共查询到20条相似文献,搜索用时 0 毫秒
1.
The effect of solvation by a continuum dielectric on the size of an ion is examined using electronic structure calculations. Various measures correlated with size are considered, including the root-mean-square radius of the electronic charge density, the amount of solute charge penetrating outside the cavity, the electronic radial distribution function, the nucleus-electron potential energy, and the electron-electron potential energy. Calculations are made on several representative ionic solutes, and it is found that every measure indicates that the application of a dielectric makes the cations larger and the anions smaller. These counterintuitive trends are examined, and a plausible explanation is offered for the observed behavior. 相似文献
2.
Sayari A Belmabkhout Y Da'na E 《Langmuir : the ACS journal of surfaces and colloids》2012,28(9):4241-4247
Adsorption of CO(2) was investigated on a series of primary, secondary, and tertiary monoamine-grafted pore-expanded mesoporous MCM-41 silicas, referred to as pMONO, sMONO, and tMONO, respectively. The pMONO adsorbent showed the highest CO(2) adsorption capacity, followed by sMONO, whereas tMONO exhibited hardly any CO(2) uptake. As for the stability in the presence of dry CO(2), we showed in a previous contribution [J. Am. Chem. Soc.2010, 132, 6312-6314] that amine-supported materials deactivate in the presence of dry CO(2) via the formation of urea linkages. Here, we showed that only primary amines suffered extensive loss in CO(2) uptake, whereas secondary and tertiary amines were stable even at temperature as high as 200 °C. The difference in the stability of primary vs secondary and tertiary amines was associated with the occurrence of isocyanate as intermediate species toward the formation of urea groups, since only primary amines can be precursors to isocyanate in the presence of CO(2). However, using a grafted propyldiethylenetriamine containing both primary and secondary amines, we demonstrated that while primary amines gave rise to isocyanate, the latter can react with either primary or secondary amines to generate di- and trisubstituted ureas, leading to deactivation of secondary amines as well. 相似文献
3.
We have investigated the performance of DFT in U(VI) chemistry. A large, representative selection of functionals has been tested, in combination with two ECPs developed in Stuttgart that have different-sized cores (60 and 78 electrons for U). In addition, several tests were undertaken with another 14 electron pseudopotential, which was developed in Los Alamos. The experimental database contained vibrational wavenumbers, thermochemical data, and (19)F chemical shifts for molecules of the type UF(6-n)Cl(n). For the prediction of vibrational wavenumbers, the large-core RECP (14 electrons) gives results that are at least as good as those obtained with the small-core RECP (32 electrons). GGA functionals are as successful as hybrid GGA for vibrational spectroscopy; typical errors are only a few percent with the Stuttgart pseudopotentials. For thermochemistry, hybrid versions of DFT are more successful than GGA, LDA, or meta-GGA. Marginally better results are obtained with a 32 electron ECP than with 14; since the experimental uncertainties are at least 25 kJ/mol for each reaction, the best functionals give results that are essentially indistinguishable from experiment. However, large-basis CCSD(T) results match experiment better than any DFT that we examined. Our findings for NMR spectroscopy are rather disappointing; no combination of pseudopotential, functional, and basis yields even a qualitatively correct prediction of trends in the (19)F chemical shifts of UF(6-n)Cl(n) species. Results yielded by the large-core RECP are, in general, slightly less bad than those obtained with the small core. We conclude that DFT cannot be recommended for predictions of NMR spectra in this series of compounds, though this conclusion should not be generalized. Our most important result concerns the good performance of the large-core Stuttgart pseudopotential. Given its computational efficiency, we recommend that it be used with DFT methods for the prediction of molecular geometries, vibrational frequencies, and thermochemistry of a given oxidation state. The hybrid GGA functionals MPW1PW91 and PBE0 give the best results overall. 相似文献
4.
5.
6.
Dupin D Armes SP Connan C Reeve P Baxter SM 《Langmuir : the ACS journal of surfaces and colloids》2007,23(13):6903-6910
A near-monodisperse styrene-functionalized poly[2-(dimethylamino)ethyl methacrylate] (PDMA) macromonomer was evaluated as a reactive steric stabilizer for the preparation of poly(2-vinylpyridine (P2VP) latexes via emulsion polymerization. The solution pH was shown to be a critical parameter for successful syntheses: stable latexes with minimal coagulum were only obtained at (or above) neutral pH. The presence of the grafted PDMA stabilizer in a near-monodisperse P2VP latex of 280 nm diameter was indicated by FT-IR spectroscopy and quantified at 6.0 wt % using 1H NMR spectroscopy. XPS studies confirmed that this stabilizer was located at the latex surface, as expected. Combined DLS and electrophoretic data indicated that these PDMA-P2VP particles exist in three states depending on the solution pH: swollen cationic microgels were obtained below pH 4.1, nonsolvated latex particles with a cationic stabilizer layer were obtained at intermediate pH, and flocculated latex particles with neutral PDMA stabilizer chains were obtained at around pH 8.5. Finally, this PDMA-P2VP latex was shown to be a superior Pickering emulsifier for stabilizing water-in-1-undecanol emulsions than either a poly(ethylene glycol)-stabilized P2VP latex or a charge-stabilized P2VP latex. This serves to illustrate the important role played by the steric stabilizer in determining particle wettability. 相似文献
7.
Monte Carlo simulation has been used to investigate the effects of linear solvent molecular size on polymer chain conformation in solutions. Increasing the solvent molecular size leads to shrinkage of the polymer chains and increase of the critical overlap concentrations. The root-mean-square radius of gyration of polymer chains (R(g)) is less sensitive to the variation of polymer concentration in solutions of larger solvent molecules. In addition, the dependency of R(g) on polymer concentration under normal solvent conditions and solvent molecular size is in good agreement with scaling laws. When the solvent molecular size approaches the ideal end-to-end distance of the polymer chain, an extra aggregation of polymer chains occurs, and the solvent becomes the so-called medium-sized solvent. When the size of solvent molecules is smaller than the medium size, the polymer chains are swollen or partially swollen. However, when the size of solvent molecules is larger than the medium size, the polymer coils shrink and segregate, enwrapped by the large solvent molecules. 相似文献
8.
The authors study how the simultaneous presence of short-range secondary and long-range tertiary interactions controls the folding and collapse behavior of a helical macromolecule. The secondary interactions stabilize the helical conformation of the chain, while the tertiary interactions govern its overall three-dimensional shape. The authors have carried out Monte Carlo simulations to study the effect of chain length on the folding and collapse behavior of the chain. They have calculated state diagrams for four chain lengths and found that the physics is very rich with a plethora of stable conformational states. In addition to the helix-coil and coil-globule transitions, their model describes the coupling between them which takes place at low temperatures. Under these conditions, their model predicts a cascade of continuous, conformational transitions between states with an increase in the strength of the tertiary interactions. During each transition the chain shrinks, i.e., collapses, in a rapid and specific manner. In addition, the number of the transitions increases with increasing chain length. They have also found that the low-temperature regions of the state diagram between the transition lines cannot be associated with specific structures of the chain, but rather, with ensembles of various configurations of the chain with similar characteristics. Based on these results the authors propose a mechanism for the folding and collapse of helical macromolecules which is further supported by the analysis of configurational, configurational, and thermodynamic properties of the chain. 相似文献
9.
Kirichenko Olga Kapustin Gennadiy Nissenbaum Vera Strelkova Anna Shuvalova Elena Shesterkina Anastasiya Kustov Leonid 《Journal of Thermal Analysis and Calorimetry》2018,132(1):233-249
Journal of Thermal Analysis and Calorimetry - Poly(butylene succinate) (PBS) nanocomposites filled with nanoprecipitated calcium carbonate (NPCC) were prepared via melt blending. The hybrid... 相似文献
10.
Hoang Tri Hai Hiroaki Kura Migaku Takahashi Tomoyuki Ogawa 《Journal of colloid and interface science》2010,341(1):194-199
A phase transformation induced by the reduction of as-synthesized γ-maghemite (γ-Fe2O3) nanoparticles was performed in solution by exploiting the reservoir of reduction gas (CO) generated from the incomplete combustion reaction of organic substances in the reactor. Results from X-ray diffraction, color indicator, and magnetic analysis using a SQUID strongly support this phase transformation. Based on this route, monodisperse magnetite (Fe3O4) nanoparticles were simply produced in the range from 260 to 300 °C. Almost all aspects of the original γ-Fe2O3 nanoparticles, such as shape, size, and monodispersity, were maintained in the produced Fe3O4 nanoparticles. 相似文献
11.
The analysis of various models assumed to represent the influence of pH on heavy metals biosorption equilibrium is presented. It shows that all of them lead to the same mathematical expressions (e.g. the Langmuir or the Flory adsorption isotherm equations) when the pH effects are neglected. Even if considering the pH effects, some of them (competitive adsorption and ion-exchange models, for instance) still lead to analogical expressions for sorption isotherm equations. The accepted mechanism of biosorption may, however, influence strongly the differences between the initial and equilibrium states of biosorption system. 相似文献
12.
N. Barros J. Salgado M. Villanueva J. Rodriquez-Añón J. Proupin S. Feijóo M. Martín-Pastor 《Journal of Thermal Analysis and Calorimetry》2011,104(1):53-60
The environmental concern on soil exploitation, linked to global warming by the Kyoto protocol, is responsible for increasing
interest in the understanding of the role of the composition and structure of the soil organic matter (SOM) on soil carbon,
C, dynamics. Thermal analysis and nuclear magnetic resonance (NMR) are applied to study the thermal properties, the structure
and composition of the SOM of six samples with different C contents in order to improve the interpretation of results given
by thermal analysis. Results showed that the direct integral of the combustion peaks obtained by DSC and the percentage of
SOM given by TG were both directly related to the quantity of total soil C. Thus, soils with higher C content showed higher
energy content too. The combustion temperatures of the curves given by DSC are those reported for labile OM. NMR results indicated
the presence of aliphatic C, carbohydrates, and a weak signal in the aromatic C band in all the samples that was not detected
in the DSC curves. Only two samples showed carboxyl/carbonyl C which was not detected by DSC also. 相似文献
13.
Linder DP Rodgers KR Banister J Wyllie GR Ellison MK Scheidt WR 《Journal of the American Chemical Society》2004,126(43):14136-14148
Recent years have seen dramatic growth in our understanding of the biological roles of nitric oxide (NO). Yet, the fundamental underpinnings of its reactivities with transition metal centers in proteins and enzymes, the stabilities of their structures, and the relationships between structure and reactivity remains, to a significant extent, elusive. This is especially true for the so-called ferric heme nitrosyls ([FeNO](6) in the Enemark-Feltham scheme). The Fe-CO and C-O bond strengths in the isoelectronic ferrous carbonyl complexes are widely recognized to be inversely correlated and sensitive to structural, environmental, and electronic factors. On the other hand, the Fe-NO and N-O bonds in [FeNO](6) heme complexes exhibit seemingly inconsistent behavior in response to varying structure and environment. This report contains resonance Raman and density functional theory results that suggest a new model for FeNO bonding in five-coordinate [FeNO](6) complexes. On the basis of resonance Raman and FTIR data, a direct correlation between the nu(Fe)(-)(NO) and nu(N)(-)(O) frequencies of [Fe(OEP)NO](ClO(4)) and [Fe(OEP)NO](ClO(4)).CHCl(3) (two crystal forms of the same complex) has been established. Density functional theory calculations show that the relationship between Fe-NO and N-O bond strengths is responsive to FeNO electron density in three molecular orbitals. The highest energy orbital of the three is sigma-antibonding with respect to the entire FeNO unit. The other two comprise a lower-energy, degenerate, or nearly degenerate pair that is pi-bonding with respect to Fe-NO and pi-antibonding with respect to N-O. The relative sensitivities of the electron density distributions in these orbitals are shown to be consistent with all published indicators of Fe-N-O bond strengths and angles, including the examples reported here. 相似文献
14.
Boehmite (BO) nanoparticles (NPs) were prepared via the injection of aqueous NaOH solution to aqueous aluminum nitrate solution at room temperature. Afterwards, a new complex of copper was immobilized on BO-NPs (Cu-Guanidine@BO-NPs). This heterogeneous nanocatalyst was used as a practical, recyclable, chemo and homoselective nanocatalyst in the organic processes, i.e. the preparation of tetrazole five-membered heterocycles and chemoselective sulfoxidation of sulfides using H2O2 as oxidant. In this sense, the prepared nanocatalyst was characterized by AAS, N2 adsorption–desorption isotherms, WDX, EDS, SEM, and TGA techniques. The reusability of this catalyst was investigated in the described organic reactions for several runs without notable loss of its catalytic activity. Moreover, all of the tetrazole and sulfoxide derivatives were isolated in high Turn Over Number (TON) and Turn Over Frequency (TOF) numbers indicating the high activity and selectivity of Cu-Guanidine@BO-NPs in the described reactions. 相似文献
15.
Graziano G 《Physical chemistry chemical physics : PCCP》2011,13(39):17689-17695
Trimethylamine N-oxide, TMAO, stabilizes globular proteins and is able to counteract the denaturing activity of urea. The mechanism of this counteraction has remained elusive up to now. A rationalization is proposed grounded on the same theoretical model used to clarify the origin of cold denaturation, and the denaturing activity of GdmCl versus the stabilizing one of Gdm(2)SO(4) [G. Graziano, Phys. Chem. Chem. Phys., 2010, 12, 14245-14252; G. Graziano, Phys. Chem. Chem. Phys., 2011, 13, 12008-12014]. The fundamental quantities are: (a) the difference in the solvent-excluded volume on passing from the N-state to the D-state, calculated in water and in aqueous osmolyte solution; (b) the difference in energetic attractions of the N-state and the D-state with the surrounding solvent molecules, calculated in water and in aqueous osmolyte solution. In aqueous 8 M urea + 4 M TMAO solution, the first quantity is so large and positive to counteract the second one that is large and negative due to preferential binding of urea molecules to the protein surface. This happens because aqueous 8 M urea + 4 M TMAO solution has a volume packing density markedly larger than that of water, rendering the cavity creation process much more costly. The volume packing density increase reflects the strength of the attractions of water molecules with both urea and TMAO molecules. This mechanism readily explains why TMAO counteraction is operative even though urea molecules are preferentially located on the protein surface. 相似文献
16.
Monica Spinu Nuno Dos Santos Nicolas Le Moigne Patrick Navard 《Cellulose (London, England)》2011,18(2):247-256
The swelling and dissolution capacity of dried and never-dried hardwood and softwood pulps and cotton linters was compared
in two aqueous solvents, N-methylmorpholine-N-oxide (NMMO)-water at 90 °C with water contents ranging from 16 to 22% and NaOH—water at −6 °C with NaOH contents ranging
from 5 to 8%. Swelling and dissolution mechanisms were observed by optical microscopy and dissolution efficiency was evaluated
by recovering insoluble fractions. The results show a contrasted picture towards the effect of the never-dried state on the
swelling and the dissolution capacity depending on the origin of the fibres and the type of aqueous solvent. In the case of
NMMO—water, the presence of water within and around the fibre does not seem to favour dissolution initiation but after 2 h
of mixing the dissolution yield appears to be similar for either dried or never-dried state. The limiting factor for dissolution
in NMMO—water is not the penetration of the solvent inside the cellulose fibres, but only the local concentration of NMMO
molecules around the fibre. For NaOH—water, both optical microscopy observations on individual fibres and dissolution yield
measurements show that the never-dried state is more reactive for softwood pulps and cotton linters and has no significant
effect on hardwood pulps. In this case, the local decrease of solvent strength is counteracted by the opening of the structure
in the never-dried state which should enable the Na+ hydrated ions to penetrate easier. 相似文献
17.
This paper compares the kinetics of exchanges of phenylethanethiolate ligands (PhC2S-) of the monolayer-protected clusters (MPCs) Au(38)(SC2Ph)(24) and Au(140)(SC2Ph)(53) with p-substituted arylthiols (p-X-PhSH), where X = NO(2), Br, CH(3), OCH(3), and OH. First-order rate constants at 293 K for exchange of the first ca. 25% of the ligands on the molecule-like Au(38)(SC2Ph)(24) MPC, measured using (1)H NMR, vary linearly with the in-coming arythiol concentration; ligand exchange is an overall second-order reaction. Remarkably, the second-order rate constants for ligand exchange on Au(38)(SC2Ph)(24) are very close to those of corresponding exchange reactions on the larger nanoparticle Au(140)(SC2Ph)(53) MPCs. These are the first results that quantitatively show that the chemical reactivity of different sized nanocrystals is almost independent of size; presumably, this is because the locus of the initial ligand exchanges is a common kind of site, thought to be the nanocrystal vertexes. The rates of later stages of exchange (beyond ca. 25%) differ for Au(38) and Au(140) cores, the latter being much slower presumably due to its larger terrace-like surface atom content. The reverse exchange reaction was studied for Au(38)(p-X-arylthiolate)(24) MPCs (X = NO(2), Br, and CH(3)), where the in-coming ligand is now phenylethanethiol. Remarkably, the rate constants of both forward and reverse exchanges display identical substituent effects, which implies a concurrent bonding of both in-coming and leaving ligands to the Au core in the rate-determining step, as in an associative mechanism. X = NO(2) gives the fastest rates, and the ratio of forward and reverse rate constants gives an equilibrium constant of K(EQ,PE) = 4.0 that is independent of X. 相似文献
18.
Gilberto Siqueira Carole Fraschini Julien Bras Alain Dufresne Robert Prud’homme Marie-Pierre Laborie 《European Polymer Journal》2011,47(12):2216-2227
Polycaprolactone (PCL)/cellulose nanocomposites were prepared by mixing PCL with surface modified sisal nanowhiskers (CNW) and microfibrillated cellulose (MFC) extracted from sisal fibers. The influence of cellulosic nanoparticles on the crystallization behavior of PCL was investigated by differential scanning calorimetry. Isothermal crystallization data were modeled with Avrami’s kinetics, Lauritzen–Hoffman secondary nucleation theory and equilibrium melting points were determined with the Hoffman–Weeks method. The cellulose nanoparticles, acting as nucleating agents, drastically accelerate the crystallization of PCL while depressing its equilibrium melting by 9–10 °C. The crystallization of MFC-nanocomposites is slightly faster than that of CNW-nanocomposites, in agreement with the slightly lower bulk activation energy for crystallization and nucleation parameter in the former. The results are discussed based on the differences of specific surface area and surface chemistry of nanoparticles, as well as the confinement phenomenon. 相似文献
19.
20.
Alireza Farrokhi Farzaneh Bivareh Saeideh Dejbakhshpour Ali Zeraatkar Moghaddam 《应用有机金属化学》2020,34(11):e5938
The photocatalytic reduction of Cr (VI) is investigated over phosphonate-based metal–organic frameworks (MOFs) structured as STA-12(M)(M = Mn, Fe, Co, Ni). The removal of Cr (VI) appears to be faster with STA-12(Fe) under natural sunlight. The correlation among responses to effective variables as main and interactions were optimized by central composite design (CCD) in response surface methodology (RSM). Moreover, STA-12(Fe) has exhibited considerable synergistic photocatalytic activity for dyes (MO and RhB) degradation and Cr (VI) reduction. The reduction ratio of Cr (VI) was increased extremely after addition of MO or RhB. The highest photocatalytic activity of dyes degradation and Cr (VI) reduction appeared under the dye/Cr (VI) weight ratio of 3:1. To determine the most important species that affected the photocatalytic reduction, trapping experiments were carried out, using various kinds of scavenger species. Finally, a probable reaction mechanism has been investigated in detail. 相似文献