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.
Adsorption of plasmid DNA to mineral surfaces and protection against DNase I.   总被引:14,自引:0,他引:14  
The adsorption of [3H]thymidine-labeled plasmid DNA (pHC314; 2.4 kb) of different conformations to chemically pure sand was studied in a flowthrough microenvironment. The extent of adsorption was affected by the concentration and valency of cations, indicating a charge-dependent process. Bivalent cations (Mg2+, Ca2+) were 100-fold more effective than monovalent cations (Na+, K+, NH4+). Quantitative adsorption of up to 1 microgram of negatively supercoiled or linearized plasmid DNA to 0.7 g of sand was observed in the presence of 5 mM MgCl2 at pH 7. Under these conditions, more than 85% of DNA adsorbed within 60 s. Maximum adsorption was 4 micrograms of DNA to 0.7 g of sand. Supercoil molecules adsorbed slightly less than linearized or open circular plasmids. An increase of the pH from 5 to 9 decreased adsorption at 0.5 mM MgCl2 about eightfold. It is concluded that adsorption of plasmid DNA to sand depends on the neutralization of negative charges on the DNA molecules and the mineral surfaces by cations. The results are discussed on the grounds of the polyelectrolyte adsorption model. Sand-adsorbed DNA was 100 times more resistant against DNase I than was DNA free in solution. The data support the idea that plasmid DNA can enter the extracellular bacterial gene pool which is located at mineral surfaces in natural bacterial habitats.  相似文献   

2.
Adsorption of DNA to sand and variable degradation rates of adsorbed DNA   总被引:11,自引:0,他引:11  
Adsorption and desorption of DNA and degradation of adsorbed DNA by DNase I were studied by using a flowthrough system of sand-filled glass columns. Maximum adsorption at 23 degrees C occurred within 2 h. The amounts of DNA which adsorbed to sand increased with the salt concentration (0.1 to 4 M NaCl and 1 mM to 0.2 M MgCl2), salt valency (Na+ less than Mg2+ and Ca2+), and pH (5 to 9). Maximum desorption of DNA from sand (43 to 59%) was achieved when columns were eluted with NaPO4 and NaCl for 6 h or with EDTA for 1 h. DNA did not desorb in the presence of detergents. It is concluded that adsorption proceeded by physical and chemical (Mg2+ bridging) interaction between the DNA and sand surfaces. Degradability by DNase I decreased upon adsorption of transforming DNA. When DNA adsorbed in the presence of 50 mM MgCl2, the degradation rate was higher than when it adsorbed in the presence of 20 mM MgCl2. The sensitivity to degradation of DNA adsorbed to sand at 50 mM MgCl2 decreased when the columns were eluted with 0.1 mM MgCl2 or 100 mM EDTA before application of DNase I. This indicates that at least two types of DNA-sand complexes with different accessibilities of adsorbed DNA to DNase I existed. The degradability of DNA adsorbed to minor mineral fractions (feldspar and heavy minerals) of the sand differed from that of quartz-adsorbed DNA.  相似文献   

3.
Adsorption and desorption of DNA and degradation of adsorbed DNA by DNase I were studied by using a flowthrough system of sand-filled glass columns. Maximum adsorption at 23 degrees C occurred within 2 h. The amounts of DNA which adsorbed to sand increased with the salt concentration (0.1 to 4 M NaCl and 1 mM to 0.2 M MgCl2), salt valency (Na+ less than Mg2+ and Ca2+), and pH (5 to 9). Maximum desorption of DNA from sand (43 to 59%) was achieved when columns were eluted with NaPO4 and NaCl for 6 h or with EDTA for 1 h. DNA did not desorb in the presence of detergents. It is concluded that adsorption proceeded by physical and chemical (Mg2+ bridging) interaction between the DNA and sand surfaces. Degradability by DNase I decreased upon adsorption of transforming DNA. When DNA adsorbed in the presence of 50 mM MgCl2, the degradation rate was higher than when it adsorbed in the presence of 20 mM MgCl2. The sensitivity to degradation of DNA adsorbed to sand at 50 mM MgCl2 decreased when the columns were eluted with 0.1 mM MgCl2 or 100 mM EDTA before application of DNase I. This indicates that at least two types of DNA-sand complexes with different accessibilities of adsorbed DNA to DNase I existed. The degradability of DNA adsorbed to minor mineral fractions (feldspar and heavy minerals) of the sand differed from that of quartz-adsorbed DNA.  相似文献   

4.
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   

5.
It is known that plasmid DNA and linear duplex DNA molecules adsorb to chemically purified mineral grains of sand and to particles of several clay fractions. It seemed desirable to examine whether plasmid DNA would also adsorb to nonpurified mineral materials taken from the environment and, particularly, whether adsorbed plasmid DNA would be available for natural transformation of bacteria. Therefore, microcosms consisting of chemically pure sea sand plus buffered CaCl2 solution were compared with microcosms consisting of material sampled directly from a groundwater aquifer (GWA) plus groundwater (GW) with respect to the natural transformation of Acinetobacter calcoaceticus by mineral-associated DNA. The GWA minerals were mostly sand with inorganic precipitates and organic material plus minor quantities of silt and clay (illite and kaolinite). The amount of plasmid DNA which adsorbed to GWA (in GW) was about 80% of the amount which adsorbed to purified sand (in buffered CaCl2 solution). Plasmid DNA adsorbed on sand transformed A. calcoaceticus significantly less efficiently than did plasmid DNA in solution. In contrast, the transformation by sand-adsorbed chromosomal DNA was as high as that by DNA in solution. In GWA/GW microcosms, the efficiency of transformation by chromosomal DNA was similar to that in sand microcosms, whereas plasmid transformation was not detectable. However, plasmid transformants were found at a low frequency when GWA was loaded with both chromosomal and plasmid DNA. Reasons for the low transformation efficiency of plasmid DNA adsorbed to mineral surfaces are discussed. Control experiments showed that the amounts of plasmid and chromosomal DNA desorbing from sand during incubation with a cell-free filtrate of a competent cell suspension did not greatly contribute to transformation in sand microcosms, suggesting that transformation occurred by direct uptake of DNA from the mineral surfaces. Taken together, the observations suggest that plasmid DNA and chromosomal DNA fragments which are adsorbed on mineral surfaces in a sedimentary or soil habitat may be available (although with different efficiencies for the two DNA species) for transformation of a naturally competent gram-negative soil bacterium.  相似文献   

6.
Monovalent ([Na+] > 10 mM) and divalent ([Ca2+], [Mg2+] > 1.0 mM) cations induced the precipitationof nucleic acid molecules. In the presence of clay minerals (montmorillonite and kaolinite), there was adsorption instead of precipitation. The cation concentration needed for adsorption depended on both the valence of the cations and the chemical nature of the nucleic acid molecules. Double-stranded nucleic acids needed higher cation concentrations than single-stranded ones to be adsorbed to the same extent on clay. Divalent cations were more efficient than monovalent ones in mediating adsorption. Adsorption to the clay occurred only when both nucleic acids and cations were present. However, once the complexes were formed, the cations could not be removed from the system by washing, indicating that they are directly involved in the association between nucleic acids and mineral surfaces.These observations indicate that cations take part directly in the formation of nucleic acid-clay complexes, acting as a `bridge' between the negative charges on the mineral surface and those of the phosphate groups of the genetic polymer. The relatively low cation concentrations needed for adsorption and the ubiquitous presence of clay minerals in the environment suggest that the adsorption of nucleic acids on mineral surfaces could have taken place in prebiotic habitats. This may have played an important role in the formation and preservation of nucleic acids and/or their precursor polymers.  相似文献   

7.
M V Norgard  K Keem  J J Monahan 《Gene》1978,3(4):279-292
The susceptibility of E. coli strain chi1776 to transformation by pBR322 plasmid DNA was examined and optimized. Maximum transformation to tetracycline (Tc) resistance was achieved when cells were harvested from L broth at 5.0--6.0 . 10(7) cfu/ml, followed by washing twice in cold 0.1 M NaCl + 5 mM MgCl2 + 5 mM Tris, pH 7.6. Cells grown in the presence of D-cycloserine (Cyc) rather than nalidixic acid (Nx) transformed markedly better. The presence of 5 mM Mg2+ ions in washing and CaCl2 solutions stimulated transformation about 2-fold. Optimal conditions for transformation included a pH range of 7.25-7.75 and a cell-to-DNA ratio of about 1.6 . 10(8) cfu/ng plasmid DNA. The frequency of transformation was highest when cells were exposed to 100 mM CaCl2 in 250 mM KCl + 5 mM MgCl2 + 5 mM Tris, pH 7.6, before mixing with DNA. A 60 min incubation period for cell + DNA mixtures held on ice produced the maximum number of Tcr transformants. In our hands, heat shocks at 37 degrees C or 42 degrees C for various times all decreased transformation to about one-half of optimal levels. Furthermore, the recovery of transformants was best when cell + DNA mixtures were plated on precooled (4 degrees C) Tc agar plates. The efficiency of plating was optimum when only 5 microliter of cell + DNA mixture was spread per plate, suggesting that non-viable background chi1776 cells on selective medium inhibited the recovery of transformants. It was also found that the presence of linear DNA molecules in cell + DNA mixtures markedly inhibited the transformation of chi1776 by pBR322 plasmid DNA. On the basis of these findings, a new procedure for the plasmid-specific transformation of E. coli chi1776 by pBR322 plasmid DNA is proposed. The use of this technique has allowed us to attain transformation frequencies in excess of 10(7) transformants/microgram pBR322 plasmid DNA.  相似文献   

8.
Prokaryotes can exchange chromosomal and plasmid genes via extracellular DNA in a process termed genetic transformation. This process has been observed in the test tube for several bacterial species living in the environment but it is not clear whether transformation occurs in natural bacterial habitats. A major constituent of terrestrial environments are solid particles such as quartz, silt and clay, which have considerable surface areas and which make up the solid-liquid interfaces of the habitat. In previous experiments the adsorption of DNA to chemically purified quartz and clay minerals was shown and the partial protection of adsorbed DNA against DNAase I. In a microcosm consisting of natural groundwater aquifer material (GWA) sampled directly from the environment and groundwater (GW) both linear duplex and supercoiled plasmid DNA molecules bound rapidly and quantitatively to the minerals. The divalent cations required to form the association were those present in the GWA/GW microcosm. The association was stable to extended elution over one week at 23°C. Upon adsorption, the DNA became highly resistant against enzymatic degradation. About 1000 times higher DNAase I concentrations were needed to degrade bound DNA to the same extent as DNA dissolved in GW. Furthermore, chromosomal and plasmid DNA bound on GWA transformed competent cells of Bacillus subtilis. However, in contrast to DNA in solution, on GWA the chromosomal DNA was more active in transformation than the plasmid DNA. The studies also revealed that in the transformation of B. subtilis Mg2+ can be replaced by Na+, K+ or NH4 The observations suggest that in soil and sediment environments, mineral material with inorganic precipitates and organic matter can harbour extracellular DNA leaving it available for genetic transformation.  相似文献   

9.
Condensation of supercoiled DNA induced by MnCl2.   总被引:6,自引:0,他引:6       下载免费PDF全文
Multivalent cations condense DNA in vitro, but it had been thought that a valence of at least + 3 was required in aqueous solution. We have found that Mn2+ can produce toroidal condensates of supercoiled plasmid DNA, but not of linearized plasmid. Mg2+ does not cause condensation, and neither MgCl2 nor NaCl can negate the effect of MnCl2, indicating that the condensation mechanism with Mn is not primarily electrostatic. Supercoiled MnDNA is more extensively digested than the linear form by S1 nuclease. Supercoiling appears to cooperate with Mn2+ in stabilizing helix distortions and also provides a "pressure" that enhances lateral association.  相似文献   

10.
We used a molecular beacon (MB) containing a 15-mer triplex-forming oligonucleotide (TFO) to probe in real-time the kinetics of triplex DNA formation in the left side of the TCl tract (502-516) of the c-src proto-oncogene in vitro. The metal ions Na+, K+, and Mg2+ stabilized triplex DNA at this site. The pseudo-first-order rate constant (kpsi) and the second-order association rate constant (k1) for the binding of the MB to the target duplex in 10 mM sodium phosphate buffer, pH 7.3, increased from 3.2 +/- 0.9 to 15 +/- 2.8 x 10(-3) s(-1) and 6.4 +/- 1.8 to 30 +/- 5.6 x 102 M(-1) s(-1), respectively, on increasing the MgCl2 concentration from 1 to 2.5 mM. Similar values were obtained for the triplex DNA stabilized by NaCl (100-250 mM). Surprisingly, the values were around 2 times higher in the presence of KCl. The AG of triplex formation in the presence of 1 mM MgCl2, 150 mM NaCl, and 150 mM KCl were -7.8 +/- 0.3, -8.2 +/- 0.3 and -8.7 +/- 0.7 kcal/mol respectively, despite significant differences in the values of deltaH and deltaS, suggesting enthalpy-entropy compensation in the stabilization of the triplex DNA by these metal ions. These results show the utility of MBs ih probing triplex DNA formation and in evaluating kinetic and thermodynamic parameters important for the design and development of TFOs as triplex DNA-based therapeutic agents.  相似文献   

11.
Winding of the DNA helix by divalent metal ions.   总被引:1,自引:0,他引:1       下载免费PDF全文
Y C Xu  H Bremer 《Nucleic acids research》1997,25(20):4067-4071
When supercoiled pBR322 DNA was relaxed at 0 or 22 degrees C by topoisomerase I in the presence of the divalent cations Ca2+, Mn2+ or Co2+, the resulting distributions of topoisomers observed at 22 degrees C had positive supercoils, up to an average delta Lk value of +8.6 (for Ca2+at 0 degrees C), corresponding to an overwinding of the helix by 0.7 degrees/bp. An increase of the divalent cation concentration in the reaction mixture above 50 mM completely reversed the effect. When such ions were present in agarose electrophoresis gels, they caused a relaxation of positively supercoiled DNA molecules, and thus allowed a separation of strongly positively supercoiled topoisomers. The effect of divalent cations on DNA adds a useful tool for the study of DNA topoisomers, for the generation as well as separation of positively supercoiled DNA molecules.  相似文献   

12.
The R-form lipopolysaccharide (LPS) from Klebsiella strain LEN-111 (O3-:K1-) forms a hexagonal lattice structure with a lattice constant of 14 to 15 nm when it is precipitated by addition of two volumes of 10 mM MgCl2-ethanol. When the LPS was suspended in various buffers (50 mM) at pH 2 to 12 for 24 hr at 4 C, at pH 2 and 3 pits of the hexagonal lattice structure markedly disappeared, at pH 4 to 8.5 the lattice structure was stable, and at pH 9 to 12 it tended to loosen somewhat. The LPS from which cations were removed by electrodialysis retained the ability of hexagonal assembly, although the lattice constant of the hexagonal lattice of the electrodialyzed LPS was large. The lattice structure of the electrodialyzed LPS was much more labile than that of the non-electrodialyzed LPS at alkaline pH levels and the former was completely disintegrated into ribbon-like structures when the LPS was suspended in 50 mM Tris buffer at pH 7.7 or higher. However, the electrodialyzed LPS formed a hexagonal lattice structure in Tris buffer at pH 8.5 containing 0.1 to 100 mM MgCl2. The lattice constants of the hexagonal lattice formed by the electrodialyzed LPS at 10 or 100 mM MgCl2 were very similar to that of the lattice of the non-electrodialyzed LPS. From these results it is concluded that the lability of the hexagonal lattice structure of the electrodialyzed LPS at alkaline conditions is due to removal of Mg2+ by electrodialysis.  相似文献   

13.
The effects of increasing concentrations of magnesium (Mg2+), calcium (Ca2+) or EDTA, and pH on the adhesion of five slime-positive strains of Staphylococcus epidermidis (Se+) to plastic were examined using an in vitro microwell assay. The addition of Mg2+ (as either MgSO4 or MgCl2) to the bacterial suspension in concentrations as low as 16 microM significantly enhanced the adhesion of all test strains to plastic (P < 0.001). Similarly, the addition of Ca2+ (as CaCl2) in concentrations exceeding 128 microM produced a significant increase in the adhesion of all test strains, but not to the extent observed with Mg2+. In contrast, the adhesion of all test strains to plastic was significantly reduced in the presence of EDTA at concentrations greater than 8 mM. However, EDTA in concentrations as low as 0.25 mM caused a significant decrease in the adhesion of two strains of Se+. The effect of pH was variable, but at a pH of 5.0 and 6.0, the adhesion of all test strains was significantly reduced compared to control values at a pH of 7.0. Two strains showed a significant increase in adhesion at a pH of 8.0. We also compared the effects of these variables on the adherence of a slime-negative phase variant derived from a slime-positive parent strain. With the exception of pH, the adhesion of both strains in response to increasing divalent cations or EDTA was similar. These data indicate that, in addition to hydrophobic interactions, ligand-specific binding, and slime production, pH and divalent cations, especially Mg2+, are important determinants of the adhesion of S. epidermidis to plastic surfaces in vitro.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
In order to determine the mechanisms involved in the persistence of extracellular DNA in soils and to monitor whether bacterial transformation could occur in such an environment, we developed artificial models composed of plasmid DNA adsorbed on clay particles. We determined that clay-bound DNA submitted to an increasing range of nuclease concentrations was physically protected. The protection mechanism was mainly related to the adsorption of the nuclease on the clay mineral. The biological potential of the resulting DNA was monitored by transforming the naturally competent proteobacterium Acinetobacter sp. strain BD413, allowing us to demonstrate that adsorbed DNA was only partially available for transformation. This part of the clay-bound DNA which was available for bacteria, was also accessible to nucleases, while the remaining fraction escaped both transformation and degradation. Finally, transformation efficiency was related to the perpetuation mechanism, with homologous recombination being less sensitive to nucleases than autonomous replication, which requires intact molecules.  相似文献   

15.
General properties of ouabain-sensitive K+ binding to purified Na+,K+-ATPase [EC 3.6.1.3] were studied by a centrifugation method with 42K+. 1) The affinity for K+ was constant at pH values higher than 6.4, and decreased at pH values lower than 6.4. 2) Mg2+ competitively inhibited the K+ binding. The dissociation constant (Kd) for Mg2+ of the enzyme was estimated to be about 1 mM, and the ratio of Kd for Mg2+ to Kd for K+ was 120 : 1. The order of inhibitory efficiency of divalent cations toward the K+ binding was Ba2+ congruent to Ca2+ greater than Zn2+ congruent to Mn2+ greater than Sr2+ greater than Co2+ greater than Ni2+ greater than Mg2+. 3) The order of displacement efficiency of monovalent cations toward the K+ binding in the presence or absence of Mg2+ was Tl+ greater than Rb+ greater than or equal to (K+) greater than NH4+ greater than or equal to Cs+ greater than Na+ greater than Li+. The inhibition patterns of Na+ and Li+ were different from those of other monovalent cations, which competitively inhibited the K+ binding. 4) The K+ binding was not influenced by different anions, such as Cl-, SO4(2-), NO3-, acetate, and glycylglycine, which were used for preparing imidazole buffers. 5) Gramicidin D and valinomycin did not affect the K+ binding, though the former (10 micrograms/ml) inhibited the Na+,K+-ATPase activity by about half. Among various inhibitors of the ATPase, 0.1 mM p-chloromercuribenzoate and 0.1 mM tri-n-butyltin chloride completely inhibited the K+ binding. Oligomycin (10 micrograms/ml) and 10 mM N-ethylmaleimide had no effect on the K+ binding. In the presence of Na+, however, oligomycin decreased the K+ binding by increasing the inhibitory effect of Na+, whether Mg2+ was present or not. 6) ATP, adenylylimido diphosphate and ADP each at 0.2 mM decreased the K+ binding to about one-fourth of the original level at 10 microM K+ without MgCl2 and at 60 microM K+ with 5 mM MgCl2. On the other hand, AMP, Pi, and p-nitrophenylphosphate each at 0.2 mM had little effect on the K+ binding.  相似文献   

16.
In order to determine the mechanisms involved in the persistence of extracellular DNA in soils and to monitor whether bacterial transformation could occur in such an environment, we developed artificial models composed of plasmid DNA adsorbed on clay particles. We determined that clay-bound DNA submitted to an increasing range of nuclease concentrations was physically protected. The protection mechanism was mainly related to the adsorption of the nuclease on the clay mineral. The biological potential of the resulting DNA was monitored by transforming the naturally competent proteobacterium Acinetobacter sp. strain BD413, allowing us to demonstrate that adsorbed DNA was only partially available for transformation. This part of the clay-bound DNA which was available for bacteria, was also accessible to nucleases, while the remaining fraction escaped both transformation and degradation. Finally, transformation efficiency was related to the perpetuation mechanism, with homologous recombination being less sensitive to nucleases than autonomous replication, which requires intact molecules.  相似文献   

17.
Catalytic properties of the HhaII restriction endonuclease   总被引:1,自引:0,他引:1  
The catalytic properties of the HhaII restriction endonuclease were studied using plasmid pSK11 DNA containing a single 5'-G-A-N-T-C HhaII cleavage site as substrate. Reactions were followed by two methods: 1) gel electrophoretic analysis of nicked circular and linear DNA products, or 2) release of 32P-labeled inorganic phosphate from specifically labeled HhaII sites in a reaction coupled with bacterial alkaline phosphatase. The enzyme is optimally active at 37 degrees C in 10 mM Tris-HCl (pH 9.1) and 4-10 mM MgCl2 without added NaCl. Activity is stabilized by the presence of 2-mercaptoethanol and 0.2% Triton X-100 or 50 microgram/ml bovine serum albumin. At enzyme concentrations below 10 nM and using pSK11 as substrate, initial kinetic rates were dependent on the order of mixing of reactants. A lag of 3-4 min was observed if enzyme or substrate was added last. Preincubation of substrate and enzyme followed by initiation of the reaction with MgCl2 or preincubation of the enzyme with nonspecific DNA followed by initiation with substrate eliminated or reduced the lag, respectively, and speeded up the reactions. Under a wide range of reaction conditions, nicked pSK11 DNA accumulated early, while linear molecules appeared later, suggesting that HhaII cleaves one strand at a time in separate binding events. The apparent Km for covalently closed pSK11 DNA molecules was approximately 17 nM, and the turnover number for the conversion of covalent to nicked sites was 1.1 single strand scissions/min. Pre-steady state kinetic analysis indicated that cleavage of the first phosphodiester bond in a site is first order with a rate constant of about 0.8 min-1, while cleavage of the second phosphodiester bond is first order with a rate constant of about 0.2 min-1.  相似文献   

18.
K Kato  M Goto  H Fukuda 《Life sciences》1983,32(8):879-887
When investigating the effects of divalent cations (Mg2+, Ca2+, Sr2+, Ba2+, Mn2+ and Ni2+) on 3H-baclofen binding to rat cerebellar synaptic membranes, we found that the specific binding of 3H-baclofen was not only dependent on divalent cations, but was increased dose-dependently in the presence of these cations. The effects were in the following order of potency: Mn2+ congruent to Ni2+ greater than Mg2+ greater than Ca2+ greater than Sr2+ greater than Ba2+. Scatchard analysis of the binding data revealed a single component of the binding sites in the presence of 2.5 mM MgCl2, 2.5 mM CaCl2 or 0.3 mM MnCl2 whereas two components appeared in the presence of 2.5 mM MnCl2 or 1 mM NiCl2. In the former, divalent cations altered the apparent affinity (Kd) without affecting density of the binding sites (Bmax). In the latter, the high-affinity sites showed a higher affinity and lower density of the binding sites than did the single component of the former. As the maximal effects of four cations (Mg2+, Ca2+, Mn2+ and Ni2+) were not additive, there are probably common sites of action of these divalent cations. Among the ligands for GABAB sites, the affinity for (-), (+) and (+/-) baclofen, GABA and beta-phenyl GABA increased 2-6 fold in the presence of 2.5 mM MnCl2, in comparison with that in HEPES-buffered Krebs solution (containing 2.5 mM CaCl2 and 1.2 mM MgSO4), whereas that for muscimol was decreased to one-fifth. Thus, the affinity of GABAB sites for its ligands is probably regulated by divalent cations, through common sites of action.  相似文献   

19.
Poliovirus adsorption by 34 minerals and soils   总被引:5,自引:0,他引:5  
The adsorption of radiolabeled infectious poliovirus type 2 by 34 well-defined soils and mineral substrates was analyzed in a synthetic freshwater medium containing 1 mM CaCl(2) and 1.25 mM NaHCO(3) at pH 7. In a model system, adsorption of poliovirus by Ottawa sand was rapid and reached equilibrium within 1 h at 4 degrees C. Near saturation, the adsorption could be described by the Langmuir equation; the apparent surface saturation was 2.5 x 10(6) plaque-forming units of poliovirus per mg of Ottawa sand. At low surface coverage, adsorption was described by the Freundlich equation. The soils and minerals used ranged from acidic to basic and from high in organic content to organic free. The available negative surface charge on each substrate was measured by the adsorption of a cationic polyelectrolyte, polydiallyldimethylammonium chloride. Most of the substrates adsorbed more than 95% of the virus. In general, soils, in comparison with minerals, were weak adsorbents. Among the soils, muck and Genesee silt loam were the poorest adsorbents; among the minerals, montmorillonite, glauconite, and bituminous shale were the least effective. The most effective adsorbents were magnetite sand and hematite, which are predominantly oxides of iron. Correlation coefficients for substrate properties and virus adsorption revealed that the elemental composition of the adsorbents had little effect on poliovirus uptake. Substrate surface area and pH, by themselves, were not significantly correlated with poliovirus uptake. A strong negative correlation was found between poliovirus adsorption and both the contents of organic matter and the available negative surface charge on the substrates as determined by their capacities for adsorbing the cationic polyelectrolyte, polydiallyldimethylammonium chloride.  相似文献   

20.
Soybean proteins have found uses in different nonfood applications due to their interesting properties. We report on the kinetics and extent of adsorption on silica and cellulose surfaces of glycinin and β-conglycinin, the main proteins present in soy. Quartz crystal microgravimetry (QCM) experiments indicate that soy protein adsorption is strongly affected by changes in the physicochemical environment. The affinity of glycinin and the mass adsorbed on silica and cellulose increases (by ca. 13 and 89%, respectively) with solution ionic strength (as it increases from 0 to 100 mM NaCl) due to screening of electrostatic interactions. In contrast, β-conglycinin adsorbs on the same substrates to a lower extent and the addition of electrolyte reduces adsorption (by 25 and 57%, respectively). The addition of 10 mM 2-mercaptoethanol, a denaturing agent, reduces the adsorption of both proteins with a significant effect for glycinin. This observation is explained by the cleavage of disulfide bonds which allows unfolding of the molecules and promotes dissociation into subunits that favors more compact adsorbed layer structures. In addition, adsorption of glycinin onto cellulose decreases with lowering the pH from neutral to pH 3 due to dissociation of the macromolecules, resulting in flatter adsorbed layers. The respective adsorption isotherms fit a Langmuir model and QCM shifts in energy dissipation and frequency reveal multiple-step kinetic processes indicative of changes in adlayer structure.  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