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 15 毫秒
1.
Abstract

Most current electrostatic surface complexation models describing ionic binding at the particle/water interface rely on the use of Poisson–Boltzmann (PB) theory for relating diffuse layer charge densities to diffuse layer electrostatic potentials. PB theory is known to contain a number of implicit assumptions whose significance in environmental applications is largely unknown. This study seeks to better quantify the impact of these assumptions by: (1) comparing potentials obtained from planar analytical solutions to the PB with those obtained from Hypernetted Chain (HNC) theory (Attard, 2006), (2) assessing the accuracy of the Ohshima et al. (1982) spherical approximate analytical solution to the PB equation by comparison with published numerical values (Loeb et al., 1961), and (3) comparing interfacial potential estimates obtained from the spherical approximate analytical solution to the PB equation at and adjacent to the particle surface with potential estimates obtained from the Entropic Balanced Surface Potential (EB) model (Loux, 1985; Loux and Anderson, 2001) and published potential estimates obtained from the Hypernetted Chain/Mean Spherical Approximation procedure (HNC/MSA; Gonzalez-Tovar and Lozada-Cassou, 1989). EB potential estimates were obtained assuming a surface volume thickness equal to the Bjerrum length (0.357 nm in a room temperature monovalent electrolyte solution). Findings from the study included: (1) the planar, surficial HNC estimates compared favourably with planar surficial PB relationships at charge densities equal to or less than 0.05 C m?2, (2) the Ohshima et al. (1982) approximate spherical analytical solution to the PB equation replicated the numerical charge density estimates required to obtain 72 datapoints over an e<img>/kT range of one to four with a maximum error of 3.37% and a coefficient of variation of 0.92%, (3) for a 0.1 μm radius particle in a room temperature 0.01 M (1 : 1) ionic strength solution, potential estimates over a surface charge density range of 0 to 0.3C m?2 occurred in the following order: ψHNC/MSA,RPB,R >ψHNC/MSA,R+0.2125nmPB,R+0.2nm ~ ψEBHNC/MSA,R+0.425nm ~ ψPB,R+0.4nm and (4) with 45 datapoints including both 1 μm and 10 nm radius particles over an ionic strength range of 1.0 to 0.001 M, the PB potential estimates 0.2 nm from the particle surface (ψPBR+02nm) closely tracked the corresponding EB estimates (ψEB) with a 5.3% coefficient of variation. If one assumes that interfacial potential values adjacent to the particle surface are most relevant for describing environmental phenomena and that a 10% coefficient of variation in potential estimates is acceptable, then presumably any of the non-surficial charge/potential relationships would be useful below an absolute charge density of 0.125 C m ?2 (with monovalent electrolyte solutions).  相似文献   

2.
Abstract

Monte Carlo simulations using a Markov process corresponding to a (generalized) Grand Canonical Ensemble have been performed for a number of spherical micropores in equilibrium with dilute external bulk solutions of primitive model electrolytes. Dilute solutions of 1:1 electrolytes with a Bjerrum parameter B = 1.546 with cations three times larger than the anions have been simulated. Also, dilute solutions of 2:1 electrolytes with ions of equal size and reduced Bjerrum parameters Br = 1.546 and 3 have been simulated. The pores are primitive pores with hard walls and the same dielectric permittivity in the wall and in the pore solution. They range from a pore radius = 5 times the mean ionic diameter to 35 times this diameter, and they carry a fixed charge equal to + 5,0 and ?5 elementary charges. The fixed charge is modelled as smoothly distributed on the pore-wall interface. In addition to the electric potential of the interfacial charge and the electric potential of the spherical double layer, a potential Δ between the pore solution and the bulk solution may be deliberately added. For single pores we may take Δ = 0, but then the pore is generally not electroneutral. In a “Swiss cheese” membrane with a lot of (equally sized) pores, the membrane phase has to approach electroneutrality for growing size of the phase. This is approximated by means of a membrane generated potential Δ in each pore (from the electrostatic interactions with the other pores). The potential A so chosen to obtain electroneutrality is the GCEMC Donnan potential. These non-ideal Donnan potentials are compared to the ideal values (with activity coefficients equal to zero). From the mean occupation numbers of cations and anions in the pores, the average pore values of the mean ionic and the single ionic activity coefficients of the ions are calculated. These are very dependent on pore sizes and on the potential in the pore. The excess energy and the electrostatic Helmholtz free energy of the ions in the pores are also simulated directly. The electrostatic entropy is found as the difference.  相似文献   

3.
Globular protein gels with a variety of structures were prepared by heating β-lactoglobulin solutions at different concentrations and different ionic strengths. The structure was analyzed in terms of the pair correlation function of the protein concentration, and the volume fraction of the gels was determined. A strong coarsening of the gel structure was observed upon increasing the NaCl concentration between 0.1 and 0.25 M. The mean square displacement of spherical particles with diameters between 0.2 and 2 μm was determined in solutions and in gels by multiparticle tracking of confocal laser scanning microscopy images. Brownian diffusion or trapping of spheres with different sizes was observed, depending on the gel structure. In few cases the diffusion was anomalous. The relationship between gel structure and particle mobility is discussed.  相似文献   

4.
Abstract

The retention of Cd by an organic soil was investigated as a function of pH and ionic strength. The adsorption of Cd at pH values from 2 to 11 at two ionic strengths (0.053 M and 0.017 M LiNO3) were found to be a function of both pH and ionic strength. Four Cd-humic complexation models were evaluated in order to test the applicability of these models to fit data from batch adsorption experiments. The models varied greatly in their complexity and implicit assumptions. Three were discrete functional group models – a simple diprotic acid model, a two diprotic acid model and the Windermere Humic Aqueous Acid (WHAM) model, and a continuous functional group model - the non-ideal competitive adsorption (NICA) model. The concentration of proton binding sites in the soil was found to be 4.51 mol kg-1. The NICA and WHAM models were more successful than either a simple diprotic acid model or a two diprotic acid model at modeling Cd complexation by the organic soil, although both underestimated adsorption at very high pH values.  相似文献   

5.
To understand the effects of ionic strength and pH on the transport of natural soil nanoparticles (NS) in saturated porous media, aeolian sandy soil nanoparticles (AS), cultivated loessial soil nano particles (CS), manural loessial soil nanoparticles (MS) and red soil nanoparticles (RS) were leached with solutions of varying pH and ionic strength. The recovery rate of soil nanoparticles decreased in the order AS > RS > MS > CS. Transport of soil nanoparticles was enhanced with increasing pH and decreasing ionic strength and was attributable to changes in the Zeta potential of NS. Deposition of NS was also affected by the composition of soil nanoparticles and the surface charge. Column experiments showed that the interaction between soil nanoparticles and saturated quartz sand was mainly due to the physical and chemical properties of soil nanoparticles. The Derjaguin–Landau–Verwey–Overbeek interaction energies between NS and sand were affected by pHs and ionic strengths. Soil nanoparticles transport through saturated porous media could be accurately simulated by the one-dimensional advection-dispersion-reaction equation.  相似文献   

6.
Generation of high-voltage high-current electron beams in a low-pressure (P = 0.1–1 Torr) gas discharge is studied experimentally as a function of the discharge voltage and the sort and pressure of the plasma-forming gas. The density of the plasma formed by a high-current electron beam is measured. Experiments on ultrahigh charging of targets exposed to a pulsed electron beam with an energy of up to 25 keV, an electron current density of higher than 1 A/cm2, a pulse duration of up to 1 μs, and a repetition rate of up to 1 kHz are described. A numerical model of ultrahigh charging of dust grains exposed to a high-energy electron beam is developed. The formation of high-energy positive ions in the field of negatively charged plane and spherical targets is calculated. The calculations performed for a pulse-periodic mode demonstrate the possibility of achieving neutron yields of higher than 106 s–1 cm–2 in the case of a plane target and about 109 s–1 in the case of 103 spherical targets, each with a radius of 250 μm.  相似文献   

7.
The planktonic bacterial populations in the surface water of the hypertrophic Himon-ya Pond were separated into four fractions (>35 μm, 35-5 μm, 5–1 μm, and <1 μm) by size fractionation of suspended particles in the water. The seasonal variations in bacterial numbers over a two year period differed for each of the four fractions. The bacterial counts in the >35 μm fraction were mainly dependent on the biomass of Microcystis colonies. Their peaks were observed in summer. In the 35-5 μm and 5–1 μm fractions, several peaks in bacterial numbers were observed, influenced by the quality and quantity of the particles associated with the bacteria. The bacterial counts in the <1 μm fraction contributed a high proportion of the total throughout all seasons except summer.  相似文献   

8.
Citrate and other polyanion binding to ferricytochrome c partially blocks reduction by ascorbate, but at constant ionic strength the citrate-cytochrome c complex remains reducible; reduction by TMPD is unaffected. At a constant high ionic strength citrate inhibits the cytochrome c oxidase reaction competitively with respect to cytochrome c, indicating that ferrocytochrome c also binds citrate, and that the citrate-ferrocytochrome c complex is rejected by the binding site at high ionic strength. At lower ionic strengths, citrate and other polyanions change the kinetic pattern of ferrocytochrome c oxidation from first-order towards zero-order, indicating preferential binding of the ferric species, followed by its exclusion from the binding site. The turnover at low cytochrome c concentrations is diminished by citrate but not the Km (apparent non-competitive inhibition) or the rate of cytochrome a reduction by bound cytochrome c. Small effects of anions are seen in direct measurements of binding to the primary site on the enzyme, and larger effects upon secondary site binding. It is concluded that anion-cytochrome c complexes may be catalytically competent but that the redox potentials and/or intramolecular behaviour of such complexes may be affected when enzyme-bound. Increasing ionic strength diminishes cytochrome c binding not only by decreasing the 'association' rate but also by increasing the 'dissociation' rate for bound cytochrome c converting the 'primary' (T) site at high salt concentrations into a site similar kinetically to the 'secondary' (L) site at low ionic strength. A finite Km of 170 microM at very high ionic strength indicates a ratio of K infinity m/K 0 M of about 5000. It is proposed that anions either modify the E10 of cytochrome C bound at the primary (T) site of that they perturb an equilibrium between two forms of bound c in favour of a less active form.  相似文献   

9.
Histone H 1-depleted chromatin made from acid extracted, intact nuclei was exposed to various ionic strengths. NaCl concentrations above 0.3 M sufficed to generate novel oligonucleosomes formerly characterized as "compact oligomers" and "spacerless dinucleosomes". Such particles could not be identified within H 1-depleted nuclei or chromatin at low ionic strengths. Their formation, proceeding within days at 0 degrees C, was accelerated by increasing ionic strengths. The data was discussed in terms of a salt-induced motion of nucleosomal core particles along the DNA to form compact oligomers.  相似文献   

10.
Abstract

Rare earth distribution coefficients, DT = (moles cm-3, cells)/(moles cm-3, solution), obtained using seawater (S = 36.4, t = 25°C, pH ~ 8.2, pCO2 ~ 345 μatm) and the marine diatom, Skeletonema costatum, exhibited a strong tendency toward the order Ce > Gd > Yb. Observations of rapid initial uptake, with subsequent gradual uptake over time, are suggestive of initial adsorption onto cell surfaces followed by slow transport to interior cell sites. The average volume concentration factors (DT) obtained in our study are: DTCe = (3.33 ± 0.9) × 105; DTGd = (2.41 ± 0.7) × 105; DTYb = (1.64 ± 0.3) × 105. Distribution coefficient results, expressed as a competition between solution and solid-state complexation terms, indicate that rare earth element complexation, both in solution and on surfaces, strongly increases with atomic number. Relatively small differences in rare earth element distribution coefficients (DT) with atomic number are the result of small differences between large solution and solid-state complexation terms.  相似文献   

11.
ABSTRACT

Metal—humate interactions in aqueous and terrestrial ecosystems control the speciation of trace metals and hence, their bioavailability and toxicity. The present study investigated the complexation interactions of copper, cadmium and lead, in their divalent states, with humic acids extracted from three different sources: the treated sewage from a treatment plant, Yamuna river bed sediment, and Yamuna river flood plain soil, all at Okhla, Delhi, employing ion selective electrode potentiometry. The conditional stability constants of the complexes were computed from Scatchard plots. The influences of the two important metal speciation factors, viz., pH and ionic strength of the reaction medium on the conditional stability constants were ascertained by investigating the reactions under three different pH: 5.0, 6.0 and 7.0 and three different ionic strengths: 0.01, 0.03 and 0.1. Stabilities of metal—humic acid (HA) complexes follow the order: Cu-HA > Pb-HA > Cd-HA for humic acids from any single source and are found to increase with a rise in pH and fall in a ionic strength of the medium. The humic acids extracted from the soil and the sediment emerged as stronger complexing agents, as compared to that extracted from sewage.  相似文献   

12.

Steady laminar axisymmetric inhalation flow and wall deposition of micron-size particles in representative triple bifurcation airways have been simulated using a commercial finite-volume code with user-enhanced programs. Assuming spherical non-interacting particles (3 μm≤ d p ≤7 μm), various inlet Reynolds numbers (Re=500-2000) and Stokes numbers (St=0.02-0.23) were considered. The resulting particle deposition patterns were analyzed and then summarized in terms of deposition efficiencies, i.e. DE=DE(Re,St) Surprisingly high DE-values occur at relatively low Reynolds numbers (e.g., Re=500 ) in the third bifurcation. The quantitative results are of interest to researchers either conducting health risk assessment studies for inhaled particulate pollutants or analyzing drug aerosol inhalation and deposition at desired lung target sites.  相似文献   

13.
Abstract The morphological characteristics and development of Oedaleus asiaticus entomopoxvirus propagated in Oeddeus infernalis are reported. This virus mainly infected host's fat bodies and hemocytes. Three kinds of spheroids were observed during propagation: big spheroid (30. 41 μm × 25. 40 μm), ellipsoid (6. 58 μm × 4. 78 μm) and small spheroid (3. 35 μm × 2. 60 μm). The virions embeded in them were oval, measuring 230 nm × 176 nm. The typical characteristic of poxviruses as spherical units with the mulberry-like surface could be seen under transmission electron microscope. The lateral body was cylinder-shaped. The rope-like substances in the core folded back only once; therefore two spots in transverse sections were seen. The development of the virions included four stages: the appearance of viro-plasm, the formation of spherical particles, the differentiation of core and capsid. The grasshoppers only in the same genus could be infected by this virus.  相似文献   

14.
Experimental data on the mean activity coefficient of CaCl2 in NaCl-CaCl2 mixtures at ionic strengths below 1 m have been used to prepare a table of activity coefficients for Ca++ in solutions of physiological interest. The establishment of an empirical calcium ion activity scale is discussed, and a number of possible assumptions are examined. The assumption γ++ = (γ±)2 is suggested as being the simplest with a theoretical basis.  相似文献   

15.
Helix pomatia beta-haemocyanin was split into dissociation products by varying the pH and the ionic strength. The purity of the solution was checked in an ultracentrifuge. Two defined dissociation products were studied in solution by small-angle X-ray scattering. In Tris-HC1 buffer, pH 8.0 and ionic strength 1 M, the following parameters of the dissociation product (tenths) could be determined: molecular weight = 7 x 10(5), volume = 1350 nm3, radius of gyration = 9.0 nm, maximal distance = 28.3 nm, radius of the spherical subunits about 2.6 nm, number of the subunits approximately 19. Tris-HC1 buffer, pH 8.7 and ionic strength 0.01 M, yielded dissociation products (twentieths) with the following parameters: molecular weight = 3.5 x 10(5), volume = 635 nm3, radius of gyration = 7.5 nm, maximal distance = 21.9 nm, radius of the spherical subunits about 2.5 nm. With this information, the assumption that the larger fragment was double the smaller one and the latest biochemical and morphological results, theoretical scattering curves of models were calculated and compared with the experimental curves. Two models are suggested which argee well with the dissociation products in radius of gyration and scattering.  相似文献   

16.
The acid-base properties of Adenosine 5'-triphosphate (ATP) in NaCl and KCl aqueous solutions at different ionic strengths (0相似文献   

17.
18.
In the presence of ATP and absence of Ca2+, muscle crossbridges have either MgATP or MgADP.Pi bound at the active site (S. B. Marston and R. T. Tregear, Nature [Lond.], 235:22:1972). The behavior of these myosin adenosine triphosphate (M.ATP) crossbridges, both in relaxed skinned rabbit psoas and frog semitendinosus fibers, was analyzed. At very low ionic strength, T = 5 degrees C, mu = 20 mM, these crossbridges spend a large fraction of the time attached to actin. In rabbit, the attachment rate constants at low salt are 10(4) - 10(5) s-1, and the detachment rate constants are approximately 10(4) s-1. When ionic strength is increased up to physiological values by addition of 140 mM potassium propionate, the major effect is a weakening of the crossbridge binding constant approximately 30-40-fold. This effect occurs because of a large decrease, approximately 100-fold, in the crossbridge attachment rate constants. The detachment rate constants decrease only 2-3-fold. The effect of ionic strength on crossbridge binding in the fiber is very similar to the effect of ionic strength on the binding of myosin subfragment-1 to unregulated actin in solution. Thus, the effect of increasing ionic strength in fibers appears to be a direct effect on crossbridge binding rather than an effect on troponin-tropomyosin. The finding that crossbridges with ATP bound at the active site can and do attach to actin over a wide range of ionic strengths strongly suggests that troponin-tropomyosin keeps a muscle relaxed by blocking a step subsequent to crossbridge attachment. Thus, rather than troponin-tropomyosin serving to keep a muscle relaxed by inhibiting attachment, it seems quite possible that the main way in which troponin-tropomyosin regulates muscle activity is by preventing the weakly-binding relaxed crossbridges from going on through the crossbridge cycle into more strongly-binding states.  相似文献   

19.
Calf lens alpha-crystallins are polydisperse globular particles made of a large number of two types of subunits, A and B, both of molecular weight congruent to 20,000. alpha-Crystallin populations consisting on average of 40 subunits or more were subjected to various changes in pH, ionic strength, temperature and urea concentration. Modifications in quaternary structure induced by variation of these physicochemical parameters were followed by means of X-ray and quasi-elastic light-scattering and quantified in terms of weight average molecular weight (M), radius of gyration (Rg) and hydrodynamic radius (Rh). High-pressure liquid chromatography was used as a control of polydispersity. Increasing the pH, decreasing the ionic strength and incubating at temperatures from 20 degrees C to 45 degrees C all resulted in the formation of particles of decreasing M, Rg and Rh values. These effects are cumulative. All monomodal alpha-crystallin populations encountered in this study, which covers a wide range of sizes and molecular weights, may be accounted for by a three-layer model with partial filling up of the layers. Applying basic principles of symmetry and postulating specific contacts between protein subunits to construct this three-layer model leads to tetrahedral symmetry, with 12, 24 and 24 sites in the first, second and third layers, respectively. Variations in probabilities of site occupancy account for both the observed quaternary structure modifications and the intrinsic polydispersity of alpha-crystallins  相似文献   

20.
The thermal polymerization of orosomucoid   总被引:1,自引:1,他引:0       下载免费PDF全文
1. Orosomucoid was prepared from the urine of a nephrotic patient and polymerized by heating it in a range of salt concentrations at pH4·1. 2. Heating at low ionic strengths produced a `chain' polymer of indefinite length but having the same width as the diameter of the monomer (5·0nm.). Similar treatment in high ionic strengths also produced a spherical (`ball') polymer of limited diameter (14·8nm.). 3. The size and shape of both polymers were determined from ultra-centrifuge, gel-filtration and electron-microscope results. The results suggest that eight monomer units condense to form the ball polymer. 4. Heating orosomucoid at pH1·8 hydrolysed the N-acetylneuraminic acid off the molecule; only chains could then be formed, even in high ionic strengths. 5. Both polymers were stable under normal conditions but could be depolymerized in 3m-guanidine hydrochloride. The monomer could be repolymerized on heating: the `chain monomer' only formed chains at all ionic strengths, but the `ball monomer' was indistinguishable from the original monomer in its immunological properties and polymerization reaction.  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