Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 16 de 16
Filtrar
Mais filtros










Base de dados
Intervalo de ano de publicação
1.
Langmuir ; 40(13): 6612-6653, 2024 Apr 02.
Artigo em Inglês | MEDLINE | ID: mdl-38509763

RESUMO

Closed-loop recycling via an efficient chemical process can help alleviate the global plastic waste crisis. However, conventional depolymerization methods for polyolefins, which compose more than 50% of plastics, demand high temperatures and pressures, employ precious noble metals, and/or yield complex mixtures of products limited to single-use fuels or oils. Superacidic forms of sulfated zirconia (SZrO) with Hammet Acidity Functions (H0) ≤ - 12 (i.e., stronger than 100% H2SO4) are industrially deployed heterogeneous catalysts capable of activating hydrocarbons under mild conditions and are shown to decompose polyolefins at temperatures near 200 °C and ambient pressure. Additionally, confinement of active sites in porous supports is known to radically increase selectivity, coking and sintering resistance, and acid site activity, presenting a possible approach to low-energy polyolefin depolymerization. However, a critical examination of the literature on SZrO led us to a surprising conclusion: despite 40 years of catalytic study, engineering, and industrial use, the surface chemistry of SZrO is poorly understood. Ostensibly spurred by SZrO's impressive catalytic activity, the application-driven study of SZrO has resulted in deleterious ambiguity in requisite synthetic conditions for superacidity and insufficient characterization of acidity, porosity, and active site structure. This ambiguity has produced significant knowledge gaps surrounding the synthesis, structure, and mechanisms of hydrocarbon activation for optimized SZrO, stunting the potential of this catalyst in olefin cracking and other industrially relevant reactions, such as isomerization, esterification, and alkylation. Toward mitigating these long extant issues, we herein identify and highlight these current shortcomings and knowledge gaps, propose explicit guidelines for characterization of and reporting on characterization of solid acidity, and discuss the potential of pore-confined superacids in the efficient and selective depolymerization of polyolefins.

2.
ChemSusChem ; 15(12): e202200535, 2022 Jun 22.
Artigo em Inglês | MEDLINE | ID: mdl-35395145

RESUMO

Decomposition of polymers by heterogeneous catalysts presents a promising approach for reuse of waste plastics. We demonstrated non-hydrogenative decomposition of model polyolefins over proton-form and metal (Cu, Ni) ion-exchanged beta (BEA) zeolites at moderate temperatures (around 300 °C). Near complete polyolefin decomposition was observed in batch reactions monitored by thermogravimetric analysis, while decomposition at partial conversion was studied in flow reactions. Ni-exchanged zeolites produced H2 at substantially higher rates (>10x) than other catalysts while also uniquely resisting deactivation over time. Application of the delplot formalism offered insights into the reaction network for polyolefin decomposition over Ni/BEA most notably that H2 is solely a primary product. We deduce that H2 production is catalyzed by activation of C-H bonds at ionic Ni sites, and H2 prevents buildup of polyaromatic coke species in Ni-exchanged zeolites that deactivate Cu-exchanged and protonic zeolites.

3.
Anal Chem ; 93(10): 4351-4360, 2021 03 16.
Artigo em Inglês | MEDLINE | ID: mdl-33645225

RESUMO

To mitigate green-house gas (GHG) emissions, governments around the world are enacting legislation to reduce carbon intensity in transportation fuels. Coprocessing biomass and petroleum-derived liquids in existing refineries is a near-term, cost-effective approach for introducing renewable carbon in fuels and enabling refineries to meet regulatory mandates. However, coprocessing biomass-derived liquids in refineries results in variable degrees of biogenic carbon incorporation, necessitating accurate quantification to verify compliance with mandates. Existing refinery control and instrumentation systems lack the means to measure renewable carbon accurately, reliably, and quickly. Thus, accurate measurement of biogenic carbon is key to ensuring refineries meet regulatory mandates. In this Perspective, we present existing methods for measuring biogenic carbon, point out their challenges, and discuss the need for new online analytical capabilities to measure biogenic carbon in fuel intermediates.


Assuntos
Petróleo , Biomassa , Carbono , Pegada de Carbono , Efeito Estufa
4.
J Chem Phys ; 141(20): 204304, 2014 Nov 28.
Artigo em Inglês | MEDLINE | ID: mdl-25429939

RESUMO

Kinetic energy thresholds have been measured for the chemisorption of N2 onto Al114 (+), Al115 (+), and Al117 (+) as a function of the cluster's initial temperature, from around 200 K up to around 900 K. For all three clusters there is a sharp drop in the kinetic energy threshold of 0.5-0.6 eV at around 450 K, that is correlated with the structural transition identified in heat capacity measurements. The decrease in the thresholds corresponds to an increase in the reaction rate constant, k(T) at 450 K, of around 10(6)-fold. No significant change in the thresholds occurs when the clusters melt at around 600 K. This contrasts with behavior previously reported for smaller clusters where a substantial drop in the kinetic energy thresholds is correlated with the melting transition.

5.
J Am Chem Soc ; 132(37): 12906-18, 2010 Sep 22.
Artigo em Inglês | MEDLINE | ID: mdl-20738096

RESUMO

Cross sections for chemisorption of N2 onto Al44(+/-) cluster ions have been measured as a function of relative kinetic energy and the temperature of the metal cluster. There is a kinetic energy threshold for chemisorption, indicating that it is an activated process. The threshold energies are around 3.5 eV when the clusters are in their solid phase and drop to around 2.5 eV when the clusters melt, indicating that the liquid clusters are much more reactive than the solid. Below the melting temperature the threshold for Al44(-) is smaller than for Al44(+), but for the liquid clusters the anion and cation have similar thresholds. At high cluster temperatures and high collision energies the Al44N2(+/-) chemisorption product dissociates through several channels, including loss of Al, N2, and Al3N. Density functional calculations are employed to understand the thermodynamics and the dynamics of the reaction. The theoretical results suggest that the lowest energy pathway for activation of dinitrogen is not dynamically accessible under the experimental conditions, so that an explicit account of dynamical effects, via molecular dynamics simulations, is necessary in order to interpret the experimental measurements. The calculations reproduce all of the main features of the experimental results, including the kinetic energy thresholds of the anion and cation and the dissociation energies of the liquid Al44N2(+/-) product. The strong increase in reactivity on melting appears to be due to the volume change of melting and to atomic disorder.

6.
J Chem Phys ; 132(3): 034302, 2010 Jan 21.
Artigo em Inglês | MEDLINE | ID: mdl-20095734

RESUMO

Heat capacities have been measured as a function of temperature for isolated aluminum nanoclusters with 84-128 atoms. Most clusters show a single sharp peak in the heat capacity which is attributed to a melting transition. However, there are several size regimes where additional features are observed; for clusters with 84-89 atoms the peak in the heat capacity is either broad or bimodal. For Al(115) (+), Al(116) (+), and Al(117) (+) there are two well-defined peaks, and for Al(126) (+), Al(127) (+), and Al(128) (+) there is a dip in the heat capacity at lower temperature than the peak. The broad or bimodal peaks for clusters with 84-89 atoms are not significantly changed by annealing to 823 K (above the melting temperature), but the dips for Al(126) (+), Al(127) (+), and Al(128) (+) disappear when these clusters are annealed to 523 K (above the temperature of the dip but below the melting temperature). Both the melting temperatures and the latent heats change fairly smoothly with the cluster size in the size regime examined here. There are steps in the melting temperatures for clusters with around 100 and 117 atoms. The step at Al(100) (+) is correlated with a substantial peak in the latent heats but the step at Al(117) (+) correlates with a minimum. Since the latent heats are correlated with the cluster cohesive energies, the substantial peak in the latent heats at Al(100) (+) indicates this cluster is particularly strongly bound.

7.
J Chem Phys ; 131(12): 124305, 2009 Sep 28.
Artigo em Inglês | MEDLINE | ID: mdl-19791879

RESUMO

Heat capacities measured as a function of temperature for Al(115)(+), Al(116)(+), and Al(117)(+) show two well-resolved peaks, at around 450 and 600 K. After being annealed to 523 K (a temperature between the two peaks) or to 773 K (well above both peaks), the high temperature peak remains unchanged but the low temperature peak disappears. After considering the possible explanations, the low temperature peak is attributed to a structural transition and the high temperature peak to the melting of the higher enthalpy structure generated by the structural transition. The annealing results show that the liquid clusters freeze exclusively into the higher enthalpy structure and that the lower enthalpy structure is not accessible from the higher enthalpy one on the timescale of the experiments. We suggest that the low enthalpy structure observed before annealing results from epitaxy, where the smaller clusters act as a nucleus and follow a growth pattern that provides access to the low enthalpy structure. The solid-to-solid transition that leads to the low temperature peak in the heat capacity does not occur under equilibrium but requires a superheated solid.

8.
J Chem Phys ; 131(4): 044307, 2009 Jul 28.
Artigo em Inglês | MEDLINE | ID: mdl-19655867

RESUMO

Heat capacities have been measured as a function of temperature for aluminum cluster anions with 35-70 atoms. Melting temperatures and latent heats are determined from peaks in the heat capacities; cohesive energies are obtained for solid clusters from the latent heats and dissociation energies determined for liquid clusters. The melting temperatures, latent heats, and cohesive energies for the aluminum cluster anions are compared to previous measurements for the corresponding cations. Density functional theory calculations have been performed to identify the global minimum energy geometries for the cluster anions. The lowest energy geometries fall into four main families: distorted decahedral fragments, fcc fragments, fcc fragments with stacking faults, and "disordered" roughly spherical structures. The comparison of the cohesive energies for the lowest energy geometries with the measured values allows us to interpret the size variation in the latent heats. Both geometric and electronic shell closings contribute to the variations in the cohesive energies (and latent heats), but structural changes appear to be mainly responsible for the large variations in the melting temperatures with cluster size. The significant charge dependence of the latent heats found for some cluster sizes indicates that the electronic structure can change substantially when the cluster melts.

9.
J Chem Phys ; 130(20): 204303, 2009 May 28.
Artigo em Inglês | MEDLINE | ID: mdl-19485445

RESUMO

The internal energy distributions for melting aluminum cluster cations with 100, 101, 126, and 127 atoms have been investigated using multicollision induced dissociation. The experimental results can be best fit with a statistical thermodynamic model that incorporates only fully solidlike and fully liquidlike clusters so that the internal energy distributions become bimodal during melting. This result is consistent with computer simulations of small clusters, where rapid fluctuations between entirely solidlike and entirely liquidlike states occur during the phase change. To establish a bimodal internal energy distribution, the time between the melting and freezing transitions must be longer than the time required for equilibration of the energy distribution (which is estimated to be around 1-2 micros under our conditions). For Al(100)(+) and Al(101)(+), the results indicate that this criterion is largely met. However, for Al(126)(+) and Al(127)(+), it appears that the bimodal energy distributions are partly filled in, suggesting that either the time between the melting and freezing transitions is comparable to the equilibration time or that the system starts to switch to macroscopic behavior where the phase change occurs with the two phases in contact.

10.
J Am Chem Soc ; 131(7): 2446-7, 2009 Feb 25.
Artigo em Inglês | MEDLINE | ID: mdl-19191490

RESUMO

The kinetic energy threshold for chemisorption of N(2) on Al(100)(+) has been measured as a function of the nanocluster's temperature from 440 to 790 K. When the Al(100)(+) cluster melts at 620-660 K, the threshold drops by approximately 1 eV (approximately 96 kJ/mol). A decrease in the activation energy of this magnitude causes a 10(8)-fold increase in the reaction rate at the melting temperature. The decrease in the activation energy may result from the mobility of the surface atoms on the liquid cluster, which allows them to move to a lower energy arrangement as the N(2) approaches.

11.
J Chem Phys ; 129(12): 124709, 2008 Sep 28.
Artigo em Inglês | MEDLINE | ID: mdl-19045050

RESUMO

Heat capacities have been measured for Al(n-1)Cu(-) clusters (n=49-62) and compared with results for pure Al(n) (+) clusters. Al(n-1)Cu(-) and Al(n) (+) have the same number of atoms and the same number of valence electrons (excluding the copper d electrons). Both clusters show peaks in their heat capacities that can be attributed to melting transitions; however, substitution of an aluminum atom by a copper atom causes significant changes in the melting behavior. The sharp drop in the melting temperature that occurs between n=55 and 56 for pure aluminum clusters does not occur for the Al(n-1)Cu(-) analogs. First-principles density-functional theory has been used to locate the global minimum energy structures of the doped clusters. The results show that the copper atom substitutes for an interior aluminum atom, preferably one with a local face-centered-cubic environment. Substitution does not substantially change the electronic or geometric structures of the host cluster unless there are several Al(n) (+) isomers close to the ground state. The main structural effect is a contraction of the bond lengths around the copper impurity, which induces both a contraction of the whole cluster and a stress redistribution between the Al-Al bonds. The size dependence of the substitution energy is correlated with the change in the latent heat of melting on substitution.

12.
J Chem Phys ; 129(14): 144702, 2008 Oct 14.
Artigo em Inglês | MEDLINE | ID: mdl-19045160

RESUMO

Dissociation energies have been determined for Al(n)(+) clusters (n=25-83) using a new experimental approach that takes into account the latent heat of melting. According to the arguments presented here, the cohesive energies of the solidlike clusters are made up of contributions from the dissociation energies of the liquidlike clusters and the latent heats for melting. The size-dependent variations in the measured dissociation energies of the liquidlike clusters are small and the variations in the cohesive energies of solidlike clusters result almost entirely from variations in the latent heats for melting. To compare with the measured cohesive energies, density-functional theory has been used to search for the global minimum energy structures. Four groups of low energy structures were found: Distorted decahedral fragments, fcc fragments, fcc fragments with stacking faults, and "disordered." For most cluster sizes, the measured and calculated cohesive energies are strongly correlated. The calculations show that the variations in the cohesive energies (and the latent heats) result from a combination of geometric and electronic shell effects. For some clusters an electronic shell closing is responsible for the enhanced cohesive energy and latent heat (e.g., n=37), while for others (e.g., n=44) a structural shell closing is the cause.

13.
J Chem Phys ; 129(1): 014503, 2008 Jul 07.
Artigo em Inglês | MEDLINE | ID: mdl-18624479

RESUMO

Heat capacities measured for isolated aluminum clusters show peaks due to melting. For some clusters with around 60 and 80 atoms there is a dip in the heat capacities at a slightly lower temperature than the peak. The dips have been attributed to structural transitions. Here we report studies where the clusters are annealed before the heat capacity is measured. The dips disappear for some clusters, but in many cases they persist, even when the clusters are annealed to well above their melting temperature. This indicates that the dips do not result from badly formed clusters generated during cluster growth, as originally suggested. We develop a simple kinetic model of melting and freezing in a system consisting of one liquidlike and two solidlike states with different melting temperatures and latent heats. Using this model we are able to reproduce the experimental results including the dependence on the annealing conditions. The dips result from freezing into a high energy geometry and then annealing into the thermodynamically preferred solid. The thermodynamically preferred solid has the higher freezing temperature. However, the liquid can bypass freezing into the thermodynamically preferred solid (at high cooling rates) if the higher energy geometry has a larger freezing rate.

14.
J Phys Chem A ; 111(33): 8056-61, 2007 Aug 23.
Artigo em Inglês | MEDLINE | ID: mdl-17672436

RESUMO

Calorimetry measurements have been performed as a function of temperature for size-selected Ga(n-1)Al+ clusters with n = 17, 19, 20, 30-33, 43, 46, and 47. Heat capacities determined from these measurements are compared with previous results for pure Ga(n)+ clusters. Melting transitions are identified from peaks in the heat capacities. Substituting an aluminum atom appears to have only a small effect on the melting behavior. For clusters that show melting transitions, the melting temperatures and latent heats for the Ga(n-1)Al+ clusters are similar to those for the Ga(n)+ analogs. For Ga(n)+ clusters that do not show first-order melting transitions (n = 17, 19, and 30) the Ga(n-1)Al+ analogs also lack peaks in their heat capacities. The results suggest that the aluminum atom is not localized to a specific site in the solid-like Ga(n-1)Al+ clusters.

15.
Rev Sci Instrum ; 78(7): 075108, 2007 Jul.
Artigo em Inglês | MEDLINE | ID: mdl-17672794

RESUMO

The translating and rotating rod or disk of a conventional laser vaporization cluster source is replaced by a liquid metal target. The self-regenerating liquid surface prevents cavities from being bored into the sample by laser ablation. The laser beam strikes a near pristine surface with each pulse, resulting in signals with much better short and long term stabilities. While this approach cannot be used for refractory metals such as tungsten and molybdenum, it is ideal for studies of bimetallic clusters, which can easily be prepared by laser vaporization of a liquid metal alloy.


Assuntos
Cristalização/instrumentação , Gases/química , Lasers , Metais/química , Desenho de Equipamento , Análise de Falha de Equipamento , Reprodutibilidade dos Testes , Sensibilidade e Especificidade , Soluções
16.
J Am Soc Mass Spectrom ; 18(1): 74-81, 2007 Jan.
Artigo em Inglês | MEDLINE | ID: mdl-17010642

RESUMO

Calorimetry measurements have been used to probe the melting of aluminum cluster cations with 63 to 83 atoms. Heat capacities were determined as a function of temperature (from 150 to 1050 K) for size-selected cluster ions using an approach based on multicollision-induced dissociation. The experimental method is described in detail and the assumptions are critically evaluated. Most of the aluminum clusters in the size range examined here show a distinct peak in their heat capacities that is attributed to a melting transition (the peak is due to the latent heat). The melting temperatures are below the bulk melting point and show enormous fluctuations as a function of cluster size. Some clusters (for example, n = 64, 68, and 69) do not show peaks in their heat capacities. This behavior is probably due to the clusters having a disordered solid-like phase, so that melting occurs without a latent heat.

SELEÇÃO DE REFERÊNCIAS
DETALHE DA PESQUISA
...