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 703 毫秒
1.
Distributions of nuclear holes (cylindrical channels of an identical diameter) in reactor track membranes were studied by computer simulation taking into account the angular spread of an ion beam. The factors of angular decrease in the number of multiple-hole channels and their dependence on the overlap multiplicity and membrane porosity P at P = 0.01–0.30 and hole multiplicity of m < 4 were obtained. The simulation results were compared with predictions made using analytical models.  相似文献   

2.
The classical problem of overlapping circular holes of equal diameters on one surface of a nuclear track membrane was considered. Using the Monte Carlo method, the probabilities of overlapping W k(P) of nuclear pores with a multiplicity of k < 32 at a porosity of P < 0.30 were determined and analytical expressions for their approximation were derived. The results were analyzed in terms of an approach based on a kinetic equation for a discrete Markov process. It was shown that the W k(P) values obtained for multiple pores with a multiplicity of k < 7 are consistent with the theory if two-step processes are taken into account.  相似文献   

3.
Cationically charged poly(allylamine) (PAA) membranes having various water contents [0.49 < H < 0.63 (g H2O/g wet membrane)] were prepared. Sorption and permeation of simple salts (sodium chloride and sodium tetraphenylborate) were investigated, taking into account the state of the water in these membranes. The weight ratios of freezable water and free water to total water (Wfz/Wt, and Wf/Wt) in the membranes were estimated by means of DSC and pulsed 1H-NMR measurements, respectively. Partition coefficients K for total water were converted into those in freezable and free water, Kfz and Kf, using Wfz/Wt and Wf/Wt. The permeability of both salts in the membranes could be interpreted satisfactorily by an equation derived from the Teorell-Meyer-Sievers theory using values of Kf. The free water is mainly involved in the permeation of simple salts through PAA membranes while bound water hardly takes part.  相似文献   

4.
Relationships between the molecular structures and zero-field splitting parameters of quintet m-phenylenedinitrenes formed during the photolysis of 1,3-diazidobenzenes in 2-methyltetrahydrofuran solutions frozen at 77 K were studied by ESR spectroscopy and B3LYP/6-31G* calculations. Simulations of the W−2/W−1, W+2/W+1, W−1/W0, W+1/W0, and W+2/W0m s = 2) transitions were performed for the first time and all signals in the ESR spectra of quintet m-phenylenedinitrenes were unambiguously assigned. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 7, pp. 1085–1089, July, 2006.  相似文献   

5.
Electrospray mass spectrometry (ESMS) has been conducted on aqueous solutions of isopolyoxotungstate systems. There is direct evidence that the desorption process in the ESMS technique has resulted in significant chemical effects, resulting in the detection of many new anions and cations. For the ammonium polyoxotungstate system, negative-ion ESMS yields ions of the form [HW m O3m+1], [W m O3m+1]2–, and [HW m O3m+2]3– (with the latter being better formulated as [H2W2m O6m+4]6–). For the alkali metal polyoxotungstate systems ions of the form [W m O3m+1A] and [W m O4m A2m–2]2– (where A=Li+, Na+, K+) were observed. For positive-ion ESMS two series were observed, namely, the [W m O4m A2m+1]+ and [W m O4m A2m+2]2+ ions. In the ammonium polyoxotungstate system, aggregates of both the [HW m O3m+1] and the [W m O3m+1]2– series can be classified as open-chained structures of tetrahedra that are corner shared, whereas the more highly charged anions [H2W2m O6m+4]6– are consistent with closed-packed structures which are based on the structure of paratungstate-B [H2W12O42]10–. For the alkali metal tungstate systems, the ESMS spectra are consistent with open-chained structures of octahedral units that are edge shared, with a terminating tetrahedral unit. Linear correlations suggest that the assembly of these aggregates occurs via an additive polymerization mechanism for which the additive moieties (WO3, WO2+ 2, and W2O8A4) in aqueous solution can be identified.  相似文献   

6.
The superposition of nanosized pyrocarbon crystallites (NPCs; an NPC superposition is n sequentially formed pyrocarbon nanocrystallites each consisting of m monolayers) was studied. NPCs were obtained by the pyrolysis of methane and deposited on a porous surface of Trumem ultrafiltration membranes with D pore = 50 and 90 nm. A new uncoupled layer effect was found. After n-fold NPC superposition of an even number of monolayers (the total number throughout the superposition; i.e., m 1 + m 2 + ... + m n is even), further deposition of pyrocarbon nanocrystallites occurs in accordance with the previously determined kinetics of topochemical dehydrogenation of methane. However, if an NPC superposition is formed from an odd number of monolayers, the reaction rate constant decreases by more than one order of magnitude. A comparative analysis of the hydrogen adsorption ability of two carbon structures (NPC superposition and orientated carbon nanotubes of graphenes (OCNTGs)) showed that OCNTGs adsorbed up to ∼14% (relative to their mass) of hydrogen, whereas NPCs did not. It was shown for the first time that hydrogen adsorbed in OCNTG affected the transport properties of membranes, decreasing their performance with respect to liquids by a factor of 4–26.  相似文献   

7.
Calorimetric measurements were carried out on the electrorefining of copper using different current densities with a Calvet type microcalorimeter at room temperature. The ratio (R) of the measured heat (Q m orW m) to the input electric energy (Q in orW in) and the excess heat (Q ex orW ex), i.e. the difference betweenQ m (orW m) andQ in (orW in) during the electrorefining process were discussed in terms of general thermodynamics. It was found thatR andQ ex were related to the current density employed in the experiment and varied as a logarithmic function. The results obtained here indicate that the heat generation under different conditions, such as different currents or voltages, may be caused partially by the irreversibility of the process or by some unknown processes.Dedicated to Prof. Menachem Steinberg on the occasion of his 65th birthdayThe authors would like to acknowledge the extreme encouragements and help of Professor Shuyi Liu (University of Science and Technology of China) and Professors Fu Tan and Guoquan Liu (Institute of Chemistry, Academia Sinica).This study was supported by the National Nature Science Foundation of China.  相似文献   

8.
5‐Coordinated methoxybenzylidene complexes M(=NAr)(=CH?C6H4?o‐OMe)(OtBuF3)2 (Ar=2,6‐iPr2C6H3; tBuF3=CMe2(CF3)) of Mo ( 1mMo ) and W ( 1mW ) were synthesized by cross‐metathesis from the corresponding neophylidene/neopentylidene precursors and o‐methoxystyrene. 1mMo and 1mW were grafted onto the surface of silica partially dehydroxylated at 700 °C to give well‐defined silica‐supported alkylidenes (≡SiO)M(=NAr)(=CH?C6H4?o‐OMe)(OtBuF3) (M=Mo ( 1Mo ), W ( 1W )). Supported methoxybenzylidene complexes were tested in metathesis of cis‐4‐nonene, 1‐nonene, and ethyl oleate, and compared to their molecular precursors and supported classical analogs (≡SiO)M(=NAr)(=CHCMe2R)(OtBuF3) (M=Mo, R=Ph ( 2Mo ), M=W, R=Me ( 2W )). Both grafted complexes 1Mo and 1W show significantly better performance as compared to their molecular precursors 1mMo and 1mW but are less efficient than the classical 4‐coordinated alkylidenes 2Mo and 2W . Noteworthy, both 1Mo and 1W can reach equilibrium conversion in metathesis of cis‐4‐nonene at catalyst loadings as low as 50 ppm.  相似文献   

9.
Plasma polymerizations of ethylene and tetrafluoroethylene are compared. In the plasma polymerization of ethylene and of tetrafluoroethylene, glow characteristics play an important role. Glow characteristic is dependent on a combined factor of W/Fm, where W is discharge power and Fm is monomer flow rate. At higher flow rates, higher wattages are required to maintain “full glow.” In the plasma polymerization of tetrafluoroethylene, simultaneous decomposition of the monomer competes with plasma polymerization. Above a certain value of W/Fm, decomposition becomes the predominant reaction, and the polymer deposition rate decreases with increasing discharge power. ESCA results indicate that the plasma polymer of tetrafluoroethylene that is formed in an incomplete glow region (low W/Fm) is a hybrid of polymers of plasma polymerization and of plasma-induced polymerization of the monomer. Polymers formed under conditions of high W/Fm to produce “full glow” are similar, regardless of the extent of decomposition of the monomer. They contain carbons with different numbers of F(CF3, ? CF2? , >CF? , >C<) and carbons bonded to other more electronegative substituents.  相似文献   

10.
The tangents to the evaporation path curves in the W/O microemulsion base of water, (W), pentanol, (P), and sodium dodecyl sulfate, (S), were extended to the W/P axis to establish the relative composition, (WV, PV), of the vapor leaving the liquid.The composition of the vapor, with which the microemulsion is in contact includes also the contribution from the relative humidity of the surrounding atmosphere. The difference between the composition of these gases is clarified using the algebraic expressions from the phase diagram, but the quantitative composition of the equilibrium vapor is not available without further numerical information. The limits of the vapor for evaporation direction under different relative humidities were clarified.  相似文献   

11.
Infrared spectra of CO2 sorbed in rubbery and glassy polymers were measured to examine the relationships between the spectroscopic data and physical properties of the polymeric membranes. The “V-shape” tendency in the plot of W1 [i.e., half-width of CO2 peak sorbed in the membranes] vs glass-transition temperature (Tg) is observed, and has exactly the same tendency that is widely known from the plot of log D (diffusion coefficient) vs Tg. It is suggested that the membranes having a wider W1 give a faster diffusion coefficient, since W1 is inversely related to the moment of inertia of CO2 in the membranes. Two distinct peaks of CO2 were not observed in the infrared spectra of CO2 sorbed in the glassy polymers. This suggests that the states of CO2 in the Henry mode and Langmuir mode in the glassy polymers are similar in the spectroscopic measurements. © 1994 John Wiley & Sons, Inc.  相似文献   

12.
We report the time‐resolved supramolecular assembly of a series of nanoscale polyoxometalate clusters (from the same one‐pot reaction) of the form: [H(10+m)Ag18Cl(Te3W38O134)2]n, where n=1 and m=0 for compound 1 (after 4 days), n=2 and m=3 for compound 2 (after 10 days), and n=∞ and m=5 for compound 3 (after 14 days). The reaction is based upon the self‐organization of two {Te3W38} units around a single chloride template and the formation of a {Ag12} cluster, giving a {Ag12}‐in‐{W76} cluster‐in‐cluster in compound 1 , which further aggregates to cluster compounds 2 and 3 by supramolecular Ag‐POM interactions. The proposed mechanism for the formation of the clusters has been studied by ESI‐MS. Further, control experiments demonstrate the crucial role that TeO32?, Cl?, and Ag+ play in the self‐assembly of compounds 1 – 3 .  相似文献   

13.
The equilibrium topology of an aqueous Janus emulsion of two oils, O1 and O2, with water, W, [(O1+O2)/W], is numerically evaluated with the following realistic interfacial tensions (γ): γO2/W=5 mN m−1, γO1/O2=1 mN m−1, and γO1/W varies within the range 4–5 mN m−1, which is the limiting range for stable Janus drop topology. The relative significance of the two independently pivotal factors for the topology is evaluated, that is, the local equilibrium at the line of contact between the three liquids and the volume fraction of the two dispersed liquids within the drop. The results reveal a dominant effect of the local equilibrium on the fraction of the O2 drop surface that is covered by O1. In contrast, for a constant volume of O2, the impact of the interfacial tension balance on the limit of the coverage is modest for an infinite volume of O1. Interestingly, when the O1 volume exceeds this value, an emulsion inversion occurs, and the O1 portion of the (O1+O2)/W topology becomes a continuous phase, generating a (W+O2)/O1 Janus configuration.  相似文献   

14.
Two kinds of hydrocarbon type monomers and three kinds of organosilicons were polymerized by a low‐temperature cascade arc argon plasma torch. Their deposition behaviors were studied as a function of experimental parameters and monomer elemental compositions. It was found that the normalized deposition rate (DR), expressed as deposition yield of DR/(FM)m, was determined by a composite operational parameter, W*(FM)c/(FM)m, where W is the power input, and (FM)c and (FM)m are the mass flow rates of carrier gases and monomers, respectively. Experimental results indicated that the deposition yield is highly dependent on the elementary compositions of monomers. Optical emission spectroscopy study on the argon plasma torch showed that the emission intensity of excited argon neutrals was proportional to the value of the parameter W*(FM)c. These results further certified that excited argon neutrals are the main energy carriers from the cascade arc column to activate monomers in the argon plasma torch. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 967–982, 1999  相似文献   

15.
Permeability coefficients P of KCl through porous cellulose triacetate (CTA) membranes were measured as a function of the water volume fraction VW and diffusion coefficients D were determined using solubility parameters K and a membrane thickness d from the relationship of P = KD/d. D increased with an increase in VW. D especially increases abruptly around VW = 0.5, which corresponds to 2% triethylene glycol (TEG) content. The percolation theory was applied to the experimental results under the conditions DA = D (VW = 1) = 1.8 × 10−5 cm2 s−1, DB = D(VW = 0) = 1.8 × 10−8 cm2 s−1, coordination number (Z) = 2.5, 3, 3.5, and 4, and packing fraction f = 1.0. A good fit was obtained at Z = 3.5 because the experimental and calculated results also shifted at the same VW below VW = 0.5. It is suggested that a phase inversion, that is, change of a discontinuous water phase to a continuous water phase, occurs around VW = 0.5. Above VW = 0.5, the experimental results agree well with the calculated line for Z = 3 or Z = 2.5 which means that the coordination numbers decrease with an increase in water content. It is thought that VW is overestimated because it is hard to completely wipe off the excess water quickly from the membrane surface. Z = 3.5 means that a pore can connect in 3.5 directions.  相似文献   

16.
The light scattering intensity distribution from rodlike crystalline superstructures is quantitatively investigated theoretically and experimentally. The arithmetic average of theoretical Hv scattered intensities at azimuthal angle μ = 0° and μ = 45° is shown to decrease with increasing scattering angle θ in proportion to W?1 at high scattering angles for a system composed of a random assembly of rodlike superstructure having very small lateral dimensions relative to the length. The quantity W is defined as 2π(L/λ) sin θ where L is the length of the rod, and λ is the wavelength of light in the medium. A method is proposed to estimate the length L by using the W?1 dependence. Effects of internal heterogenity, polydispersity in rod length, and finite lateral dimensions of the rodlike superstructure are considered to account for experimental deviation of the scattered intensity distributions from the W?1 dependence. The effect of finite lateral dimensions turns out to be the most important.  相似文献   

17.
The density, viscosity, and ultrasonic velocity of solutions of two schiff bases in 1,4-dioxane and dimethylformamide (DMF) were measured at 318.15 K. Various acoustical properties, such as the specific impedance (Z), isentropic compressibility (κs), Rao’s molar sound function (R m), van der Waals constant (b), molar compressibility (W), intermolecular free length (L f), relaxation strength (r), relative association (R A), and free volume (V f), were calculated. The results were interpreted in terms of molecular interactions occurring in the solutions. Published in Russian in Zhurnal Fizicheskoi Khimii, 2006, Vol. 80, No. 7, pp. 1206–1210. The text was submitted by the authors in English.  相似文献   

18.
Based on our earlier three-dimensional DWBA theory, we discuss angular distributions and the roles of various angular momenta. The theory predicts a relatively small number of partial waves, backward scattering, and broadening of angular distributions with increased energy, for the reaction F+ H2 (va = 0, ja = 0) → HF (vb = 2 jb = 0) + H  相似文献   

19.
Summary Van der Waals’ volumes (V W) and surface areas (S W) of alkanes, (E)-azoalkanes and structurally similar alkenes (R1-X=X-R2, X=N, CH) were calculated by a semiempirical quantum-chemical method (AM1). The calculated data are in reasonable agreement with the experimental values of Bondi and good correlations were found between the calculated data and Kovats’ retention indices (I R). While theV Ws of alkanes with the same carbon number are very close to one another, theS Ws follow the scatter of theI R values for branched alkanes. The difference in theI R of (E)-azo compounds and the structurally similar alkenes can be explained by the difference inV Ws.  相似文献   

20.
The thermal degradation of an amphiphilic block copolymer poly(ethylene)-b-poly(ethylene oxide)-carboxylic acid terminated (PE-b-80%PEO–CH2COOH) and its salt obtained as intermediary product from chemical oxidation of the end group of poly(ethylene)-b-poly(ethylene oxide) (PE-b-80%PEO) has been studied using a thermogravimetric mass spectrometry (TG/MS) coupled system. The isothermal fragmentation of PE-b-80%PEO–CH2COOH showed a more complex fragmentation pattern than PE-b-80%PEO owing to the simultaneous occurrence of the polyether block and the carboxylic end group fragmentations. This led to the appearance of four overlapping ion current peaks of fragments with m/z 44 and two peaks relative to m/z 18 at different times by acid-terminated copolymer. For the PE-b-80%PEO copolymer, two ion current peaks associated to m/z 44 and one large peak relative to m/z 18 fragments were detected. The intermediary product (PE-b-80%PEO–CH2COO K+) showed differences related to the fragmentation behavior. It has more defined ion current signals and presented characteristic peaks attributed to m/z 43 fragment at the very beginning of the thermal degradation process, which it not detected in the acid copolymer.  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