首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The authors of the title paper (J Therm Anal Calorim 2012,110:873–878) report to have grown a dipeptide hydrochloride crystal namely glycyl-l-alanine hydrochloride by the slow evaporation of an aqueous solution containing stoichiometric amounts of l-alanine and glycine and an excess of hydrochloric acid. In this letter, we prove that no such dipeptide hydrochloride can be crystallized simply by mixing two amino acids in aqueous hydrochloric acid.  相似文献   

2.
This article deals with a semi‐analytical study on the electrophoresis of charged spherical rigid colloid by considering the effects of relaxation and ion size. The particle surface is taken to be either hydrophilic or hydrophobic in nature. In order to consider the ion size effect we have invoked the Carnahan and Starling model (J. Chem. Phys. 1969, 51, 635‐636). The mathematical model is based on Stokes equation for fluid flow, modified Boltzmann equation for spatial distribution of ionic species and Poisson equation for electric potential. We adopt a linear perturbation technique under a weak electric field assumption. An iterative numerical technique in employed to solve the coupled set of perturbed equations. We have validated the numerically obtained electrophoretic mobility with the corresponding analytical solution derived under low potential limit. Going beyond the widely employed Debye‐Hückel linearization, we have presented the results for a wide range of surface charge density, electrolyte concentration, and slip length to Debye length ratio. We have also identified several interesting features including occurrence of local maxima and minima in the mobility for critical choice of pertinent parameters.  相似文献   

3.
In polyhedral crystallography, non-Euclidean possibilities come into considerations as well. The so-called football manifolds appearing in the Bolyai?CLobachevskian hyperbolic space (H 3) can model ??fullerens?? very probably. This may give us the feeling that??in atomic measure??space can wear non-Euclidean structures, e.g., hyperbolic (H 3), spherical (S 3) and other ones, respectively. Such phenomena as pentagonal or decagonal symmetry and the discovery of quasicrystals might strengthen these impressions. We examine here some Archimedean solids: (4, 6, 6) in E 3 (the truncated octahedron in the classical Euclidean space, for analogue as motivation, Fig.?1); (5, 6, 6) the ??real football?? (now in H 3, Fig.?2, from Molnár in Proceedings of International Conference on Differential Geometry and Its Applications, 1988), (3, 10, 10) in H 3 (Fig.?3 from Molnár in the previous Proceedings); an extra solid (4, 6, 8) with two manifolds on it in H 3 (Figs.?4, 5), published first time in this paper. Furthermore (3, 8, 8) will be realized in the spherical space S 3 (Fig.?6). We can study this classical example as a simple non-Euclidean one also for motivation to our method (see also Molnár in Acta Cryst A61:542?C552, 2005, as an introductory paper). Each of them will be a manifold, i.e., it can be endowed with an appropriate face pairing (topological glueing), so that every point has a ball-like neighborhood in the corresponding space, especially of constant curvature, but completely in the classical analogy. In addition, each manifold will be compact (bounded and closed) whose fundamental group can be generated by two screw motions.  相似文献   

4.
A minimum-basis diatomics-in-molecules (DIM) model previously developed for singly-ionized argon clusters is applied to neon clusters, Ne n + , forn=3, 4,...,22. A search for the global minimum energy of each cluster yields structures with the positive charge localised on a dimer-ion. This appears to be due largely to the linear unsymmetrical configuration which the model finds for Ne 3 + . For this reason, the structures of the clusters at their minimum energy are different from those for Ar n + computed with the same model. On the other hand, the behaviour of the charge distribution as a function of the geometrical configuration is similar to that for Ar n + , as are the overall shapes of the potential energy surfaces. The results are discussed in terms of the charge distributions and the ratios of equilibrium properties of the dimers and dimer-ions which constitute the input to the model.  相似文献   

5.
Polycrystalline thorium(IV) phosphate-triphosphate, Th2(PO4)(P3O10) (1), was obtained by (NH4)2Th(PO4)2·H2O (2) heating from room temperature to 1,273 K. 1 crystallizes in the orthorhombic space group Pn21 a (a = 11.6846(2) Å, b = 7.1746(1) Å, c = 12.9320(3) Å, Z = 4). Combining powder synchrotron X-ray diffraction data and DFT geometry optimization, a structural model is proposed for 1. The structure is built on ThO8 polyhedral chains along the b-axis. PO4 3? and P3O10 5? groups coexist in the structure and the latter group forms non-linear chains. Cohesion of the structure is made by the linkage of ThO8 chains by PO4 and P3O10 groups. Thermal transformation from 2 to 1 was monitored by thermogravimetric analysis (activation energy as a function of the extent of conversion was obtained from Kissinger–Akahira–Sunose (KAS) isoconversional method) and powder X-ray thermo-diffraction. For 2, the dehydration process takes place in two steps, with the apparition of a layered intermediate phase, (NH4)2Th(PO4)2·nH2O (0 < n < 1, d = 6.42 Å), previously to the formation of (NH4)2Th(PO4)2 (d = 6.31 Å). The condensation process produces an amorphous material that crystallizes to α-ThP2O7 (3) when the temperature increases. At 1,273 K, 3 slowly transforms to 1.  相似文献   

6.
We first introduce the latest experimental results, i.e., production of the fine nanostructured and near fully dense transparent Y3Al5O12 (YAG) bulks at high pressure and modest temperature (2.0–5.0 GPa and 300–500 °C). And then, we employ the first-principles plane wave pseudopotential density functional theory method to calculate the equilibrium lattice parameters and the thermodynamic properties of YAG. The obtained lattice parameters are consistent with the experimental data and the available theoretical data of others. Through the quasi-harmonic Debye model, the dependences of the normalized primitive volume V/V 0 on pressure P, the Debye temperature $ \Uptheta_{\rm{D}} $ , and the heat capacity C V on pressure P and temperature T, as well as the variation of the thermal expansion α with temperature and pressure are obtained successfully.  相似文献   

7.
The electrophoresis of a charged soft particle with charged rigid core is considered under a weak imposed field condition. The rigid core of the soft particle is considered to have a finite dielectric permittivity and a fixed volume charge density. The electric potential distribution is determined by solving the Poisson-Boltzman equation out side the rigid core and a Poisson equation within the core along with continuity conditions on the core-shell interface. We have extended the analytic expression of Ohshima (Electrophoresis 27:526–533, 2006) for the electrophoretic mobility of a soft particle with a charged shell to include the effect of the volume charge density of the rigid core. Mobility based on the present expression matches exactly with the existing analytical solutions for a soft particle with an uncharged core. We have also made a comparison of our solution for mobility with an uncharged rigid core with the existing experimental results. The impact of the core charge density on the soft particle mobility is analyzed.  相似文献   

8.
We investigated geometry, energy, ${\nu_{{\text{N--H}}}}$ harmonic frequencies, 14N nuclear quadrupole coupling tensors, and ${n_{\rm O}\to \sigma _{{\text{N--H}}}^\ast}$ charge transfer properties of (acetamide) n clusters, with n = 1 ? 7, by means of second-order Møller-Plesset perturbation theory (MP2) and DFT method. Dependency of dimer stabilization energies and equilibrium geometries on various levels of theory was examined. B3LYP/6-311++G** calculations revealed that for acetamide clusters, the average hydrogen-bonding energy per monomer increases from ?26.85 kJ mol?1 in dimer to ?35.12 kJ mol?1 in heptamer; i.e., 31% cooperativity enhancement. The n-dependent trend of ${\nu_{{\text{N--H}}}\,{and}\,^{14}}$ N nuclear quadrupole coupling values were reasonably correlated with cooperative effects in ${r_{{\text{N--H}}}}$ bond distance. It was also found that intermolecular ${n_{\rm O}\to \sigma_{{\text{N--H}}}^\ast}$ charge transfer plays a key role in cooperative changes of geometry, binding energy, ${\nu_{{\text{N--H}}}}$ harmonic frequencies, and 14N electric field gradient tensors of acetamide clusters. There is a good linear correlation between 14N quadrupole coupling constants, C Q (14N), and the strength of Fock matrix elements (F ij ). Regarding the ${n_{\rm O}\to \sigma_{{\text{N--H}}}^\ast}$ interaction, the capability of the acetamide clusters for electron localization, at the N–H· · ·O bond critical point, depends on the cluster size and thereby leads to cooperative changes in the N–H· · ·O length and strength, N–H stretching frequencies, and 14N quadrupole coupling tensors.  相似文献   

9.
In many scientific fields, the half-life of radionuclides plays an important role. The accurate knowledge of this parameter has direct impact on, e.g., age determination of archeological artifacts and of the elemental synthesis in the universe. In order to derive the half-life of a long-lived radionuclide, the activity and the absolute number of atoms have to be analyzed. Whereas conventional radiation measurement methods are typically applied for activity determinations, the latter can be determined with high accuracy by mass spectrometric techniques. Over the past years, the half-lives of several radionuclides have been specified by means of multiple-collector inductively coupled plasma mass spectrometry (MC-ICP-MS) complementary to the earlier reported values mainly derived by accelerator mass spectrometry. The present paper discusses all critical aspects (amount of material, radiochemical sample preparation, interference correction, isotope dilution mass spectrometry, calculation of measurement uncertainty) for a precise analysis of the number of atoms by MC-ICP-MS exemplified for the recently published half-life determination of $^{60}$ Fe (Rugel et al, Phys Rev Lett 103:072502, 2009).  相似文献   

10.
Collisional charge exchange between mass selected alkali cluster ions and Cs has been studied and cross sections have been determined for the processes Na n + + Cs and K n + + Cs, withn=1–21 andn=1–14, respectively. A strong dependence of the cross sections on the energy defect as well as on cluster size and collision energy is found. The results are analysed by a coupled two state density matrix model, taking account of the relaxation of electronic amplitudes due to interaction with the nuclear motion in the cluster.  相似文献   

11.
Cellulose beads were prepared from water-based solvent and oxidised by modified Anelli’s reaction at 20– $80\,^\circ \hbox {C}$ for 2–48 h (Fig. 1). The maximum amount of anionic groups (AGs) was $1.85\,\hbox {mmol}\,\hbox {g}^{-1}$ . The distribution of AGs was verified by absorption of cationic dyes and imaging with confocal fluorescent microscopy. Structural changes were studied spectroscopically and with electron microscopy. Oxidation of the beads drastically increased the swelling capacity of air-dried beads. Uptake of model drug was more than doubled in never-dried beads. This is due to the changes in pore size distribution, mainly opening and widening of the closed pores and narrow cavities. Release profiles of the drug were studied at physiological pH of 7.4 and showed a controlled release rate independently of the amount of the drug encapsulated and amount of AGs.  相似文献   

12.
13.
We present the first theoretical investigation of solvent effects on the Faraday ${\mathcal{B}}$ term of magnetic circular dichroism (MCD) at the density–functional level of theory. In our model, the solvent is described by the polarizable continuum model in its integral-equation formulation. We present the extensions required for including electron correlation effects using density–functional theory (DFT) as well as the necessary extensions for including the effects of a dielectric continuum. The new code is applied to the study of the Faraday ${\mathcal{B}}$ term of MCD in a series of benzoquinones. It is demonstrated that electron correlation effects, as described by DFT, are essential in order to recover the experimentally observed signs of the ${\mathcal{B}}$ term. Dielectric continuum effects increase, in general, the magnitude of the ${\mathcal{B}}$ term, leading to an overestimation of the experimental observations in most cases.  相似文献   

14.
We have studied the electronic structure (ground and excited states) and g matrix of a model compound for oxomolybdenum enzymes featuring the MoO–dithiolate moiety in C s symmetry, by means of multiconfigurational second-order perturbation theory (CASPT2) for a range of fold angles (5–29°), i.e. the angle between the S–Mo–S and S–C–C–S planes of the dithiolate ligand. We found no evidence of a suggested 3-center pseudo-σ bonding interaction between the singly occupied orbital of the ground state and the symmetric in-plane dithiolate orbital (Inscore et al. in Inorg Chem 38:1401–1410, 1999). This is complemented with our alternative assignment of band 4 in the electronic spectrum as the transition out of the a″ instead of the a′ dithiolate in-plane orbital into the singly occupied ground-state orbital, believed to probe the dominant hole superexchange pathway (Inscore et al. in Inorg Chem 38:1401–1410, 1999; Burgmayer et al. in J Inorg Biochem 101:1601–1616, 2007). Principal g values of 1.9652, 1.9090, 2.0003 were obtained at a fold angle of 21°. The latter value is so close to the free electron g e factor is due to an important positive contribution from the LMCT transition corresponding to band 4, counteracting the negative contributions from the ligand field transitions.  相似文献   

15.
The density functional theory (DFT) calculations within the framework of generalized gradient approximation have been employed to systematically investigate the geometrical structures, stabilities, and electronic properties of Pd n Si q (n = 1–7 and q = 0, +1, ?1) clusters and compared them with the pure ${\text{Pd}}_{n + 1}^{q}$ (n = 1–7 and q = 0, +1, ?1) clusters for illustrating the effect of doping Si atom into palladium nanoclusters. The most stable configurations adopt a three-dimensional structure for both pure and Si-doped palladium clusters at n = 3–7. As a result of doping, the Pd n Si clusters adopt different geometries as compared to that of Pd n+1. A careful analysis of the binding energies per atom, fragmentation energies, second-order difference of energies, and HOMO–LUMO energy gaps as a function of cluster size shows that the clusters ${\text{Pd}}_{4}^{ + }$ , ${\text{Pd}}_{4}$ , ${\text{Pd}}_{8}^{ - }$ , ${\text{Pd}}_{5} {\text{Si}}^{0, + , - }$ , and ${\text{Pd}}_{7} {\text{Si}}^{0, + , - }$ possess relatively higher stability. There is enhancement in the stabilities of palladium frameworks due to doping with an impurity atom. In addition, the charge transfer has been analyzed to understand the effect of doped atom and compared further.  相似文献   

16.
Standard transfer Gibbs energies, $ \Updelta_{\text{tr}} G^{^\circ } $ , of a series of homologues α-amino acids have been evaluated by determining the solubility of glycine, alanine, amino butyric acid and norvaline gravimetrically at 298.15 K. Standard entropies of transfer, $ \Updelta_{\text{tr}} S^{^\circ } $ , of the amino acids have also been evaluated by extending the solubility measurement to five equidistant temperatures ranging from 288.15 to 308.15 K. The chemical contributions $ \Updelta_{\text{tr,ch}} G^{^\circ } (i) $ of α-amino acids, as obtained by subtracting theoretically computed contributions to $ \Updelta_{\text{tr}} G^{ \circ } $ due to cavity and dipole–dipole interaction effects from the corresponding experimental $ \Updelta_{\text{tr}} G^{ \circ } $ , are indicative of the superimposed effect of increased basicity and dispersion and decreased hydrophobic hydration (hbh) in DMF–water solvent mixtures as compared to those in water, while, in addition, $ T\Updelta_{\text{tr,ch}} S^{^\circ } (i) $ is guided by structural effects. The computed chemical transfer energies of the –CH2– group, $ \Updelta_{\text{tr,ch}} P^{^\circ } $ (–CH2–) [P = G or S] as obtained by subtracting the value of lower homologue from that of immediately higher homologue, are found to change with composition indicating involvement of several opposing factors in the calculation of the chemical interactions. The $ \Updelta_{\text{tr,ch}} G^{^\circ } $ (–CH2–) values are found to be guided by the decreased hydrophobic effect in DMF–water mixtures, and are indicative of the nature of the three dimensional structure of the aquo-organic solvent system around each solute.  相似文献   

17.
The simulation results from a basic polyelectrolyte chain consisting of an anionic string of 150 univalent negatively charged particles connected under various harmonic-like potential interactions with each other in the presence of a similar number of positive and free counter ions found in Jesudason et al. (EPJE $30$ 30 :341–350, 2009) forms the focal point for further discussion on chain models based on a survey of more recent developments in general polyelectrolyte theory. The topics discussed include persistence length definition, forcefields and methods of controlling simulation parameters, and thermodynamics. The data for the basic system was derived for the temperature range $0.1$ 0.1 $10.0$ 10.0 in reduced units (corresponding to $\xi =10\text{-- }0.1$ ξ = 10 -- 0.1 ); the augmented data involves a $360$ 360 monomer chain. The data include the total and Coulombic energies, radial distribution functions, radii of gyration, end-to-end distances and snapshots of the system which are all discussed anew. Polyelectrolyte systems have been overwhelmingly associated with biophysical interpretations, but here it is suggested that these detailed studies and the consequent theoretical formulations could be extended further afield; non-biological ionic liquid systems with catalytic and energy storage applications are some of many other possibilities. However, the approach used by MD simulations to validate ionic liquid systems as carriers of molecules with catalytic moieties often refer to CPMD and DFT quantum methods, which is not the current norm as judged by the literature in especially coarse grained polyelectrolyte MD. The quantum approach could also be used for more detailed analysis of biophysical systems where one trend seems to be that the incorporation of details in simulations accounts for phenomena not explicable in coarser grained MD, for instance if conventional atomic ionic charges are assigned to all atom modeling. This is illustrated by a linear chain modeling a DNA polymer using different charge and size assignments for the same linear charge density. The trends are such that it might be expected that some form of routine standardization of force fields in the spirit of the Jorgensen OPLS-AA method that incorporates quantum calculations specific to a system will be implemented as a routine as refinements are seen to lead to more comprehensive rationalization.  相似文献   

18.
The densities, ρ 12, and speeds of sound, u 12, of 1-ethyl-3-methylimidazolium tetrafluoroborate (1) + N-methylformamide or N,N-dimethylformamide (2) binary mixtures at (293.15. 298.15. 303.15, 308.15 K), and excess molar enthalpies, $ H_{12}^{\text{E}} $ H 12 E , of the same mixtures at 298.15 K have been measured over the entire mole fraction range using a density and sound analyzer (Anton Paar DSA-5000) and a 2-drop microcalorimeter, respectively. Excess molar volume, $ V_{12}^{\text{E}} $ V 12 E , and excess isentropic compressibility, $ \left( {\kappa_{S}^{\text{E}} } \right)_{12} $ ( κ S E ) 12 , values have been calculated by utilizing the measured density and speed of sound data. The observed data have been analyzed in terms of: (i) Graph theory and (ii) the Prigogine–Flory–Patterson theory. Analysis of the $ V_{12}^{\text{E}} $ V 12 E data in terms of Graph theory suggest that: (i) in pure 1-ethyl-3-methylimidazolium tetrafluoroborate, the tetrafluoroborate anion is positioned over the imidazoliun ring and there are interactions between the hydrogen atom of (C–H{edge}) and proton of the –CH3 group (imidazolium ring) with fluorine atoms of tetrafluoroborate anion, and (ii) (1 + 2) mixtures are characterized by ion–dipole interactions to form a 1:1 molecular complex. Further, the $ V_{12}^{\text{E}} $ V 12 E , $ H_{12}^{\text{E}} $ H 12 E and $ \left( {\kappa_{S}^{\text{E}} } \right)_{12} $ ( κ S E ) 12 values determined from Graph theory compare well with their measured experimental data.  相似文献   

19.
In a previous work the equilibrium geometrical and electronic structures of Xe n + clusters had been established using a non-empirical model hamiltonian. The same model is used to determine the energetic barriers between the nearly degenerate isomers; the movement of the neutral atoms around the Xe 3 + or Xe 4 + ionized linear cores are quite easy (ΔE?0.9 kcal/mole), the changes from a Xe 3 + to a Xe 4 + core are more difficult (ΔE?2.0 kcal/mole). The energetically possible fissions from a vertical photoionization \(Xe_n \xrightarrow{{h v}}Xe_n^{v + } \to Xe_p^ + + Xe_{n - p} \) forn≦19,p=1–9 and 12–14 and mass exchanges Xe p + +Xe q →Xe p+m + +Xe q?m (m=1,2,3) from relaxed Xe p + clusters are given forp+m≦9 and 12–14 andq≦19. Surprisingly the reverse reactions are shown to occur for some values ofp andq. Numerous processes lead to Xe 13 + , which is especially stable.  相似文献   

20.
A Poisson equation for nuclear motions in diatomic molecules is derived. The working formula is whereV α 2 is the Laplacian operator for the position of nucleusα, W is the Born-Oppenheimer molecular energy, is the atomic number ofα, and ? β (α) is the electronic charge density evaluated atα due to orbitals centered onβ. Harmonic, anharmonic and quartic equilibrium force constants are calculated using Hartree-Fock molecular and atomic electronic charge densities, for a number of first and second row diatomic molecules. A charge-model field gradient formula for harmonic force constants $$k_e = {3 \mathord{\left/ {\vphantom {3 {R_e^3 ,}}} \right. \kern-\nulldelimiterspace} {R_e^3 ,}}$$ wherek e is the force constant andR e the equilibrium internuclear distance, which offers general improvement over a similar formula due to Bratoz, is presented.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号