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.
Good's zwitterionic buffers are widely used in biological and biochemical research in which hydrogen peroxide is a solution component. This study was undertaken to determine whether Good's buffers exhibit reactivity toward H(2)O(2). It is found that H(2)O(2) oxidizes both morpholine ring-containing buffers (e.g., Mops, Mes) and piperazine ring-containing zwitterionic buffers (e.g., Pipes, Hepes, and Epps) to produce their corresponding N-oxide forms. The percentage of oxidized buffer increases as the concentration of H(2)O(2) increases. However, the rate of oxidation is relatively slow. For example, no oxidized Mops was detected 2h after adding 0.1M H(2)O(2) to 0.1M Mops (pH 7.0), and only 5.7% was oxidized after 24h exposure to H(2)O(2). Thus, although all of these buffers can be oxidized by H(2)O(2), their slow reaction does not significantly perturb levels of H(2)O(2) in the time frame and at the concentrations of most biochemical studies. Therefore, the previously reported rapid loss of H(2)O(2) produced from the ferroxidase reaction of ferritin is unlikely due to reaction of H(2)O(2) with buffer, a conclusion supported by the fact that H(2)O(2) is also lost rapidly when the solution pH of the ferroxidase reaction is controlled by a pH stat apparatus in the absence of buffer.  相似文献   

2.
The ω-aminohexyl diamine immobilized as ligand on CNBr- and bisoxirane-activated agarose gel was evaluated for the purification of human immunoglobulin G (IgG) from serum and plasma by negative affinity chromatography. The effects of matrix activation, buffer system, and feedstream on recovery and purity of IgG were studied. A one-step purification process using Hepes buffer at pH 6.8 allowed a similar recovery (69–76%) of the loaded IgG in the nonretained fractions for both matrices, but the purity was higher for epoxy-activated gel (electrophoretically homogeneous protein with a 6.5-fold purification). The IgG and human serum albumin (HSA) adsorption equilibrium studies showed that the adsorption isotherms of IgG and HSA obeyed the Langmuir–Freundlich and Langmuir models, respectively. The binding capacity of HSA was high (210.4 mg mL?1 of gel) and a positive cooperativity was observed for IgG binding. These results indicate that immobilizing ω-aminohexyl using bisoxirane as coupling agent is a useful strategy for rapid purification of IgG from human serum and plasma.  相似文献   

3.
《Free radical research》2013,47(3):149-160
Iron autoxidation in Mops and Hepes buffers is characterized by a lag phase that becomes shorter with increasing FeCl2 concentration and pH. During iron oxidation in these buffers a yellow colour develops in the solution. When the reaction is conducted in the presence of nitro blue tetrazolium (NBT), blue formazan is formed. Of the many OH' scavengers tested, mannitol and sorbitol are most effective in inhibiting Fe2+ oxidation, yellow colour development and NBT reduction. Some inhibition was also noted with catalase. The iron product of the oxidative reaction differs from Fe3+ in its absorption spectrum and its low reactivity with thiocyanate. Similar results are obtained when iron autoxidation is studied in unbuffered solutions brought to alkaline pH with NaOH. In phosphate buffer, no lag phase is evident and the absorption spectrum of the final solution is identical to that of Fe3+ in this buffer. The iron product reacts immediately with thiocyanate. When iron oxidation is conducted in the presence of NBT the formation of formazan is almost undetectable. Of the many compounds tested only catalase inhibits iron autoxidation in this buffer. The sequence of reactions leading to iron autoxidation in Good-type buffers1 thus resembles that occurring in unbuffered solutions brought to alkaline pH with NaOH and greatly differs from that occurring in phosphate buffer. These results are in agreement with the observation that these buffers have very low affinity for iron.1 The data presented define experimental conditions where Fe2+ is substantially stable for a considerable length of time in Mops buffer.  相似文献   

4.
l-Histidine, intended as a pseudobiospecific ligand, was immobilized on poly(ethylenevinyl alcohol) hollow fibre membranes after their activation with epichlorohydrin or butanediol diglycidyl ether. The affinity membranes obtained allowed the one-step separation of immunoglobulin G (IgG) from untreated human serum. Elution was possible under mild conditions with discontinuous pH or salt gradients. IgM was also adsorbed to a certain extent and partially separated from IgG by pH gradient elution. The bound IgG fractions showed pI values between 8 and 9.5 and contained IgG1 and IgG3. The dissociation constants for IgG on the bisoxirane- and epichlorohydrin-activated membranes coupled with histidine were determined by equilibrium binding analysis to be 2.5·10−5 and 2.0·10−5 M, respectively. The maximum binding capacity of the affinity hollow fibre membranes was 80 and 70 mg of IgG per gram of support, respectively. With a cartridge of surface area 1 m2 (about 19 g of fibres), during a 60-min run, theoretically up to 1.5 g of IgG can be removed from human serum. The histidine affinity membranes are very stable owing to the simple nature of the ligand and the coupling via an ether linkage. Reproducible results were obtained over more than 1 year even with untreated human serum being used regularly.  相似文献   

5.
The depyrogenation of different IgG solutions using the histidine-linked hollow fiber membrane developed in our laboratory is presented here. Three strategies for endotoxin (ET) removal were investigated according to the immobilized histidine's ability to bind different immunoglobulins: (1) ET removal from 1 mg/ml non histidine-binding mouse monoclonal IgG1 (MabCD4) solution was achieved in the presence of acetate buffer (pH 5.0) without any protein loss. (2) For contaminated human IgG, combined adsorption of ET and IgG in the presence of MOPS or Tris buffer was tested, followed by differential elution using increasing salt concentrations. This attempt was not successful since ET were quantitatively found in the IgG elution fraction. (3) Alternatively, it was proposed to adsorb selectively ET in the presence of acetate buffer (pH 5.0) under non binding conditions for human IgG. Human IgG could then be purified if necessary with the same membrane in the presence of MOPS buffer (pH 6.5). With a 1 m2 histidine-PEVA module under these operating conditions, it is estimated that the depyrogenation of 3 1 of 1 mg/ml IgG (human or murine) solution containing 80 EU/ml of ET should be possible.  相似文献   

6.
Iminodiacetic acid (IDA) and tris(2‐aminoethyl)amine (TREN) chelating ligands were immobilized on poly(ethylene vinyl alcohol) (PEVA) hollow‐fiber membranes after activation with epichlorohydrin or butanediol diglycidyl ether (bisoxirane). The affinity membranes complexed with Cu(II) were evaluated for adsorption of human immunoglobulin G (IgG). The effects of matrix activation and buffer system on adsorption of IgG were studied. Isotherms of batch IgG adsorption onto finely cut membranes showed that neither of the chelates, IDA‐Cu(II) or TREN‐Cu(II), had a Langmuirean behavior with negative cooperativity for IgG binding. A comparison of equilibrium and dynamic maximum capacities showed that the dynamic capacity for a mini‐cartridge in a cross‐flow filtration mode (52.5 and 298.4 mg g?1 dry weight for PEVA‐TREN‐Cu(II) and PEVA‐IDA‐Cu(II), respectively) was somewhat higher than the equilibrium capacity (9.2 and 73.3 mg g?1 dry weight for PEVA‐TREN‐Cu(II) and PEVA‐IDA‐Cu(II), respectively). When mini‐cartridges were used, the dynamic adsorption capacity of IDA‐Cu(II) was the same for both mini‐cartridge and agarose gel. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

7.
Human IgG4 antibody shows therapeutically useful properties compared with the IgG1, IgG2, and IgG3 subclasses. Thus IgG4 does not activate complement and shows conformational variability. These properties are attributable to its hinge region, which is the shortest of the four IgG subclasses. Using high throughput scattering methods, we studied the solution structure of wild-type IgG4(Ser222) and a hinge mutant IgG4(Pro222) in different buffers and temperatures where the proline substitution suppresses the formation of half-antibody. Analytical ultracentrifugation showed that both IgG4 forms were principally monomeric with sedimentation coefficients s20,w0 of 6.6–6.8 S. A monomer-dimer equilibrium was observed in heavy water buffer at low temperature. Scattering showed that the x-ray radius of gyration Rg was unchanged with concentration in 50–250 mm NaCl buffers, whereas the neutron Rg values showed a concentration-dependent increase as the temperature decreased in heavy water buffers. The distance distribution curves (P(r)) revealed two peaks, M1 and M2, that shifted below 2 mg/ml to indicate concentration-dependent IgG4 structures in addition to IgG4 dimer formation at high concentration in heavy water. Constrained x-ray and neutron scattering modeling revealed asymmetric solution structures for IgG4(Ser222) with extended hinge structures. The IgG4(Pro222) structure was similar. Both IgG4 structures showed that their Fab regions were positioned close enough to the Fc region to restrict C1q binding. Our new molecular models for IgG4 explain its inability to activate complement and clarify aspects of its stability and function for therapeutic applications.  相似文献   

8.
Because buffers can act as metal ligands, they can effect several reactions necessary for DNA oxidation by ferric iron and thiols, such as iron reduction. Therefore, these reactions were studied in Hepes and phosphate buffers and unbuffered NaCl. Reduction of Fe3+ by dithiothreitol (DTT) and cysteine was observed in either Hepes or NaCl solutions, but not in phosphate buffer. Thiyl radicals were observed in Hepes, but there was much less thiyl radical production in the saline or phosphate solutions. Redox cycling between either DTT or cysteine and Fe3+ also resulted in dioxygen consumption in Hepes buffer. Reduction of Fe3+ and O2 resulted in the formation of an oxidant capable of producing 8-hydroxy-2′-deoxyguanosine (8-OHdG) in calf-thymus DNA. The highest levels of 8-OHdG were detected when DTT or cysteine and Fe3+ were incubated in Hepes, while much less DNA oxidation was detected when the experiment was done in a saline solution, and almost no DNA oxidation occurred in the phosphate buffer. These results demonstrate that the use of different buffers can greatly affect the ability of thiols to promote iron-dependent oxidations. © 1997 John Wiley & Sons, Inc. J Biochem Toxicol 12: 125–132, 1998  相似文献   

9.
The K 562 is a transformed human erythroid stemcell and is used as a target cell for NK-T-cells. In this study the presence of insulin receptors in K 562 is established.The best binding and negative cooperativity was found in the two Hepes containing buffers whereas no cooperativity was obtained in the Krebs-Ringer buffer. The calculated affinity constants and receptor number per cell varied according to the buffer. Preincubation with insulin caused a down-regulation of the insulin binding capacity. 10 ng/ml caused a lowering of the affinity, with an unchanged number of receptors. 100 ng/ml caused a decrease in receptor number with unchanged affinity. These results were found in both Hepes and Krebs-Ringer phosphate buffer. IGF-I shows cross-reactivity with the insulin receptor, with a potency of 12 and 100 times less than insulin in Krebs-Ringer phosphate buffer and G-buffer respectively. However, no specific IGF-I receptors were found.The presence of receptors on K 562 cells suggests a biological role for insulin. The different results in the different buffers, indicate that a buffer containing Hepes and/or Tris, is required to expose negative cooperativity and make the receptors more accessible to insulin.  相似文献   

10.
The adsorbent ω-aminodecyl-agarose was evaluated as to its feasibility for the adsorption of human serum and plasma proteins, aiming at the purification of immunoglobulin G (IgG). The contribution of electrostatic and hydrophobic interactions (mixed-mode) and the effects of buffer system on the adsorption of serum proteins were also studied. The adsorption isotherm parameters of human serum albumin (HSA) and IgG were evaluated, pointing to the existence of cooperative effects in the process. A positive (n = 2.30 ± 0.38) and negative cooperativity (n = 0.63 ± 0.12) were observed for IgG and HSA binding, respectively. High purity IgG was obtained (based on total protein concentration and nephelometric analysis of HSA, transferrin, and immunoglobulins A, G, and M) with a 75% recovery in Hepes 25 mmol L−1 pH 6.8 feeding human serum. These results indicate that the use of ω-aminodecyl-agarose is a potential technique for purification of IgG from human serum.  相似文献   

11.
The applicability of dye-ligands attached to an expanded bed chromatography quartz base matrix (Streamline™) for the affinity bioseparation of rabbit immunoglobulin G (IgG) was investigated. Reactive Green 5 (RG-5) immobilized onto adsorbent was selected for capturing of rabbit-IgG due to its higher binding capacity compared to other dye-ligands possessing similar ligand density. Adsorption parameters such as pH, temperature, ionic strength and initial rabbit-IgG concentration were optimized for the adsorption of rabbit-IgG on the RG-5-immobilized adsorbent. The highest rabbit-IgG adsorption was recorded in pH 7.0, while the maximum binding capacity for BSA was achieved at pH 4.0. The adsorption of rabbit-IgG on RG-5-immobilized adsorbent was declined as the increase of ionic strength. There is no significant influence of temperature against adsorption efficiency of RG-5-immobilized adsorbent for rabbit-IgG. The adsorption phenomenon of rabbit-IgG on RG-5-immobilized adsorbent appeared to follow the Langmuir–Freundlich adsorption isotherm model. The theoretically maximum binding capacity (qm) of RG-5-immobilized adsorbent estimated from this isotherm was 49.3 mg ml−1, which is very close to that obtained experimentally (49.0 mg ml−1). About 50% of bound BSA on RG-5-immobilized adsorbent in binary adsorption system was removed with washing buffer containing 1 M NaCl.  相似文献   

12.
Data are presented regarding the establishment of the pH (designated pH*) of a standard buffer solution suitable as a pH reference in 50 mass% glycerol/water mixtures at temperatures ranging from −20 to 25 °C. The buffer material selected was the ampholyte Mops [(3-N-morpholino)-propane sulfonic acid], and the reference standard consists of equal molal amounts of Mops and its sodium salt. The assignment of pH* values is based on measurements of the electromotive force (emf) of cells without liquid junction of the type: Pt;H2(g,1 atm) ¦ Mops, Na Mopsate, NaCl ¦ AgCl;Ag and the pH* was derived from a determination of K2, the equilibrium constant for the dissociation process (Mops)±/ah (Mopsate) + H +. The standard emf of the silver-silver chloride electrode in 30, 40, and 50 mass% glycerol/water mixtures was determined from emf measurements of the cell at subzero temperatures with HCl solutions replacing the buffer-chloride mixtures.  相似文献   

13.
The interaction of DNA with various neutral pH, amine-based buffers has been analyzed by free solution capillary electrophoresis, using a mixture of a plasmid-sized DNA molecule and a small DNA oligonucleotide as the reporter system. The two DNAs migrate as separate, nearly Gaussian-shaped peaks in 20-80 mM TAE (TAE, Tris-acetate-EDTA; Tris, tris[hydroxymethyl]aminomethane) buffer. The separation between the peaks gradually increases with increasing TAE buffer concentration because of differences in solvent friction between large and small DNA molecules. The two DNAs form complexes with the borate ions in TBE (Tris-borate-EDTA) buffer, with mobilities that depend on the DNA/borate ratio. In 45 mM TBE buffer, the two DNAs comigrate as a single sharp peak, with a mobility that is faster than either of the constituent DNAs in the same buffer. Hence, the mixed DNA-borate complex is stabilized by the binding of additional borate ions, possibly forming bridges between the different DNAs. The mixed DNA-borate complex is gradually dissociated into its component DNAs by increasing the TBE concentration, possibly because the borate binding sites become saturated at high buffer concentrations. Other neutral pH, amine-based buffers, such as Mops (3-[N-morpholino]propanesulfonic acid), Hepes (N-[2-hydroxyethyl]piperazine-N'-[2-ethanesulfonic acid]), Bes (N,N-bis[2-hydroxyethyl]-2-aminoethanesulfonic acid), Tes (N-tris[hydroxymethyl]methyl-2-aminoethanesulfonic acid), and tricine (N-tris[hydroxymethyl]methylglycine) also form complexes with DNA, giving distorted peaks in the electropherograms. The combined results indicate that borate buffers and most neutral pH, amine-based buffers interact with DNA.  相似文献   

14.
We have developed a protocol to quantify polymer DNA cleavage which replaces the traditional radiolabeling and scintillation counting with fluorescent staining and digital imaging. This procedure offers high sensitivity, speed, and convenience, while avoiding waste and error associated with traditional 32P radiolabeling. This protocol was used to measure cleavage of pBR322 plasmid DNA by EcoRV, a type II restriction enzyme. EcoRV was found to exhibit an order of magnitude difference in binding in two apparently similar buffers used in previous investigations. To determine the origin of this effect, we measured reaction kinetics in buffers of different chemical nature and concentration: Tris, bis-Tris propane, Tes, Hepes, and cacodylate. We found that buffer concentration and identity had significant effects on EcoRV reaction velocity through large changes in specific binding and nonspecific binding (reflected in the Michaelis constant Km and the dissociation constant for nonspecific binding Kns). There were only small changes in Vmax. The source of the buffer effect is the protonated amines common to many pH buffers. These buffer cations likely act as counterions screening DNA phosphates, where both the protonated buffer structure and concentration affect enzyme binding strength. It appears that by choosing anionic buffers or zwitterionic buffers with a buried positive charge, buffer influence on the protein binding to DNA can be largely eliminated.  相似文献   

15.
The functional characteristics of hemoglobin (Hb) depend on oxygenation-linked proton and anion binding and thus on solvent buffer groups and ionic composition. This study compares the oxygenation properties of human Hb in ionic [tris(hydroxymethyl)aminomethane (Tris) and BisTris] buffers with those in zwitterionic N-2-hydroxy-ethylpiperazine-N'-2-ethanesulfonic acid (HEPES) buffer under strictly controlled chloride concentrations at different pH values, two temperatures, and in the absence and presence of the erythrocytic cofactor, 2,3-diphosphoglycerate (DPG). In contrast to earlier studies (carried out at the same or different chloride concentrations) it shows only small buffer effects that are manifested at low chloride concentration and high pH. These observations suggest chloride binding to the Tris buffers, which reduces the interaction with specific chloride binding sites in the Hb. The findings indicate that HEPES allows for more accurate assessment of Hb-oxygen affinity and its anion and temperature sensitivities than ionic buffers and advocates standard use of HEPES in studies on Hb function. Precise oxygen affinities of Hb dissolved in both buffers are defined under standard conditions.  相似文献   

16.
An affinity adsorbent for phospholipase A2 (EC 3.1.1.4) was prepared by reacting 10-O-p-toluenesulfonyldecane-1-O-phosphocholine with AH-Sepharose 4B. Phospholipases A2 bind to the immobilized ligand in the presence of Ca2+ and can be eluted with buffers containing EDTA. This principle held not only for soluble phospholipase A2 from porcine pancreas and Crotalus adamanteus, but proved also effective in the purification of phospholipases A2 solubilized from the membranes of rat liver mitochondria and rat platelets.  相似文献   

17.
The membrane-binding properties of a class A amphipathic peptide (18D) were investigated using two different immobilized model membrane systems. The first system involved the use of surface plasmon resonance (SPR) to study the binding of 18D to dimyristylphosphatidylcholine (DMPC) and dimyristylphosphatidylglycerol (DMPG), which allowed peptide binding to be monitored in real time. The SPR experiments indicated stronger binding of 18D to DMPG than DMPC, which kinetic analysis revealed was due to a faster on-rate. The second model membrane system involved immobilized membrane chromatography in which the binding of 18D to either DMPC or DMPG monolayers covalently linked to silica particles was analysed by elution chromatography. Stronger binding affinity of 18D was also obtained with the negatively charged phosphatidylglycerol (PG) monolayer compared to the phosphatidylcholine (PC) monolayer, which was consistent with the SPR results. Non-linear binding behaviour of 18D to the immobilized lipid monolayers was also observed, which suggests that the peptide undergoes conformational and orientational changes upon binding to the immobilized PC and PG ligands. Significant band broadening was also observed on both monolayers, with larger bandwidths obtained on the PC surface, indicating slower binding and orientation kinetics with the zwitterionic surface. The dependence of logk' on the percentage of methanol also demonstrated a bimodal interaction whereby hydrophobic forces predominated at higher temperatures and methanol concentrations, while at lower temperatures, electrostatic and other polar forces also made a contribution to the affinity of the peptides for the lipid monolayer particularly. Overall, these results demonstrate the complementary use of these two lipid biosensors which allows the role of hydrophobic and electrostatic forces in peptide–membrane interactions to be studied and insight gained into the kinetic factors associated with these interactions.  相似文献   

18.
α-Chymotrypsin (α-CT) activity was tested in aqueous media with the following cetyltrialkylammonium bromide surfactants in the series methyl, ethyl, propyl and butyl, different in the head group size, and for the sake of comparison also with the anionic sodium n-dodecyl sulfate and the zwitterionic myristyldimethylammonium propanesulfonate. N-glutaryl-l-phenylalanine p-nitroanilide hydrolysis rate was monitored at surfactant concentration above the critical micellar one. Only some cationic surfactants gave superactivity and the head group size had a major weight. The highest superactivity was measured in the presence of cetyltributylammonium bromide. The effect of both nature and concentration of three different buffers was also investigated. There is a dependence of enzyme superactivity on buffer type. Michaelis–Menten kinetics were found. The binding constants of substrate with micellar aggregates were determined in the used buffers and the effective improvement of reaction rate (at the same free substrate concentration in the medium) was calculated. kcat significantly increased while Km was little changed after correction to free substrate concentration. The ratio of kcat to Km was between 12 and 35 times higher than in pure buffer, depending on buffer and surfactant concentrations. The increase of α-CT activity (30%) was less important in the presence of 1×10−2 M tetrabutylammonium bromide, a very hydrophobic salt, unable to micellise. Fluorescence spectra showed differences of enzyme conformation in the presence of various surfactants.  相似文献   

19.
The conditional stability constant at pH 7.4 for Cu(II) binding at the N-terminal site (NTS) of human serum albumin (HSA) was determined directly by competitive UV–vis spectroscopy titrations using nitrilotriacetic acid (NTA) as the competitor in 100 mM NaCl and 100 mM N-(2-hydroxyethyl)piperazine-N′-ethanesulfonic acid (Hepes). The log K NTSc value of 12.0 ± 0.1 was determined for HSA dissolved in 100 mM NaCl. A false log log K NTSc value of 11.4 ± 0.1 was obtained in the 100 mM Hepes buffer, owing to the formation of a ternary Cu(NTA)(Hepes) complex. The impact of the picomolar affinity of HSA for Cu(II) on the availability of these ions in neurodegenerative disorders is briefly discussed.  相似文献   

20.
The mechanism of increasing effect of CuCl2 on specific [3H]cimetidine binding was examined in brain membranes of rats. CuCl2-Induced elevation of [3H]cimetidine binding was high in Krebs-Ringer solution (pH 7.4) compared to those in 50 mM Na, K-phosphate buffer (pH 7.4) and in 50 mM Tris-HCl buffer (pH 7.4). CaCl2 (5–50 mM) inhibited effect of CuCl2, but NaCl (25–200 mM), KCl (5–100 mM) or MgCl2 (5–50 mM) did not. CuCl2 (50 μM) elevated 9.3- and 2.5-fold the binding in phosphate- and Tris—HCl buffer, respectively. EDTA-2Na decreased the binding elevated by 50 μM CuCl2 in phosphate buffer to the similar level in Tris-HCl buffer, whereas it did not affect those in Tris-HCl buffer. The absorption spectra of cimetidine and CuCl2 mixture showed a peak at 317 nm in phosphate buffer that was not observed in Tris-HCl buffer. It is suggested that cimetidine-Cu2+ chelate complex could be formed in phosphate buffer, resulting in higher amount of binding in phosphate buffer than in Tris-HCl buffer. PdCl2 also caused a marked elevation in [3H]cimetidine binding, seeming to be due to formation of cimetidine-Pd2+ chelate complex. There were two types of [3H]cimetidine binding in the presence of 20 nM PdCl2: high affinity binding with Kd = 0.7 ± 0.1 nM and low affinity binding with Kd = 44.3 ± 3.0 nM. It is suggested that cimetidine-Cu2+ complex binds to cimetidine binding sites in brain with higher affinity than cimetidine alone.  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