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 78 毫秒
1.
Sphingomyelinase of Bacillus cereus proved to be specifically adsorbed onto mammalian erythrocyte membranes in the presence of either Ca2+ or Ca2+ plus Mg2+ in the order of sphingomyelin content; i.e., sheep, bovine greater than porcine greater than rat erythrocytes. No appreciable adsorption was observed in the presence of Mg2+ alone nor in the absence of divalent metal ions. The enzyme adsorption onto bovine erythrocytes was dependent upon the incubation temperature. By shifting the temperature from 37 to 0 degrees C, sphingomyelinase once adsorbed onto the surface of bovine erythrocytes was released into the supernatant. Ca2+ proved to be an essential factor for the enzyme adsorption: The addition of 1 mM Ca2+ enhanced the adsorptive process, but inhibited sphingomyelin hydrolysis and hot or hot-cold hemolysis of erythrocytes, while the addition of 1 mM Ca2+ plus 1 mM Mg2+ enhanced sphingomyelin breakdown and hemolysis as well as the enzyme adsorption. However, when the amount of sphingomyelin fell off to 0.2-0.7 nmol/ml or less by the action of sphingomyelinase, the enzyme once adsorbed was completely released from the surface of erythrocytes. The result indicates that the major binding site for sphingomyelinase is sphingomyelin. In the presence of 1 mM Mg2+ alone, the enzymatic hydrolysis of sphingomyelin and hemolysis proceeded whereas the enzyme adsorption was not encountered during 60 min incubation at 37 degrees C. The change in the molar ratio of Ca2+ to Mg2+ affected the enzyme adsorption and sphingomyelin breakdown; the higher Ca2+ enhanced the adsorption whereas the higher Mg2+ stimulated sphingomyelin hydrolysis.  相似文献   

2.
The structure and surface activity of the extracellular fraction of pulmonary surfactant known as tubular myelin are Ca2+ dependent. Previous studies have demonstrated surfactant-specific proteins with monomeric molecular weights of 28,000-36,000 (SP28-36) are associated with this fraction. In reassembled lipoprotein mixtures, SP28-36 promotes the Ca2+-induced aggregation and surface activity of surfactant lipids, but the detailed interactions between Ca2+, SP28-36, and surfactant lipids have not been established. In this study, we investigated the effect of various cations on the aggregation of surfactant lipid liposomes in the presence of SP28-36. SP28-36 reduced the threshold ion concentration for liposome aggregation from greater than 10 to 0.5 mM for Ca2+, Ba2+, and Sr2+ but not Mg2+ or Mn2+. The liposome aggregation was reversed by ethylenediaminetetraacetic acid and not associated with leakage of carboxyfluorescein. SP28-36 promoted similar liposome aggregation at pH less than 5 in the absence of divalent cations. Surfactant lipids adsorbed slowly to an air-fluid interface in all ionic conditions unless SP28-36 was present. Both Ca2+ and H+ induced rapid lipid adsorption in the presence of SP28-36. The surface activity of native surfactant had a similar ion dependence. Electron micrographs of native surfactant showed typical tubular myelin structures at pH 7.4 only in the presence of Ca2+. At pH 4.4 in the absence of Ca2+, similar but not identical structures were seen. In the reconstituted system, SP28-36 in the presence of Ca2+ induced the formation of larger multilayered structures including parallel bilayers and small areas of squares and triangles with dimensions similar to structures found in the native material.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
An analysis of Ca2+ influence on the thermotropic behaviour of different phosphatidylcholine-vitamin D3 mixtures was carried out by differential scanning calorimetry technique (DSC). A competitive effect between Ca2+ and vitamin D3 on the membrane fluidity was detected. The observed shifts of the gel-liquid crystal transition temperature were correlated with the mole fraction of vitamin D3 dissolved within the lipid bilayer as well as with the Ca2+ concentration in the surrounding medium. These shifts were rationalized on the ground of a simple microscopic model through the calculation of the internal pressure exerted by the adsorbed Ca2+ on the lipid matrix by the Clapeyron equation. The experimental results and the obtained equations accord with each other and support the idea of micro-domain formation richer in one lipid component. The extent of such lateral phase separation of the lipid components seems to be favoured by the adsorption of Ca2+ at the membrane-water interface.  相似文献   

4.
The activity of phospholipase C from Clostridium perfringens on 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) as a monolayer at an air/water interface was examined. With a pure POPC monolayer, sharp cut-off of the enzyme activity was observed on increase in surface pressure. However, this cut-off disappeared on addition of a 0.3 molar fraction of 1,2-dioleoylglycerol (1,2-DO) to the monolayer. An abrupt change in the enzyme activity was observed with molar fractions of between 0.2 and 0.3 1,2-DO in the POPC monolayer at an initial surface pressure of 35 mN/m. For examination of the effect of 1,2-DO on the phospholipase C activity, the quantity of [125I]phospholipase C adsorbed to the surface was determined. The enzyme was found to be adsorbed nonspecifically to all lipid films except that of POPC only. The adsorption of enzyme was not affected by the presence or absence of Ca2+ and Zn2+. The rate constant for enzyme adsorption to a 1,2-DO film was 4.5 times that for its adsorption to a POPC film. The adsorption decreased linearly with increase in the surface concentration of POPC, and increased with increase in the surface concentration of 1,2-DO. These data suggest that 1,2-DO (a reaction product) regulates the interaction of phospholipase C with films containing substrate and may also regulate the enzyme activity.  相似文献   

5.
The isotherm of ferritin adsorption onto a hydrophobic surface was studied by transmission electron microscopy. Adsorbed ferritin was found to be distributed in molecular clusters. The adsorption process was diffusion-rate-limited after 20 h adsorption time at bulk concentrations below 1 mg/1. The clusters formed during the diffusion-rate-limited adsorption had a fractal dimension D approximately 1.0 when averaged over all clusters. The pair distribution function g(r) showed an increased probability of finding nearest neighbours at distances less than 30 nm. The surface concentration of adsorbed ferritin was weakly dependent on the bulk concentration of ferritin in the range 10 mg/1-10 g/1 and the average number of nearest neighbour molecules was constant in this concentration range. The mass distribution of adsorbed ferritin c(r) had a fractal dimension D = 1.8 at a bulk concentration of 10 g/l and a surface concentration corresponding to theta = 0.45 +/- 0.05. The pair correlation function g(r) showed decreasing probability of finding nearest neighbour molecules over long distances as in percolating clusters. The results indicate that ferritin adsorbs strongly to the surface at low surface concentrations and weakly at high surface concentrations. The stability of ferritin adsorption was correlated to the average number of nearest neighbour molecules, indicating a possibility that desorption is a critical supramolecular phenomenon.  相似文献   

6.
This present article describes a new and simple method for preparing model lipid bilayers. Stable and reproducible surface layers were produced at silica surfaces by co- adsorbing lipid with surfactant at the silica surface from mixed micellar solutions. The adsorption was followed in situ by use of ellipsometry. The mixed micellar solution consisted of a lipid (L-alpha-dioleoyllecithin) and a non-ionic sugar-based surfactant (n-dodecyl-beta-maltoside). The latter showed, by itself, no affinity for the surface and could, therefore, easily be rinsed off the surface after the adsorption step. By first adsorbing from solutions with high lipid and surfactant concentrations and then, in succession, rinsing and re-adsorbing from solutions with lower lipid-surfactant concentrations, a dense-packed lipid bilayer was produced at the silica surface. The same result can be achieved in a one-step process where the rinsing, after adsorption from the concentrated solution, is performed very slowly. The thickness of the adsorbed lecithin bilayer after this treatment found was to be about 44 +/- 3 A, having a mean refractive index of 1.480 +/- 0.004. The calculated surface excess of lipids on silica was about 4.2 mg m(-2), giving an average area per lipid molecule in the two layers of 62 +/- 3 A2. The physical characteristic of the adsorbed bilayer is in good agreement with previously reported data on bulk and surface supported lipid bilayers. However, in contrast to previous investigations, we found no support for the presence of a thicker multi-molecular water layer located between the lipid layer and the solid substrate.  相似文献   

7.
A quartz crystal microbalance with dissipation (QCM-D) is used to determine the adsorption rate of a supercoiled plasmid DNA onto a quartz surface and the structure of the resulting adsorbed DNA layer. To better understand the DNA adsorption mechanisms and the adsorbed layer physicochemical properties, the QCM-D data are complemented by dynamic light scattering measurements of diffusion coefficients of the DNA molecules as a function of solution ionic composition. The data from simultaneous monitoring of variations in frequency and dissipation energy with the QCM-D suggest that the adsorbed DNA layer is more rigid in the presence of divalent (calcium) cations compared to monovalent (sodium) cations. Adsorption rates are significantly higher in the presence of calcium, attaining a transport-limited rate at about 1 mM Ca2+. Results further suggest that in low ionic strength solutions containing 1 mM Ca2+ and in moderately high ionic strength solutions containing 300 mM NaCl, plasmid DNA adsorption to negatively charged mineral surfaces is irreversible.  相似文献   

8.
The Ca2+-ATPase of sarcoplasmic reticulum from rabbit skeletal muscle was incorporated into vesicles made from dimyristoylphosphatidylcholine or dipalmitoylphosphatidylcholine. The Ca2+-ATPase activity of these reconstituted membranes became appreciable above 20 degrees C and 30 degrees C, respectively, in accord with the results of previous investigators. Measurement by the spin-labeling technique of the fluidity of the bulk lipid revealed the gel-to-liquid crystalline phase transition at 29 degrees C and 39 degrees C, respectively, while the fluidity of the boundary lipid in both samples was found to be low throughout the temperature range studied. The rotational mobility of the Ca2+-ATPase protein in both samples, measured by saturation transfer electron spin resonance, was also very low throughout the temperature range studied and its temperature-dependence did not show any break or jump corresponding to the phase transition of the bulk lipid. On the other hand, the structural fluctuation of the Ca2+-ATPase protein in dimyristoylphosphatidylcholine-recombinant, measured in terms of hydrogen-deuterium exchange reaction kinetics, showed a jump at about 27 degrees C, apparently in accordance with the phase transition of the bulk lipid. Results obtained in this study suggested that the Ca2+-ATPase protein molecules are in an aggregated state in these reconstituted membranes and that the Ca2+-ATPase activity is neither directly correlated to the fluidity of the boundary lipid nor to the rotational mobility of the Ca2+-ATPase, contrary to the suggestions of previous investigators (Hesketh et al. (1976) Biochemistry 15, 4145-4151; Hidalgo et al. (1978) J. Biol. Chem. 253, 6879-6887).  相似文献   

9.
Adsorption of reovirus by minerals and soils.   总被引:7,自引:1,他引:7       下载免费PDF全文
Adsorption of [35S]methionine-labeled reovirus by 30 dry soils, minerals, and finely ground rocks suspended in synthetic freshwater at pH 7 was investigated to determine the conditions necessary for optimum virus removal during land application of wastewaters. All of the minerals and soils studied were excellent adsorbents of reovirus, with greater than 99% of the virus adsorbed after 1 h at 4 degrees C. Thereafter, virus remaining in suspension was significantly inactivated, and within 24 h a three to five log10 reduction in titer occurred. The presence of divalent cations, i.e., Ca2+ and Mg2+, in synthetic freshwater enhanced removal, whereas soluble organic matter decreased the amount of virus adsorbed in secondary effluent. The amount of virus adsorbed by these substrates was inversely correlated with the amount of organic matter, capacity to adsorb cationic polyelectrolyte, and electrophoretic mobility. Adsorption increased with increasing available surface area, as suspended infectivity was reduced further by the more finely divided substrates. However, the organic content of the soils reduced the level of infectious virus adsorbed below that expected from surface area measurements alone. The inverse correlation between virus adsorption and substrate capacity for cationic polyelectrolyte indicates that the adsorption of infectious reovirus particles is predominately a charged colloidal particle-charged surface interaction. Thus, adsorption of polyelectrolyte may be useful in predicting the fate of viruses during land application of sewage effluents and sludges.  相似文献   

10.
Sharp JS  Forrest JA  Jones RA 《Biochemistry》2002,41(52):15810-15819
We consider the effects that different lipid surfaces have upon the denaturation and subsequent formation of amyloid fibrils of bovine insulin. The adsorption and unfolding kinetics of insulin being adsorbed onto the different lipid surfaces under denaturing conditions are studied using FTIR ATR spectroscopy and are compared to the bulk solution behavior of the protein. Atomic force microscopy studies are also performed to compare the fibrils growing on the different surfaces. This study shows that both the adsorption and unfolding kinetics of insulin can be described by a sum of exponential processes and that different surfaces behave differently, with respect both to one another and to the bulk protein solution. The proteins adsorbed onto the surfaces are observed to have faster unfolding kinetics than those in the bulk, and the fibril-like structures formed at the surfaces are shown to be different in a number of ways from those found in bulk solution. The beta-sheet content and growth kinetics of the adsorbed proteins also differ from those of the bulk system. An attempt is made to describe the observed behavior in terms of simple physical arguments involving adsorption, unfolding, and aggregation of the proteins.  相似文献   

11.
With the limited stirring procedure used in the present work, substrate and enzyme together form a segregated and well-defined system on the surface. The lipase molecules responsible for the lipolysis are only those that are adsorbed on the glyceride monolayer. After a study of the stirring procedure, two series of systematic experiments were done: a) the bulk concentration of the enzyme was varied with different constant surface concentrations of the substrate, and b) the surface concentration of the substrate was varied with different constant bulk concentrations of the enzyme. The influence of the surface concentration of substrate on a) the rate of lipolysis, V,; b) the enzyme activity, a,; and c) the enzyme adsorption, Ze, were each determined by a different procedure. The values obtained verify the enzymic activity equation (a = V/Ze). The roles of other factors (Ca2+ ions, and pH) which govern the adsorption of the enzyme and its specific activity were also studied in preliminary experiments.  相似文献   

12.
The interaction of a peripheral protein with a lipid-water interface can show a pronounced dependence on the composition and two-dimensional packing density of the lipids that comprise the interface. We report a novel optical method for measuring the adsorption of macromolecules, such as proteins and nucleic acids, and smaller solutes, such as drugs, to lipid monolayers at the gas-liquid interface. Using fluorescence emission from proteins and a small molecule, we demonstrate that the emissions from these solutes when in the aqueous phase and when associated with the monolayer can be temporally separated. Such separation allows measurement of the extent of solute adsorption, spectral characterization of the adsorbed solute, and characterization of lipid organization using adsorption kinetics. The method does not require, but is compatible with, the solute having different spectral properties in the bulk and surface phases. Indeed, if optical signals from adsorbed and soluble solute are the same or their relationship is known, absolute surface excess of adsorbed solute can be calculated without independent calibration. With appropriate instrumental configuration, the method should be adaptable for screening solutes for interaction with planar monolayers having both well-defined composition and adjustable lipid packing density.  相似文献   

13.
The fluorescent probe 2-(p-toluidinyl)-naphthalene-6-sodium sulfonate (TNS) was used to study the surface adsorption of sublytic concentrations of the anionic surfactant sodium dodecylsulfonate (C(12)-SO(3)) on phosphatidylcholine (PC) bilayers. The number of adsorbed molecules was quantified by determination of the electrostatic potential (psi(o)) of the bilayers. The abrupt decrease in the fluorescence intensity detected even 10 s after the surfactant addition and the slight fluorescence variations with time indicated that the surfactant adsorption was very fast and almost complete. For a given number of monomers adsorbed a linear dependence between the lipid and C12-SO3 concentrations was obtained, indicating similar adsorption mechanism regardless of the surfactant concentration. Hence, a monomeric adsorption is assumed even in systems with a C12-SO3 concentration above its CMC. In addition, this linear correlation allowed us to determine the surfactant/lipid molar ratios (Re) (inversely related to the C12-SO3 ability to be adsorbed on liposomes) and the bilayer/aqueous phase coefficients (K). The fact that the lowest values for Re were always reached after 10 s of incubation corroborates the rapid kinetics of the process. The decrease in the C12-SO3 partitioning (K) when the number of surfactant molecules exceeded 15000 was possibly due to the electrostatic repulsion between the free and the adsorbed monomers, which could hinder the incorporation of new monomers on the charged surface of liposomes.  相似文献   

14.
Sialic acid (14.6 mug/mg protein) was quantitated in the non-cellular material removed from the lung of Beagle dogs by lavage. Sialic acid did not affect the dynamic surface tension properties of either the total alveolar lipid removed by lavage or of its major lipids, dipalmitoyl lecithin (DPL) and dipalmitoyl phosphatidyl glycerol (DPG). The presence of divalent cation (Ca2+, Mg2+, Mn2+ and Zn2+) or a lowered pH in the subphase medium lowered the surface tension during the expansion phase of the total alveolar lipid film when it was compressed and expanded on a Wilhelmy trough. Films of DPL behaved similarly, but no pH effect was observed with DPG monolayers. The cation effect manifested itself in the same direction as the value of the individual stability constants (Zn2+ greater than Mn2+ greater than Mg2+ greater than Ca2+) which suggests ionic binding of the cations to the phosphate group of the phospholipids. A physiological advantage of such an effect may lie in the conservation of the energetically favorable low surface tension state achieved during film compression with a minimum of surfactant lipid.  相似文献   

15.
The relationship between Ca2+- and hydroxyapatite-binding sites in salivary acidic proline-rich phosphoproteins A and C was investigated. Coating of hydroxyapatite with protein before adsorption had no effect on Ca2+ binding to the mineral, but simultaneous adsorption of Ca+ and protein to hydroxyapatite caused additional Ca2+ binding to the solid. The additional amount of Ca2+ adsorbed, measured in mol of Ca2+/mol of protein adsorbed to hydroxyapatite, was approx. 2 for protein C, 4 for protein A, 9 for the N-terminal tryptic peptide and 2 for dephosphorylated protein A. It is suggested that the ability of the proteins to inhibit hydroxyapatite formation is related to the binding of the proteins to crystal growth sites on the mineral, which prevents access of Ca2+ from the surrounding liquid.  相似文献   

16.
R Marsault  M Murgia  T Pozzan    R Rizzuto 《The EMBO journal》1997,16(7):1575-1581
Theoretical models and indirect experimental observations predict that Ca2+ concentrations at the inner surface of the plasma membrane may reach, upon stimulation, values much higher than those of the bulk cytosol. In the past few years, we have shown that the Ca2+-sensitive photoprotein aequorin can be intracellularly targeted and utilized for specifically monitoring the [Ca2+] of various organelles. In this work, we extend this approach to the study of the cytoplasmic rim beneath the plasma membrane. We have constructed a new aequorin chimera by fusing the photoprotein with SNAP-25, a neuronal protein which is recruited to the plasma membrane after the post-translational addition of a lipid anchor. The SNAP-25-aequorin chimera, expressed in the rat aortic smooth muscle cell line A7r5, appears correctly sorted as revealed by immunocytochemistry. Using this probe, we demonstrate that the mean [Ca2+] of this cytoplasmic region ([Ca2+]pm) can reach values >10-fold higher than those of the bulk cytosol ([Ca2+]c) upon activation of Ca2+ influx through plasma membrane channels. In unstimulated cells, the mean [Ca2+]pm appears also to be higher than the bulk cytosol, presumably reflecting the existence of microdomains of high [Ca2+].  相似文献   

17.
Neutral phospholipids play an important role in Ca2+ binding to biomembranes, in particular if the membrane carries a net negative surface charge due to charged lipids or proteins. The concentration of Ca2+ ions in the plane of the phospholipid head groups can be enhanced by at least two orders of magnitude compared to bulk solution. Ca2+ binding furthermore changes the orientation of the phospholipid head groups which is accompanied by variations of the local membrane dipole potential of the order of 10(5) V/cm. Such high electric fields could entail conformational changes of membrane-bound proteins and the Ca2(+)-induced reorientation of the lipid dipoles could thus play a regulatory role in membrane function.  相似文献   

18.
Lipid dependence of surface conformations of protein kinase C   总被引:1,自引:0,他引:1  
The change of conformation of protein kinase C interacting with the surface of a mercury electrode directly from a solution or through a lipid monolayer was inferred from the number of cystine residues exposed and reduced on the electrode and from their reduction potentials. Soluble protein kinase C was estimated to have 5-6 disulfide bonds which could potentially react with the mercury electrode. Two major reduction peaks of cystine at different microenvironments within the protein molecule adsorbed to a mercury surface. They were observed in a.c. polarograms and cyclic voltamograms at two distinct potentials. The potential of these peaks became more negative as the pH of the solution increased, which was consistent with relaxation or decrease in alpha-helicity (ordered structure) of the protein as determined by circular dichroism (CD) estimations of secondary structure. The peak at the more positive potentials (-0.46 V relative to NAg/AgCl electrode at pH 7.4) tended to vanish upon cyclic reduction and reoxidation of the cystine, while the more negative peak (-0.62 V at pH 7.4) was enhanced. Addition of Mg2+ or Ca2+ had no significant effect on the potential but there was a reduction in their amplitude which appeared to affect the disappearance of these peaks upon pH adjustment. This suggests that the tertiary structure of the molecule is stabilized by Ca2+ and Mg2+, as substantiated by CD spectral analysis of secondary structures. Protein kinase C penetrated lipid monolayers to some extent. Addition of diacylglycerol or phorbol ester to the lipid monolayers facilitated this penetration. These compounds stabilized the protein surface conformation by destabilizing the monolayer at more positive potentials, resulting in an enhanced reduction peak at -0.42 V. This phenomenon was not significantly affected by Mg2+ or by Ca2+. The region of the protein kinase C (PKC) sequence which penetrated the monolayer contains cysteines and a primary amine(s), and may have homology to a region of phospholipase A2 which has been proposed as a phospholipid binding site for the two enzymes. Additionally, these polarographic studies suggest that PKC associates with and penetrates monolayers in a divalent cation-independent manner in agreement with our previous physical analyses of PKC interactions with lipid bilayers.  相似文献   

19.
Pulmonary surfactant is a lipid:protein complex containing dipalmitoyl-phosphatidylcholine (DPPC) as the major component. Recent studies indicate adsorbed surfactant films consist of a surface monolayer and a monolayer-associated reservoir. It has been hypothesized that the monolayer and its functionally contiguous reservoir may be enriched in DPPC relative to bulk phase surfactant. We investigated the compositional relationship between the monolayer and its reservoir using paper-supported wet bridges to transfer films from adsorbing dishes to clean surfaces on spreading dishes. Spreading films appear to form monolayers in the spreading dishes. We employed bovine lipid extract surfactant [BLES(chol)] containing [3H]DPPC and either [14C]palmitoyl, oleoyl-phosphatidylcholine (POPC), [14C]dipalmitoyl-phosphatidylglycerol (DPPG), [14C]palmitoyl, oleoyl-phosphatidylglycerol (POPG), or [14C]cholesterol. Radiolabeled phosphatidylglycerols were prepared using phospholipase D. The studies demonstrated that the [3H]DPPC-[14C] POPC ratios were the same in the prepared BLES dispersions as in Langmuir-Blodgett films, indicating a lack of DPPC selectivity during film formation. Furthermore, identical 3H-14C isotopic ratios were observed with DPPC and either 14C-labeled POPC, DPPG, POPG, or cholesterol in the original dispersions, the bulk phases in adsorption dish D1, and monolayers recovered from spreading dish D2. These relationships remained unperturbed with 2-fold increases in bulk concentrations in D1 and 10-fold variations in D1-D2 surface area. These results indicate adsorbed surfactant monolayers and their associated reservoirs possess similar lipid compositions and argue against selective adsorption of DPPC.  相似文献   

20.
The adsorption of different proteins in a single biospecific and hydrophobic adsorbent particle for preparative protein chromatography has been observed directly by confocal laser scanning microscopy as a function of time at a constant bulk concentration c(b). The bulk concentration was in the non-linear part of the adsorption isotherm. At all times the concentration of free protein at the particle surface was almost equal to the bulk content indicating that external mass transfer resistance is not rate limiting for the adsorption under these conditions. Inside the particles a distinct maximum in adsorbed and free protein concentration that moved inside to a distance of approximately 0.2 R (R particle radius) from the particle surface, was observed. This is due to a decreasing solid-phase density and adsorptive capacity in the particle between 0.8 R and R indicating that the fraction of macropores (or void space) is larger in the outer than in the inner part of the adsorbent particles. By increasing the bulk concentration by a factor of 10 the equilibration time was reduced by about the same magnitude. This is in agreement with the concentration dependence of the effective pore diffusion coefficient D(p,eff)=D(p)/[epsilon(p)[1+nK/(K +c)(2)]] derived from the mass conservation relations describing the adsorption process. The time dependence protein adsorption up to approximately 90% of the equilibration value q* could be described by a bilinear free driving force model. The rapid equilibration in the outer part of the particle with a half-life time of approximately 100 s in the studied systems accounted for 0.3-0.4 q*. The slower equilibration with a up to ten times longer half-life time, was the adsorption in the inner part of the particle that outside 0.5 R accounts for 0.5-0.6 q*. These data were compared with literature data for batch adsorption of proteins in biospecific, hydrophobic and ion-exchange adsorbents. They could also be described by a bilinear free driving force model, with about the same quantitative results as obtained for similar conditions in the single particle experiments. The static adsorption parameters, maximum binding site concentration n, and dissociation constant for the protein binding to a binding site K, were determined from Scatchard plots. For the same protein-adsorbent system the plots changed from linear to non-linear with increasing n. This change occurred when the average distance between adjacent binding sites become of the same order of magnitude as the size of the binding site or adsorbed protein. This causes a shielding of free binding sites increasing with n and the concentration of adsorbed protein, yielding a concentration dependence in K. These results show that for a high throughput and rapid adsorption in preparative chromatography, the adsorption step should be carried out in the non-linear part of the adsorption isotherm with concentrations up to c(b) where q*/c(b)>/=10 to obtain high protein recoveries. To avoid tailing due to the flow of adsorbed proteins in the inner part of the particles further into the particles at the start of the desorption, and to speed up desorption rates, protein adsorption in the particle within 0.5 R from the particle center should be avoided. This requires the further development of suitable pellicular particles for preparative protein chromatography that meet this requirement.  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