Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 19 de 19
Filtrar
Mais filtros










Base de dados
Intervalo de ano de publicação
1.
Phys Chem Chem Phys ; 14(28): 9826-37, 2012 Jul 28.
Artigo em Inglês | MEDLINE | ID: mdl-22710404

RESUMO

This article reports a theoretical study of the stereodynamics of Ar + NO(X(2)Π, v = 0, j = 1/2, Ω = 1/2, ε = ±1) rotationally inelastic collisions. First, quantum scattering data are used to calculate all differential polarisation moments of the reagent and product molecules; this leads to the observation that the orientations of the reagent and product angular momenta are very strongly correlated. Next, canonical collision mechanisms theory [Aldegunde et al., Phys. Chem. Chem. Phys., 2008, 10, 1139] is used to separate and characterise the stereodynamics of the two independent collision mechanisms that contribute to the collision dynamics; this leads to the observation that the average product orientation is determined by the relative contributions of the two canonical mechanisms, which have comparable importance but are associated with starkly contrasting angular momentum orientations. These observations lead to a new and rigorous explanation of the experimental results reported a decade ago by Lorenz et al. [Science, 2001, 293, 2063]. The central fact of the new explanation is the incoherent, interference-free superposition of two independent collision mechanisms. This makes the new explanation radically different from the only one previously suggested, namely that the experimental observations might be due to quantum interference in a single collision mechanism.

2.
Phys Chem Chem Phys ; 13(18): 8345-58, 2011 May 14.
Artigo em Inglês | MEDLINE | ID: mdl-21279213

RESUMO

The stereodynamics and mechanism of the F + HD(v = 0, j = 1) → HF (DF) + D (H) reactions have been thoroughly analysed at collision energies in the 0-160 meV range. Specifically, this study is focused on (i) the comparison between the stereodynamics of the collisions leading to HF and DF formation, and (ii) the stereodynamical fingerprints of the resonance that occurs at low collision energies in the HF channel and whose manifestation in the total cross section is greatly diminished for initial j > 0. While previous studies were limited to the analysis of integral cross sections (ICS), differential cross sections (DCS) and reaction probabilities, in the present work we have included the analysis of vectorial quantities such as the direction of the initial rotational angular momentum and internuclear axis, and their effect on reactivity. In particular, polarisation parameters (PP) and polarisation dependent differential cross sections (PDDCS), quantities that describe how the intrinsic HD rotational angular momentum and molecular axis polarisations contribute to reaction, are calculated and examined. The evolution of the PPs with the collision energy differs markedly between the two reaction channels. For the DF channel, the PP values are small and change very little in the energy range in which DF formation is appreciable. In contrast, rapid fluctuations in the magnitude and sign of the PPs are observed in the HF channel at low collision energies in and around the resonance. As the collision energy increases, direct (non-resonant) scattering prevails, and the various quantities are reasonably well accounted for by the QCT calculations, as in the case of the DF channel. The intrinsic directional information has been used to access the extent of control that can be achieved through polarisation of the HD molecule prior to collision. It was found that the same extrinsic preparation leads to very different outcomes on the HF channel DCS when the collision energy is close to the resonance. It is also shown that polarisation of the HD internuclear axis along the initial relative velocity enhances the effect of the resonance and allows its clear identification. Finally, the effect of different extrinsic preparations on the angle-velocity DCS is found to be strong, thus allowing considerable control of product angular distributions.

3.
Inorg Chem ; 49(23): 11127-32, 2010 Dec 06.
Artigo em Inglês | MEDLINE | ID: mdl-21067148

RESUMO

The reaction of CuX(2) (X(-) ≠ F(-)) salts with 1 equiv of 3-pyridyl-5-tert-butylpyrazole (HL) in basic methanol yields blue solids, from which disk complexes of the type [Cu(7)(µ(3)-OH)(4)(µ-OR)(2)(µ-L)(6)](2+) and/or the cubane [Cu(4)(µ(3)-OH)(4)(HL)(4)](4+) can be isolated by recrystallization under the appropriate conditions. Two of the disk complexes have been prepared in crystalline form: [Cu(7)(µ(3)-OH)(4)(µ-OCH(2)CF(3))(2)(µ-L)(6)][BF(4)](2) (2) and [Cu(7)(µ(3)-OH)(4)(µ-OCH(3))(2)(µ-L)(6)]Cl(2)·xCH(2)Cl(2) (3·xCH(2)Cl(2)). The molecular structures of both compounds as solvated crystals can be described as [Cu⊂Cu(6)(µ-OH)(4)(µ-OR)(2)(µ-L)(6)](2+) (R = CH(2)CF(3) or CH(3)) adducts. The [Cu(6)(µ-OH)(4)(µ-OR)(2)(µ-L)(6)] ring is constructed of six square-pyramidal Cu ions, linked by 1,2-pyrazolido bridges from the L(-) ligands and by basal, apical-bridging hydroxy or alkoxy groups, while the central Cu ion is bound to the four metallamacrocyclic hydroxy donors in a near-regular square-planar geometry. The L(-) ligands project above and below the metal ion core, forming two bowl-shaped cavities that are fully (R = CH(2)CF(3)) or partially (R = CH(3)) occupied by the alkoxy R substituents. Variable-temperature magnetic susceptibility measurements on 2 demonstrated antiferromagnetic interactions between the Cu ions, yielding a spin-frustrated S = (1)/(2) magnetic ground state that is fully populated below around 15 K. Electrospray ionization mass spectrometry, UV/vis/near-IR, and electron paramagnetic resonance measurements imply that the heptacopper(II) disk motif is robust in organic solvents.

4.
Phys Chem Chem Phys ; 12(41): 13626-36, 2010 Nov 07.
Artigo em Inglês | MEDLINE | ID: mdl-20852814

RESUMO

This article analyses the mechanisms of inelastic and reactive H + D(2)(v = 0, j = 2) collisions that result in highly vibrationally excited products when the collision energy is 1.70 eV. The analytical method is entirely quantum mechanical and focuses on correlations between the polarization of the reactant molecule and the direction of product scattering. Two viewpoints are used. The "intrinsic" viewpoint reveals the reactant polarizations that lead to the largest cross section at each value of the scattering angle (the angle between the reactant-approach and product-recoil directions); the "extrinsic" viewpoint reveals how the dependence of the collision cross section on the scattering angle changes when the reactant polarization is fixed at each one of a set of experimentally feasible alternatives. Comparison of processes correlating with the same range of impact parameters is also used, to facilitate isolation and identification of directional effects. When products are scattered in the backward and sideways regions, the results for inelastic and reactive collisions are rather similar. When products are scattered in the forward region, the results for inelastic and reactive collisions are clearly different: a side-on collision geometry that largely increases the inelastic cross section hardly affects the reactive cross section. This feature is the quantum mechanical signature of the so-called "tug-of-war" mechanism.


Assuntos
Deutério/química , Hidrogênio/química , Teoria Quântica , Termodinâmica , Vibração
5.
J Chem Phys ; 132(16): 161102, 2010 Apr 28.
Artigo em Inglês | MEDLINE | ID: mdl-20441249

RESUMO

Following upon our recent work on vector correlations in the Ar-NO collisions [Lemeshko and Friedrich, Phys. Chem. Chem. Phys. 12, 1038 (2010)], we compare model results with close-coupling calculations for a range of channels and collision energies for the He-NO system. The striking agreement between the model and exact polarization moments indicates that the stereodynamics of rotationally inelastic atom-molecule collisions at thermal energies is governed by diffraction of matter waves from a two-dimensional repulsive core of the atom-molecule potential. Furthermore, the model polarization moments characterizing the He-NO, He-O(2), He-OH, and He-CaH stereodynamics are found to coalesce into a single, distinctive pattern, which can serve as a "fingerprint" to identify diffraction-driven stereodynamics in future work.

6.
J Phys Chem A ; 113(52): 14943-51, 2009 Dec 31.
Artigo em Inglês | MEDLINE | ID: mdl-20028174

RESUMO

This article introduces the concept of intrinsic entropy, S, of a molecular collision. Defined in rigorously quantum mechanical terms as the von Neumann entropy of the intrinsic density matrices of reagents and products, the intrinsic entropy is a dimensionless number in the 0 < or = S < or = 1 range. Its limits are associated with situations where the collision cross section is due to a single combination of reagent and product polarizations (S = 0) or where there is absolutely no selectivity with respect to the molecular polarizations (S = 1). The usefulness of the intrinsic entropy as a quantifier of the sensitivity of a molecular collision to molecular polarizations is demonstrated with examples for the benchmark H + D(2) reaction.

7.
J Chem Phys ; 131(5): 054308, 2009 Aug 07.
Artigo em Inglês | MEDLINE | ID: mdl-19673563

RESUMO

Direct current slice and crush velocity map imaging has been used to probe the photodissociation dynamics of nitrogen dioxide above the second dissociation limit. The paper is a companion to a previous publication [J. Chem. Phys. 128, 164318 (2008)] in which we reported results for the O((3)P(J)) + NO((2)Pi(Omega)) adiabatic product channel. Here we examine the O((1)D(2)) + NO((2)Pi(Omega)) diabatic product channel at similar excitation energies. Using one- and two-color imaging experiments to observe the velocity distributions of state selected NO fragments and O atoms, respectively, we are able to build a detailed picture of the dissociation dynamics. We show that by combining the information obtained from velocity map imaging studies with mass-resolved resonantly enhanced multiphoton ionization spectroscopy it is possible to interpret and fully assign the NO images. By recording two-color images of the O((1)D(2)) photofragments with different polarization combinations of the pump and probe laser fields we also measure the orbital angular momentum alignment in the atomic fragment. We find that the entire O((1)D(2)) photofragment distribution is similarly aligned with most of the population in the M(J) = +/-1 magnetic sublevels. The similarity of the fragment polarizations is interpreted as a signature of all of the O((1)D(2)) atoms being formed via the same avoided crossing. At the photolysis energy of 5.479 52 eV we find that the NO fragments are preferentially formed in v = 1 and that the vibrationally excited fragments exhibit a bimodal rotational distribution. This is in contrast to the unimodal rotational profile of the NO fragments in v = 0. We discuss these observations in terms of the calculated topology of the adiabatic potential energy surfaces and attribute the vibrational inversion and rotational bimodality of the v = 1 fragments to the symmetric stretch and bending motion generated on excitation to the (2) (2)B(2) state.

8.
Transplantation ; 87(6): 919-25, 2009 Mar 27.
Artigo em Inglês | MEDLINE | ID: mdl-19300197

RESUMO

BACKGROUND: Clinical demand for liver transplant steadily grows while organs offer has reached a plateau years ago. To expand the donor liver pool, various options have been considered including acceptance of suboptimal donors and steatotic grafts, with a risk of poorer outcomes. The latter risk and its relation to the grade of liver graft steatosis have been studied in this prospective clinical study. METHODS: One hundred eighteen consecutive liver transplantation (115 patients) performed between May 2002 and March 2008 were prospectively analyzed. According to the grade of steatosis on a 2 hr postreperfusion biopsy, four groups were considered: absence (<5%) (n=34), mild (<30%) (n=40), moderate (30%-60%) (n=23), or severe steatosis (> or = 60%) (n=21). Donors and recipients demographic data, and patients and grafts survival rates were compared among the four groups. RESULTS: Eighty-four (71%) grafts presented some degree of steatosis (macrosteatosis: 19.5%, microsteatosis: 47%, mix type: 33.5%). Patient and graft survival were significant lower in the "severe steatosis" group, as a whole. Grafts with less than 30% predominant macro-, or microsteatosis also had poorer outcomes with lower patient and graft survival rates. CONCLUSION: Steatotic liver grafts were used on a large scale (71%) in this clinical series. The analysis confirms that using grafts with moderate (>30%) and severe steatosis (>60%) have a negative impact on outcomes. The authors conclude that using these grafts allow a significant increase in organ offer that counterbalances the negative outcome for patients who are not offered a transplant, and this supports the need for further clinical research.


Assuntos
Fígado Gorduroso/patologia , Falência Hepática/cirurgia , Transplante de Fígado/fisiologia , Doadores de Tecidos/provisão & distribuição , Coleta de Tecidos e Órgãos/métodos , Adulto , Alanina Transaminase/sangue , Aspartato Aminotransferases/sangue , Biópsia , Cadáver , Causas de Morte , Fígado Gorduroso/classificação , Feminino , Humanos , Testes de Função Hepática , Transplante de Fígado/métodos , Masculino , Pessoa de Meia-Idade , Estudos Prospectivos , Estudos Retrospectivos , Sódio/sangue , Coleta de Tecidos e Órgãos/estatística & dados numéricos
9.
Neotrop Entomol ; 38(6): 827-33, 2009.
Artigo em Português | MEDLINE | ID: mdl-20098930

RESUMO

The causal agent of citrus variegated clorosis, Xylella fastidiosa, is transmitted by leafhoppers of the subfamily Cicadellinae, whose species vary regionally. The goal of this study was to identify potential vectors of this pathogen in citrus groves of Bahia North Coast, Brazil. The survey was done from March/2002 to February/2003 in three seven- to nine-year-old sweet orange (Citrus sinensis, Pêra variety) groves located in Rio Real, BA. Fifteen yellow sticky cards (8.5x11.5 cm) were installed 40 m apart in each grove, hanged at 1.5 m high on the upper north side of citrus canopies, and replaced fortnightly. A sweep net was periodically used to sample leafhoppers on herbaceous weeds inside the groves, by selecting five points at random and performing 30 sweeps in each point. A total of 1,360 specimens of 49 Auchenorrhyncha species were collected in this study, mostly in the family Cicadellidae (90.2%). The subfamily Cicadellinae, which includes the sharpshooter vectors of X. fastidiosa, showed the largest number of species (14) and specimens (84.8%). Acrogonia flagellata Young, A. citrina Marucci & Cavichioli, Homalodisca spottii Takiya, Cavichioli & McKamey and an unidentified Cicadellini (species 1) were the dominant species trapped on citrus canopies, while Hortensia similis (Walker) and Erythrogonia dubia (Medler) were dominant in the weedy vegetation. Among the Cicadellinae species already known as vectors of X. fastidiosa in citrus, only A. citrina, Bucephalogonia xanthophis (Berg) e Ferrariana trivittata (Signoret) were found. The two latter species were accidentally trapped by sweep net in the weedy vegetation.


Assuntos
Citrus/microbiologia , Vetores de Doenças , Hemípteros/microbiologia , Doenças das Plantas/microbiologia , Xylella/fisiologia , Animais , Brasil
10.
Phys Chem Chem Phys ; 10(8): 1139-50, 2008 Feb 28.
Artigo em Inglês | MEDLINE | ID: mdl-18270616

RESUMO

This article considers the extent to which one can control the reactivity of atom-diatom systems through reactant polarisation. Three different limits for reactivity manipulation are defined: "absolute" limits that do not depend on the reaction dynamics but can only be obtained for particular combinations of quantum numbers, "unconstrained" limits that depend on dynamics but not on constraints imposed by any particular experimental setup, and "constrained" limits that depend on dynamics and also on the constraints imposed by a particular experimental setup. Methods for calculation of these limits are presented and applied to the benchmark F + H2 reaction. The variations of the maximum and minimum reactivity one can obtain are analysed in terms of reaction mechanisms and steric constraints. Tables listing the minimum and maximum values of angular momentum polarisation moments of rank up to 4, and integer and half-integer quantum numbers up to 5, are also presented.

12.
Phys Chem Chem Phys ; 9(43): 5794-808, 2007 Nov 21.
Artigo em Inglês | MEDLINE | ID: mdl-19462575

RESUMO

This article introduces a definition of the concept of elementary reaction mechanism that, while conforming to the traditional view of reaction mechanisms as dynamical processes whereby reagents are transformed into products, sharpens it by requiring reagent and product states to be completely specified and fully correlated. This leads to well-defined mathematical requirements for classification of a dynamical process as a reaction mechanism and also to a straightforward mathematical procedure for the determination of a special class of independent collision mechanisms that are dubbed "canonical". Canonical mechanisms result from an exact decomposition of the differential cross section of the reaction and form a complete orthogonal basis in terms of which all reaction mechanisms can be described. Examples involving the benchmark F + H2 and D + H2 reactions at energies ranging from ultralow to hyperthermal illustrate how canonical and other reaction mechanisms can be visualised and also how analysis of a reaction in terms of its canonical mechanisms can provide insight into its dynamics.

13.
J Phys Chem A ; 109(28): 6200-17, 2005 Jul 21.
Artigo em Inglês | MEDLINE | ID: mdl-16833960

RESUMO

This article presents theoretical methods for the description of the directional effect of reactant rotation on the reactivity of atom-diatom systems and suggests an experiment that could be used to test theoretical predictions. The theory can be used in conjunction with both quantum reactive scattering and quasiclassical trajectory calculations, and is stated in general terms, which allows it to deal with arbitrary reactant polarizations. The illustrative results obtained for the benchmark H + D2 reaction are also presented and show that under experimentally achievable conditions one can largely control reactive cross sections and product state distributions, while at the same time gaining valuable and at times surprising information on the reaction mechanism.

14.
J Chem Phys ; 121(20): 9830-43, 2004 Nov 22.
Artigo em Inglês | MEDLINE | ID: mdl-15549856

RESUMO

We have recently reported a derivation of the relationship between the quantum and classical descriptions of angular momentum polarization [M. P. de Miranda and F. Javier Aoiz, Phys. Rev. Lett. 93, 083201 (2004)]. This paper presents a detailed account of the derivation outlined in that paper, and discusses the implications of the new results. These include (i) a new expression of the role of the uncertainty principle in the broadening of angular momentum distributions, (ii) the attribution of azimuthal fluctuations of angular momentum distributions to spatial quantum beats, (iii) the definition of a new Fourier transform of the density matrix, distinct from those suggested in the past, that provides an alternative view of how the quantum description of angular momentum polarization approaches the classical one in the correspondence principle limit, (iv) a prescription for the determination of a quasiclassical angular momentum distribution function that does not suffer from problems encountered with its purely classical counterpart, and (v) a description of how angular momentum distributions commonly visualized with recourse to the classical vector model can be depicted with exact and well-defined quantum mechanics.

15.
Phys Rev Lett ; 93(8): 083201, 2004 Aug 20.
Artigo em Inglês | MEDLINE | ID: mdl-15447182

RESUMO

This Letter presents a derivation of the relationship between the quantum and classical descriptions of angular momentum polarization. The results involve an "uncertainty broadening" term that directly expresses the restrictions imposed by the uncertainty principle. It is argued that neglect of this term can lead to error in the interpretation of theoretical or experimental angular momentum polarization data. Functions that take the uncertainty broadening into account, appropriate for use in quantum or quasiclassical descriptions of spatial distributions of angular momenta, are defined.

16.
Dalton Trans ; (1): 59-64, 2004 Jan 07.
Artigo em Inglês | MEDLINE | ID: mdl-15356742

RESUMO

Reaction of CuX2(X-=Cl- or Br-) with 2 molar equivalents of 3[5]-(2,4,6-trimethylphenyl)pyrazole (HpzMes) in MeOH in the presence of NaOH yields [Cu3X(HpzMes)2(micro-pzMes)3(micro3-OMe)]X (X-=Cl- or Br-). Crystal structures of these compounds show almost identical triangles of Cu(II) ions, centred by a triply bridging methoxide ligand and with three edge-bridging pyrazolide groups. The mesityl substituents on the bridging pyrazolide ligands are arranged in HT, HH, TT fashion. chi(M)T for both compounds decreases steadily with decreasing temperature, reaching 0.40 cm(3) mol(-1) K at 70 K before decreasing further below 40 K. This low temperature behaviour could not be interpreted using conventional superexchange Hamiltonians, but was reproduced by an alternative model that incorporated an additional antisymmetric exchange term. This interpretation was confirmed by the Q-band EPR spectra of the two compounds. NMR experiments show that the structures of these compounds are not retained in solution, in contrast to other closely related tricopper compounds. These are the first examples of triangular Cu(II) compounds bearing a [Cu3micro3-OR)]5+(R is not equal to H) core motif, and the first triangular compounds showing antisymmetric exchange to have been analysed by both susceptibility and EPR measurements.

17.
Chemistry ; 10(7): 1827-37, 2004 Apr 02.
Artigo em Inglês | MEDLINE | ID: mdl-15054770

RESUMO

Reaction of CuX(2) (X(-)=Cl(-), Br(-), NO(3) (-)), NaOH, and 3[5]-tert-butylpyrazole (Hpz(tBu)) in a 1:1:2 molar ratio in MeOH at 293 K for three days affords [[Cu(3)(Hpz(tBu))(6)(mu(3)-X)(mu(3)-OH)(3)](2)Cu]X(6) (X(-)=Cl(-), 1; X(-)=Br(-), 2; X(-)=NO(3)(-), 3) in moderate yields. These compounds contain a centrosymmetric, vertex-sharing double-cubane [[Cu(3)(Hpz(tBu))(6)(mu(3)-X)(mu(3)-OH)(3)](2)Cu](6+) core, surrounded by a belt of six hydrogen-bonded X(-) ions. For 1 and 2, the ring of guest anions has near C(3) symmetry, that is slightly distorted owing to the axis of Jahn-Teller elongation at the central Cu ion. For 3 only, the NO(3)(-) guest ions are crystallographically disordered, reflecting their poor complimentarity with complex host. A similar reaction employing CuF(2) yields [[Cu(3)(Hpz(tBu))(4)(mu-pz(tBu))(2)(mu-F)(2)(mu(3)-F)](2)]F(2) (4), whose structure contains a cyclic hexacopper core with approximate C(2v) symmetry. Finally, an analogous reaction using Cu(NCS)(2) gives a mixture of trans-[Cu(NCS)(2)(Hpz(tBu))(2)] (5) and [Cu(2)(NCS)(2)(mu-pz(tBu))(2)(mu-Hpz(tBu))(Hpz(tBu))(2)] (6). The latter compound contains a Hpzt(Bu) ligand bridging the two Cu ions in an unusual kappa(1),mu-coordination mode. The variable temperature magnetic properties of 1-3 show antiferromagnetic behavior, leading to a S=1/2 ground state in which the seven copper(II) ions are associated into three mutually independent distinct spin systems. In confirmation of this interpretation, Q-band EPR spectra of solid 1 and 2 at 5 K also demonstrate a S= 1/2 spin system and exhibit hyperfine coupling to three (63,65)Cu nuclei. Unusually, the coupling is manifest as an eight-line splitting of the parallel feature, rather than the usual 10 lines. This has been rationalized by a spin-projection calculation, and results from the relative magnitudes of coupling to the three Cu nuclei. UV/Vis and mass spectrometric data show that 1-4 decompose to lower nuclearity species in solution.

19.
Chem Commun (Camb) ; (24): 2978-9, 2002 Dec 21.
Artigo em Inglês | MEDLINE | ID: mdl-12536775

RESUMO

The complex [[Cu3(HpztBu)4(mu-pztBu)2(mu-F)2(mu 3-F)]2]F2 (HpztBu = 3[5]-tert-butylpyrazole) has a cyclic, C2v-symmetric hexacopper core. The two non-coordinated F- anions are encapsulated within cavities formed by three HpztBu ligands.

SELEÇÃO DE REFERÊNCIAS
DETALHE DA PESQUISA
...