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 diffusion deposition of point particles from a Stokes–Brinkman transverse stationary flow in a model monolayer membrane composed of contacting spherical hollow grains (capsules) with porous permeable shells formed from nanoparticles, is calculated. Monolayers with square and hexagonal packings of the grains are considered. Approximation equations are constructed for the dependences of grain drag force on shell thickness, Brinkman permeability parameter S, and internal shell radius ξ. Efficiencies of point particle collection on the grains are calculated as depending on the Peclet number, S, and ξ, and it is shown that layers of hollow permeable grains possess the highest filtration performance criterion among layers of impermeable and permeable uniform porous grains provided that the zero-concentration boundary condition is fulfilled at the outer radius of the grain.  相似文献   

2.
Uniform spreading of oil on solid surfaces is important in many processes where proper lubrication is required and this can be controlled using surfactants. The role of oil–solid interfacial self-assembled surfactant structure (SASS) in oil spreading is examined in this study for the case of hexadecane-surfactant droplet spreading on a flat horizontal copper surface, with triphenyl phosphorothionate surfactants having varying chain lengths (0 to 9). It is shown that the frictional forces (FSASS) as determined by the SASS regulate droplet spreading rate according to surfactant chain length; surfactants with longer chains led to higher reduction in the spreading rate. The extent of such forces, FSASS, depends on the surfactant density of the evolving SASS, and specific configuration the evolving SASS exhibit as per the orientations of the surfactant chains therein. Thus, FSASS = [k1 + k2(t)] Γδ(t), where Γδ(t) is the surfactant adsorption density of SASS at time ‘t’ during evolution, and, k1 and k2(t) are the force coefficients for Γδ(t) and orientations (as a function of spreading time) of the surfactant chains respectively. As a SASS evolves/grows along with adsorption of surfactants at the spreading induced fresh interface, the k1Γδ(t) component of FSASS increases and contributes to reduction in the net spreading force (S). With a decrease in the net spreading force, the existence of a cross-over period, during which the transition of the spatial dynamics of the chains from disordered to realignment/packing induced ordered orientation occurs, has been inferred from the FSASS vs. chain length relationships. Such relationships also suggested that the rate of realignment/packing is increased progressively particularly due the realignment/packing induced decrease in the net spreading force. Therefore, the realignment process is a self-induced process, which spans a measurable period of time (several minutes), the cross-over period, during which the net spreading force decreases essentially due to such self-induced process.  相似文献   

3.
We studied interaction of hydrophilic polymer chain and hydrophilic silica nanoparticles in a dilute aqueous system using an idealized model system comprised of a well characterized polyvinyl alcohol of 100 Å Rg and hard spherical LUDOX® silica of 80 Å radii. Interaction among the polymer chains forming polymer clusters with collective polymer structure factor induced by the polymer-mediated potentials of mean force between the nanoparticles, was observed. However, Gaussian nature of individual polymer chain remains unaltered. The dilute system of polymer with low silica volume fraction has the scattering form which was appropriately modeled as the sum of the individual profiles of spherical silica particles and polymer cluster of interchain packing. With increasing silica volume fraction in the dilute solution, the spatial range parameter between the particles is reduced; hence there is a net increase in the mean potential force and consequently to stronger interaction between the silica and polymer. In the dilute systems of high silica with low polymer volume fraction, the polymer chain apparently attracted closer to the silica and concurrently absorbed to the silica hard surface and their scattering data were excellently fit with a model form factor as comprising of one unit forming the core of the spherical silica particles and the interacting polymer as the corona. This result of severe change in polymer interchain conformation in a dilute system corroborated with reduced polymer viscosity observed.  相似文献   

4.
Stability and coalescence of emulsions stabilized with solid particles is determined by the energy of particle attachment at the liquid–liquid interface (the energy of adhesion) and by the value of capillary pressure arising in the emulsion film in the process of its thinning under the lower pressure when two layers of solid particles (on the opposite film sides) draw together up to their direct contact and formation of menicsi in the porous space between particles. We calculated maximal (critical) capillary pressure P c, max whose exceeding leads to the film rupture as a function of contact angle and the size of solid particles needed to form the adsorption layer of monodisperse spherical particles with a dense hexagonal packing. Capillary pressure isotherms P c(h) (h is the thickness of emulsion film) were also calculated. The deviation of meniscus shape from spherical was considered using the Mayer, Stowe, and Princen method. Determination of capillary pressure in a model emulsion film containing hexagonal-packed transparent glass spheres demonstrated that, at various degrees of particle hydrophobicity, experimental data are in good agreement with theoretical calculations of the P c, max value and P c(h) isotherm.  相似文献   

5.
This paper discusses the dynamics (i.e. the rates of the various charge displacement processes) of electrical double layers after disequilibration, taking the AgI-electrolyte solution interface as the model system. Relaxation studies performed with Ag/AgI electrodes lead to the conclusion that there is a relatively slow step, connected with the transfer of charge through the solid-solution interface. Evidence is given that this step occurs also with colloidal particles under interaction. It is inferred that in AgI sols interaction takes place at constant total particle charge, but at variable charge density distribution. The effect of particle radius a is also considered, an important parameter being the ratio τlint between the characteristic time scales of lateral charge flow and interaction. This ratio is dominated by the hydrodynamic drag. It appears that these dynamic factors level off the effect of the particle radius on the rate of coagulation. Other implications for colloid stability are discussed.  相似文献   

6.
Drag force on a metallic or nonmetallic spherical particle exposed to a plasma flow is studied for the extreme case of a free-molecule regime. Analytical expressions are derived for the drag components due to, respectively, atoms, ions, and electrons and for the total drag on the whole sphere due to all the gas species. It has been shown that the drag is proportional to the square of the particle radius or the drag coefficient is independent of the particle radius. At low gas temperatures with a negligible degree of ionization, the drag is caused mainly by atoms and could be predicted by using the well-known drag expression given in ordinary-temperature rarefied gas dynamics. On the other hand, the drag is caused mainly by ions at high plasma temperatures with a great degree of ionization. The contribution of electrons to the total drag is always negligible. Ignoring gas ionization at high plasma temperatures would overestimate the particle drag. There is a little difference between metallic and nonmetallic spheres in their total drag forces, with a slightly higher value for a metallic sphere at high plasma temperatures, but usually such a small difference could be neglected in engineering calculations. The drag increases rapidly with increasing gas pressure or oncoming speed ratio. For a two-temperature plasma, the drag increases at low electron temperatures but decreases at high electron temperatures with the increase in the electron/heavy-particle temperature ratio.Nomenclature C d Drag coefficient - e Elementary charge - f D,F D Local and total drag (N/m 2 andN) - f Velocity distribution function for incident gas particles - f + Velocity distribution function for reflected gas particles - k Boltzmann's constant - m Gas particle mass (kg) - n Number density of gas species (m –3) - P ,P + Surface pressure due to incident and reflected gas particles - R 0 Sphere radius (m) - S Speed ratio,S j=U/(2kT j/mj)1/2 - T e,T h Electron and heavy-particle (atom, ion) temperature - T w Wall temperature - U Oncoming plasma flow velocity - v x, vy, vz Velocity components of gas particles in thex, y, andz directions (m/sec) - v Thermal motion speed of gas particles,v j =(8kT j /m j )1/2 - v ze Smallestv z of electrons which could reach the sphere surface,v ze=(2e/m e)1/2 (m/sec) - v zw Value ofv z of ions or electrons as arriving at the sphere surface (m/sec) - Center angle - Gas density (kg/m3) - Shear stress (N/m2) - Absolute value of the floating potential (V) - , Local and total particle fluxes incident to the surface - a Atoms - e Electrons - h Heavy particles - i Ions - j jth gas species - m Metallic sphere - mn Nonmetallic sphere A preliminary version of this paper was presented at the Eighth International Symposium on Plasma Chemistry held in Tokyo, September 1987.  相似文献   

7.
Analytical expressions are presented for the drag force acting on an evaporating or nonevaporating particle immersed in a plasma flow for the extreme case of free-molecule flow regime and thin plasma .sheath. It is shown that the drag force on a spherical particle is proportional to the square of the particle radius and to the relative velocity between the particle and the bulk plasma at low speed ratios. The existence of a relative velocity between the particle and the plasma results in a nonuniform heat flux distribution with its rnaximum value at the frontal stagnation point of tire sphere. This nonuniform distribution of the local heat fux density causes a nonuniforrn distribution of the local evaporated-mass flux and vapor reaction force around the surface of an evaporating particle, and thus induces an additional force on the particle. Consequently, the drag force acting on art evaporating particle is always greater than that on a nonevaporating one. This additional drag force due to particle evaporation is more significant for nonmetallic particles and for particle materials with lower latent heat of evaporation and lower vapor molecular mass. It increases with increasing plasma temperature and with decreasing gas pressure at the high plasma temperatures associated with appreciable gas ionization. The drag ratio increases with increasing electron/heavy-particle temperature ratio at high electron temperatures for a two-temperature plasma.  相似文献   

8.

The hydrodynamic force (drag) on spherical and irregularly shaped particles significantly increases when the particles move close to solid and permeable boundaries. The overall effect of the increased hydrodynamic drag is to hinder the particle movement in the vicinity of boundaries and this includes the Brownian movement and electrophoresis. The Monte Carlo simulation method is used to model the Brownian movement, the resulting diffusion, and the electrophoresis of spherical particles in narrow, cylindrical pores, filled with Newtonian fluids. It is observed that the effect of the pore walls is a significant reduction of the space-averaged electrophoretic velocity of the particles, which implies reduced particle flux through the pores. The hindered electrophoresis is primarily a geometric phenomenon, caused by the increased drag and depends on the size of the particles and the pore-to-particle diameter ratio. The temperature of the fluid slightly affects the hindered electrophoresis through its effect on the viscosity, which is a determinant of the Brownian force, the diffusivity and the electrophoretic velocity. The hindered electrophoresis is almost independent of the other fluid and particle properties, such as density. Based on the simulation results a non-linear correlation for the flux of particles is derived, valid in the ranges 5?<?R/α?<?120, 5 nm?<?α?<?100 nm and 273 K?<?T?<?355 K.

  相似文献   

9.
Molecular spoked wheels with D3h and Cs symmetry are synthesized by Vollhardt trimerization of C2v-symmetric dumbbell structures with central acetylene units and subsequent intramolecular ring closure. Scanning tunneling microscopy of the D3h-symmetric species at the solid/liquid interface on graphite reveals triporous chiral honeycomb nanopatterns in which the alkoxy side chains dominate the packing over the carboxylic acid groups, which remain unpaired. In contrast, Cs-symmetric isomers partially allow for pairing of the carboxylic acids, which therefore act as a probe for the reduced alkoxy chain nanopattern stabilization. This observation also reflects the adsorbate substrate symmetry mismatch, which leads to an increase of nanopattern complexity and unexpected templating of alkoxy side chains along the graphite armchair directions. State-of-the-art GFN-FF calculations confirm the overall structure of this packing and attribute the unusual side-chain orientation to a steric constraint in a confined environment. These calculations go far beyond conventional simple space-filling models and are therefore particularly suitable for this special case of molecular packing.

Scanning tunneling microscopy investigations of phenylene-based molecular spoked wheels with D3h and Cs symmetries on graphite show the competitive or complementary effects of carboxylic acid groups and alkoxy chains on the nanopattern formation.  相似文献   

10.
Nano-LC columns of different lengths (14–35 cm), 75 μm i.d., were packed with solid-core C18 particles using a conventional HPLC system at low pressure (375 bar) and without expensive tools and fittings. Solid-core particles consist of a solid, non-porous core surrounded with a shell of a porous layer with a very narrow particle size distribution. This geometry allows a faster diffusion of the analytes compared to porous particles, reducing the C term of the Knox plot. Different slurries of packing material were evaluated and tested. The packing procedure was carried out at room temperature and at 70 °C to evaluate the influence of this parameter on the overall process. The synergic action of pressure, temperature and sonication contributed to columns of various lengths in the packing process. The columns were tested at room temperature taking into account the following parameters: Knox plots, specific permeability and peak capacity. Reduced heights of theoretical plates, h, ranged between 3.8 and 5.1 at ν between 2 and 6. An LC-MS test was carried out with a Direct-EI LC-MS instrument.  相似文献   

11.
Particle removal by electro-osmotic flow was investigated by comparison with the removal by ordinary flow of water without electrokmetic effect. The relationship between adhesion and removal of particles in terms of force acting on the particle was also discussed. Experiments were carried out in an aqueous solution using nylon particles and a quartz plate. The adhesive force,F T, for the particles which adhered to the quartz plate in secondary minima in the total potential energy of interaction versus separation distance curves was calculated. Particle removal experiments were carried out applying electro-osmotic and Poiseuille flows. The hydrodynamic force,F d, which was required to remove particles from the plate was estimated using flow velocities. The effectiveness of electro-osmotic flow on particle removal was larger than that of Poiseuille flow. In the particle removal by electro-osmotic flow, the minimum of the ratioF d/F t for particle removal was found to be 50 and the ratio for removal efficiency of 0.5 was about 140.  相似文献   

12.
Abstract

With optimised packing procedures, spherical shaped silica gel particles produce 1.5 to 2 times more plates in HPLC than irregular shaped silica gel particles. The lowest reduced plate height obtained by us so far is for 5 μm ROSiL-C18-HL-D and is h: 1.62 for k': 4.5. It is suggested to transform h into 100/h% and to name this the “Chromatographic efficiency”, or a % of the ideal 100% limit. This limit would be an h value equal to the mean particle diameter. Spherical and irregular silica gel particles of 5 and 10 μm particle diameter and with similar physical characteristics have the same permeability in HPLC columns.

Whether a correct column packing procedure is used can be shown by the constancy of plate number and column permeability in function of different packing pressures.  相似文献   

13.
In this paper the conformational properties and dynamics of protein-like lattice chains confined in an infinite cylinder are investigated by using Monte Carlo simulation method and the hybrid (3 1 0) lattice model. When the diameter of infinite cylinder D is greater than 0.5-0.7RF, the shape of protein-like chains remains unchanged, here Flory radius RF is RF = 3.0N0.6. When D < 0.5RF, the confinement of the cylinder help to form the helices, however, it hinders to form the coils, and the shape of protein-like chains changes from sphere-like structure to rod-like one. Different confinements affect distinctly to form the secondary structure of proteins. There exist a different dynamics behavior between protein-like chains confined in an infinite cylinder and general protein-like chains without confinement. These investigations may provide some insights into the influence of crowding condition to the conformation properties and dynamics of proteins.  相似文献   

14.
《Chemical physics letters》1986,124(6):557-560
The onset of light absorption of various Q-CdS samples (materials of extremely small particle size showing size quantization effects) was determined to obtain the dependence of Λthreshold as a function of particle radius R. We describe the different experimental methods used for determining the size of the CdS particles. For the quantum-mechanical theory, which interprets the effect in terms of the energy levels of a spatially confined exciton, some new model calculations are reported. The previous model can be improved by introducing a finite depth for the spherical potential box representing the crystallite.  相似文献   

15.
The Stokes–Brinkman flow field has been calculated in a model deposit, i.e., a row of parallel chains formed from porous spherical clusters of nanoparticles and oriented perpendicularly to the gas flow direction. The force of drag to an air flow has been calculated for the row of chains taking into account their permeability and the distance to neighboring chains. The drag forces have been found for nanodendrites composing clusters with allowance for the gas slip effect. Corresponding approximation formulas have been derived. A method has been proposed for calculating the pressure drop across a highly porous deposit of clusters of aerosol nanoparticles deposited onto a filter.  相似文献   

16.
The effects of particle size on polyacrylamide (PAAm,M w =59×104, 500×104) adsorption were investigated using a series of well-characterized hematite (-Fe2O3) dispersions. The -Fe2O3 particles with highly monodisperse and nearly spherical shape ranged in radius from 23 nm to 300 nm. the maximum amount of PAAm adsorption (M m ) in each system, showed a steady increase with decreasing particle radius and was influenced strongly by particle concentrations in the medium. Furthermore, it was realized that the diameter of -Fe2O3 particles after treatment with PAAm under different particle concentrations decreased with increasing particle concentration. The relation between particle concentration in the medium and particle size after treatment was also influenced by the medium pH, i.e., at the medium pH close to the isoelectric point of -Fe2O3 particles (pHo=9.2), the particle size after treatment increased with increasing particle concentration. All these results suggest that in the system of ultra-fine particles, the mixing process between particle-particle and polymerparticle will play an important role on the conformation of adsorbed polymer layer.  相似文献   

17.
18.
The derivation of ionic Stokes radii, r iSt, from ion mobilities (conductivities ?? i ?? ) and their relation to the actual ionic radii, of the bare or the hydrated ions, is followed through the years. It is concluded that such quantities as r iSt do not convey any useful information beyond what is already available in the ?? i ?? , and in particular that the r iSt should not be used for the estimation of solvation numbers. The concept of a Stokes radius of ions that are smaller than or commensurate with molecules of the solvent should be scrapped altogether.  相似文献   

19.
Particles in nonthermal dusty plasmas tend to charge negatively. However several effects can result in a significant fraction of the particles being neutral or positively charged, in which case they can deposit on surfaces that bound the plasma. Monte Carlo charging simulations were conducted to explore the effects of several parameters on the non-negative particle fraction of the stationary particle charge distribution. These simulations accounted for two effects not considered by the orbital motion limited theory of particle charging: single-particle charge limits, which were implemented by calculating electron tunneling currents from particles; and the increase in ion current to particles caused by charge-exchange collisions that occur within a particle’s capture radius. The effects of several parameters were considered, including particle size, in the range 1–10 nm; pressure, ranging from 0.1 to 10 Torr; electron temperature, from 1 to 5 eV; positive ion temperature, from 300 to 700 K; plasma electronegativity, characterized in terms of n +/n e ranging from 1 to 1000; and particle material, either SiO2 or Si. Within this parameter space, higher non-negative particle fractions are associated with smaller particle size, higher pressure, lower electron temperature, lower positive ion temperature, and higher electronegativity. Additionally, materials with lower electron affinities, such as SiO2, have higher non-negative particle fractions than materials with lower electron affinities, such as Si.  相似文献   

20.
The viscoelastic properties of concentrated microlatex dispersions were investigated using oscillatory measurements. The latices were prepared by polymerisation of styrene-in-water microemulsions using UV and azobiisobutyronitrile initiator. The complex modulus, G*, storage modulus, G′ and loss modulus, G″ were measured as a function of strain amplitude (to obtain the linear viscoelastic region) and frequency at various latex volume fractions. Two latices with radii of 3.9 and 15.1 nm were investigated at 20°C. The results showed a change from predominantly viscous to a predominantly elastic response at a critical volume fraction, φc With the smaller latex, the concentration of the free surfactant in bulk solution was relatively low (2.6%) and the dispersions remained stable. φc was found to be 0.161. Assuming random packing of the particles (volume fraction=0.64), an estimate was obtained for the adsorbed layer thickness and this was found to be 1.4 nm, which is small for a surfactant chain with 15 ethylene oxide units. However, since the surfactant layer is a mixture of chains with 4 and 15 EO units, it is likely that the larger PEO chains will undergo interpenetration and/or compression on close approach of the particles. With the larger latex, on the other hand, there was high free surfactant concentration (9.1%) and this led to depletion flocculation. This results in a lower φc than would be the case in the absence of flocculation.  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