.


Assuntos
Gases/química , Cianeto de Hidrogênio/química , Modelos Químicos , Compostos Orgânicos/química , Piridinas/química , Pirimidinas/química , Espectrometria de Massas por Ionização por Electrospray/métodos , Sítios de Ligação , Cátions , Simulação por Computador , Radicais Livres/química , Ligação de Hidrogênio , Modelos Moleculares , Transição de Fase , Prótons
18.
J Am Chem Soc ; 135(1): 155-66, 2013 Jan 09.
Artigo em Inglês | MEDLINE | ID: mdl-23205891

RESUMO

Here, we present evidence from laboratory experiments for the formation of nitrogen-containing complex organic ions by sequential reactions of acetylene with the pyridinium and pyrimidinium ions in the gas phase and within ionized pyridine-acetylene binary clusters. Additions of five and two acetylene molecules onto the pyridinium and pyrimidinium ions, respectively, at room temperature are observed. Second-order rate coefficients of the overall reaction of acetylene with the pyridinium and pyrimidinium ions are measured as 9.0 × 10(-11) and 1.4 × 10(-9) cm(3) s(-1), respectively, indicating reaction efficiencies of about 6% and 100%, respectively, at room temperature. At high temperatures, only two acetylene molecules are added to the pyridinium and pyrimidinium ions, suggesting covalent bond formation. A combination of ion dissociation and ion mobility experiments with DFT calculations reveals that the addition of acetylene into the pyridinium ion occurs through the N-atom of the pyridinium ion. The relatively high reaction efficiency is consistent with the absence of a barrier in the exothermic N-C bond forming reaction leading to the formation of the C(7)H(7)N(•+) covalent adduct. An exothermic addition/H-elimination reaction of acetylene with the C(7)H(7)N(•+) adduct is observed leading to the formation of a bicyclic quinolizinium cation (C(9)H(8)N(+)). Similar chemistry is observed in the sequential reactions of acetylene with the pyrimidinium ion. The second acetylene addition onto the pyrimidinium ion involves an exclusive addition/H-elimination reaction at room temperature leading to the formation of a bicyclic pyrimidinium cation (C(8)H(7)N(2)(+)). The high reactivity of the pyridinium and pyrimidinium ions toward acetylene is in sharp contrast to the very low reactivity of the benzene cation, which has a reaction efficiency of 10(-4)-10(-5). This indicates that the presence of a nitrogen atom within the aromatic ring enhances the ring growth mechanism by the sequential addition of acetylene to form nitrogen-containing polycyclic hydrocarbon ions. The observed reactions could explain the possible formation of nitrogen-containing complex organics by gas-phase ion-molecule reactions involving the pyridinium and pyrimidinium ions with acetylene under a wide range of temperatures and pressures in astrochemical environments such as the nitrogen-rich Titan's atmosphere. The current results suggest searching for spectroscopic evidence for these organics in Titan's atmosphere.


Assuntos
Acetileno/química , Nitrogênio/química , Compostos Policíclicos/síntese química , Compostos de Piridínio/química , Pirimidinas/química , Cátions/síntese química , Cátions/química , Gases/química , Íons/química , Compostos Policíclicos/química , Teoria Quântica
19.
J Phys Chem A ; 117(6): 1069-78, 2013 Feb 14.
Artigo em Inglês | MEDLINE | ID: mdl-22671581

RESUMO

Equilibrium thermochemical measurements using the ion mobility drift cell technique have been utilized to investigate the binding energies and entropy changes associated with the stepwise association of HCN and CH(3)CN molecules with the benzene radical cation in the C(6)H(6)(•+)(HCN)(n) and C(6)H(6)(•+)(CH(3)CN)(n) clusters with n = 1-6 and 1-4, respectively. The binding energy of CH(3)CN to the benzene cation (14 kcal/mol) is stronger than that of HCN (9 kcal/mol) mostly due to a stronger ion-dipole interaction because of the large dipole moment of acetonitrile (3.9 D). However, HCN can form hydrogen bonds with the hydrogen atoms of the benzene cation (CH(δ+)···NCH) and linear hydrogen bonding chains involving HCN···HCN interaction. HCN molecules tend to form externally solvated structures with the benzene cation where the ion is hydrogen bonded to the exterior of HCN chains. For the C(6)H(6)(•+)(CH(3)CN)(n) clusters, internally solvated structures are formed where the acetonitrile molecules are directly interacting with the benzene cation through ion-dipole and hydrogen bonding interactions. The lack of formation of higher clusters with n > 4, in contrast to HCN, suggests the formation of a solvent shell at n = 4, which is attributed to steric interactions among the acetonitrile molecules attached to the benzene cation and to the presence of the blocking CH(3) groups, both effects make the addition of more than four acetonitrile molecules less favorable.


Assuntos
Acetonitrilas/química , Benzeno/química , Cianeto de Hidrogênio/química , Termodinâmica , Cátions/química , Radicais Livres/química , Estrutura Molecular , Teoria Quântica
20.
J Chem Phys ; 139(8): 084304, 2013 Aug 28.
Artigo em Inglês | MEDLINE | ID: mdl-24006994

RESUMO

Equilibrium thermochemical measurements using an ion mobility drift cell technique have been utilized to investigate the binding energies and entropy changes associated with the stepwise hydration of the biologically significant ions pyrimidine radical cation and protonated pyrimidine. The binding energy of the hydrated pyrimidine radical cation is weaker than that of the proton-bound dimer pyrimidineH(+)(H2O) consistent with the formation of a weak carbon-based CH(δ+)··OH2 hydrogen bond (11.9 kcal/mol) and a stronger NH(+)··OH2 hydrogen bond (15.6 kcal/mol), respectively. Other proton-bound dimers such as pyrimidineH(+)(CH3OH) and pyrimidineH(+)(CH3CN) exhibit higher binding energies (18.2 kcal/mol and 22.8 kcal/mol, respectively) due to the higher proton affinities and dipole moments of acetonitrile and methanol as compared to water. The measured collisional cross sections of the proton-bound dimers provide experimental-based support for the DFT calculated structures at the M06-2x/6-311++G (d,p) level. The calculations show that the hydrated pyrimidine radical cation clusters form internally solvated structures in which the water molecules are bonded to the C4N2H4(●+) ion by weak CH(δ+)··OH2 hydrogen bonds. The hydrated protonated pyrimidine clusters form externally solvated structures where the water molecules are bonded to each other and the ion is external to the water cluster. Dissociative proton transfer reactions C4N2H4(●+)(H2O)(n-1) + H2O → C4N2H3(●) + (H2O)(n)H(+) and C4N2H5(+)(H2O)(n-1) + H2O → C4N2H4 + (H2O)(n)H(+) are observed for n ≥ 4 where the reactions become thermoneutral or exothermic. The absence of the dissociative proton transfer reaction within the C4N2H5(+)(CH3CN)n clusters results from the inability of acetonitrile molecules to form extended hydrogen bonding structures such as those formed by water and methanol due to the presence of the methyl groups which block the extension of hydrogen bonding networks.


Assuntos
Acetonitrilas/química , Metanol/química , Pirimidinas/química , Água/química , Cátions/química , Radicais Livres/química , Prótons , Teoria Quântica , Solubilidade
SELEÇÃO DE REFERÊNCIAS
DETALHE DA PESQUISA