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 375 毫秒
1.
Light scattering from fibrin structures, obtained by exposure of fibrinogen to thrombin, Batroxobin (Reptilase) or coagulant fraction extracted from Contortrix venom at 20 and and 37°C, show in every case that rod-like intermediates are formed in the beginning of the aggregation process. The fibrils differ in the extent of branching and in lateral aggregation. Contortrix enzyme causes the highest branching density but the lowest lateral aggregation. Thrombin and Batroxobin give almost identical results. A change of temperature from 20 to 37°C yields an increase in branching density and lateral aggregation for the fibrin structures induced by the two snake venoms. With thrombin, however, the branching density decreases with the elevated temperature while the lateral aggregation strongly increases. Mostly opaque clots are obtained, with the exception of the clots induced by thrombin at 37°C, where a fine or traslucent gel is obtained. A very low extent of branching and translucent gels are also found with thrombin at 20°C and pH 7.3 but at pH 9.5 no correlation between a preferential cleavange of fibrinopeptide B and the lateral aggregation could be detected. The opacity is discussed as being the result of inhomogeneity in both branching and lateral aggregation. A quantitative analysis of the angular dependence of the scattered light indicates that non-activated human fibrinogen exists at least in the two conformations of a long rod, L = 95 ± 5 nm, and a short rod of 47.5 ± 5 nm, with mass fractions of ~ 70 and 30%, respectively. Only the long rod conformation of the monomer is built in the fibril. The model of a pure end-to-end aggregation is shown to be unlikely and the possibility of an overlapping of the monomeric rods over a region of ~ 8 nm is discussed.  相似文献   

2.
The small-angle X-ray scattering (SAXS) method using a synchrotron radiation source was applied to the study of the self-aggregation process of tobacco mosaic virus protein (TMVP) at a concentration of 5.0 or 12.0 mg ml-1 in 50 mM or 100 mM-phosphate buffer (ionic strengths approx. 0.1 and 0.2, respectively) at pH 7.2 in the temperature region of 4.8 to 25.0 degrees C. This paper presents the results of static measurements of SAXS. Sedimentation velocity experiments were performed simultaneously under the same conditions. These results are qualitatively parallel to those of the SAXS measurements, although the size of stacked disks derived from the SAXS measurements is larger than that derived from the sedimentation experiments, suggesting a change in the equilibrium conditions in the centrifugal field. Qualitative analysis of the SAXS data with model simulation calculations implies that the aggregation of TMVP consists of two steps: (1) the aggregation of A-protein comprising a few subunits to form double-layered disks; and (2) the random polymerization of double-layered disks by disk-stacking. Increase in temperature, ionic strength or protein concentration induced TMVP to polymerize to form a double-layered disk or a quadruple-layered short rod with consumption of A-proteins, accompanied by a small number of multi-layered short rods. The SAXS results indicate that the A-protein and the multilayered short rods are polydisperse with respect to size and shape, i.e. the mixture of A-protein, double-layered disks and multi-layered short rods coexists in the equilibrium state without pressure-induced partial dissociation of TMPV as observed during normal ultracentrifugation, and even under solution conditions in which the formation of double-layered disks or higher-order aggregates is favored.  相似文献   

3.
The pH change of 10 different buffering systems with temperature ranging from room to 5 °C was examined; three were conventional buffers which included phosphate yolk, citrate yolk, and skim milk. Seven were Good's buffers with egg yolk which included TES, TRIS, BES, MOPS, PIPES, MES, and TEST. The pH of the three conventional buffers did not change with decreasing temperature, but Good's buffers showed an increase in pH with decreasing temperature from room to 5 °C. The pH change due to temperature was measured for TEST buffer solution with and without 20% egg yolk containing 2 or 6% of five different cryoprotective compounds. The pH at 5 °C was significantly higher than at room temperature. The addition of egg yolk and/or cryoprotective compound did not alter the pH significantly during cooling, even though a slight drop in pH was noted with the addition of egg yolk indicating that the change in pH is primarily due to the buffer. The pH of TEST yolk buffer (pH 7.2 at room temperature) was measured continuously from 37 °C to below freezing (?18 °C). The pH increased with decreasing temperature to 8.0 ± 0.2 from 37 to ?14 °C at which point it dropped abruptly to pH 6.5 ± 0.2.  相似文献   

4.
To determine the stage at which H+ ions are bound during the entropy-driven polymerization of tobacco mosaic virus protein, acid-base titrations were carried out at a concentration of 5 mg/ml in 0.1 m-KCl from pH 8 to pH 5.2 and back to pH 8 at 4, 10, 15 and 20 °C. The titration was always completely reversible when the addition of acid or base was so slow that the experiment required seven hours in each direction. When the titration was started at pH 7 and performed down and up twice as rapidly, a hysteresis loop, indistinguishable from one previously published, was obtained at 20 °C.Ultracentrifugation experiments were carried out at selected pH values at the four temperatures. H+ ion uptake, as determined from the reversible titration curves, is correlated with the disappearance of the 4 S component and is independent of whether the polymerized species is in a 20 S or higher state of aggregation. At pH 7, approximately 1 mole of H+ ion is bound per mole of monomer. At pH values between 6.56 and 6.05, 1.5 moles of H+ ion are bound per mole of monomer upon polymerization. At pH 6.05, 0.5 mole of H+ ion is bound before any polymerization takes place.Tobacco mosaic virus protein at 20 °C in an unbuffered 0.1 m-KCl solution at pH 7.18 at a concentration of 41 mg/ml, largely in the 20 S state, was depolymerized entirely to the 4 S state by dilution with 0.1 m-KCl adjusted to the same pH. Under these conditions, there was no pH change, indicating that no H + ions are released.These seemingly contradictory findings can be explained by assuming that the 4 S component polymerizes to form either double discs without binding H+ ions, or, alternatively, two-turn helices accompanied by the binding of H+ ions. Both double discs and two-turn helices sediment at approximately 20 S. Whether polymerization in the neighborhood of pH 7 leads to helices or discs depends upon the availability of H+ ions.  相似文献   

5.
The Hsp60-type chaperonin GroEL assists in the folding of the enzyme human carbonic anhydrase II (HCA II) and protects it from aggregation. This study was aimed to monitor conformational rearrangement of the substrate protein during the initial GroEL capture (in the absence of ATP) of the thermally unfolded HCA II molten-globule. Single- and double-cysteine mutants were specifically spin-labeled at a topological breakpoint in the β-sheet rich core of HCA II, where the dominating antiparallel β-sheet is broken and β-strands 6 and 7 are parallel. Electron paramagnetic resonance (EPR) was used to monitor the GroEL-induced structural changes in this region of HCA II during thermal denaturation. Both qualitative analysis of the EPR spectra and refined inter-residue distance calculations based on magnetic dipolar interaction show that the spin-labeled positions F147C and K213C are in proximity in the native state of HCA II at 20 °C (as close as ~8 Å), and that this local structure is virtually intact in the thermally induced molten-globule state that binds to GroEL. In the absence of GroEL, the molten globule of HCA II irreversibly aggregates. In contrast, a substantial increase in spin–spin distance (up to >20 Å) was observed within minutes, upon interaction with GroEL (at 50 and 60 °C), which demonstrates a GroEL-induced conformational change in HCA II. The GroEL binding-induced disentanglement of the substrate protein core at the topological break-point is likely a key event for rearrangement of this potent aggregation initiation site, and hence, this conformational change averts HCA II misfolding.  相似文献   

6.
The bone morphogenetic property of bone matrix is degraded at 25 ° to 37 °C within 24 hours after a bone is removed from the body. The degradation occurs in the intact undemineralized bone from the action of endogenous enzymes, presumably neutral proteinases at pH optima of 7 · 0 to 7 · 4. Degradation is: more rapid at physiologic than at acid pH; heat inactivated in the range between 40 ° and 60 °C; slow at 2 °C over a period of 7 days in EDTA at pH 7 · 4. Degradation is inhibited by iodoacetic acid at concentrations as low as 3 · 0 mmoles per liter either in phosphate buffer or EDTA. Degradative activity of endogenous enzymes, as measured by the yield of bone from implants of matrix, is comparable to those obtained from matrix treated with trypsin at 15 °C, pH 7 · 6 over a period of 12 hours. These enzymes include a neutral proteinase (BMP-ase) which degrades bone morphogenetic protein (BMP) without mobilizing bone collagen hydroxy-proline as rapidly and as selectively as a specific functional entity. Observations on carboxypeptidase A and thermolysin cleavage of phenylalanine groups and data on acetylation of tyrosyl groups reducing bone yield suggest aromatic amino acids may be necessary for the biologically active conformation of BMP.  相似文献   

7.
Purified colicin E7 was analyzed by CD spectrum and gel filtration chromatography in a mimicking membrane-translocation phase. It was found that the CD spectra of colicin E7 at pH 7 and pH 2.5 were similar. Although the melting temperature of the protein shifted from 54.5°C to 34°C at low pH, the thermal denaturation curves of colicin E7 at different pH conditions still fit a two-state model. These experimental results imply that a minor structural change, triggered by acidic pH, for instance, may reduce the energy required for protein melting. In contrast to the minor change in secondary structure at different pH conditions, we observed that, in vitro, all monomeric colicin E7s converted into multimer-like conformations after recovering from the partial unfolding process. This multimeric form of colicin can only be dissociated by formamide and guanidine hydrochloride, indicating that this protein complex is indeed formed by aggregation of the monomeric colicins. Most interestingly, the aggregated colicins still perform in vivo bacteriocidal activity. We suggest that in a partial unfolding state the colicin is prepared for binding to the specific targets for translocation through the membrane. However, in the absence of specific targets in vitro these unfold intermediates may therefore aggregate into the multimeric form of colicins. Proteins 32:17–25, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

8.
The effects of temperature (4–20°C), relative humidity (RH, 0–100%), pH (3–7), availability of nutrients (0–5 g/l sucrose) and artificial light (0–494 μmol/m2/s) on macroconidial germination of Fusarium graminearum were studied. Germ tubes emerged between 2 and 6 h after inoculation at 100% RH and 20°C. Incubation in light (205 ± 14 μmol/m/s) retarded the germination for approximately 0.5 h in comparison with incubation in darkness. The times required for 50% of the macroconidia to germinate were 3.5 h at 20°C, 5.4 h at 14°C and 26.3 h at 4°C. No germination was observed after an incubation period of 18 h at 20°C in darkness at RH less than 80%. At RH greater than 80%, germination increased with humidity. Germination was observed when macroconidia were incubated in glucose (5 g/l) or sucrose (concentration range from 2.5 × 10?4 to 5 g/l) whereas no germination was observed when macroconidia were incubated in sterile deionized water up to 22 h. Macroconidia germinated quantitatively within 18 h at pH 3–7. Repeated freezing (?15°C) and thawing (20°C) water agar plates with either germinated or non‐germinated macroconidia for up to five times did not prevent fungal growth after thawing. However, the fungal growth rate of mycelium was negatively related to the number of freezing events the non‐germinated macroconidia experienced. The fungal growth rate of mycelium was not significantly affected by the number of freezing events the germinated spores experienced. Incubation of macroconidia at low humidity (0–53% RH) suppressed germination and decreased the viability of the spores.  相似文献   

9.
In the summer of 2008, the abundance, local aggregation structure, growth rate, and life span of the gastropod Nucella freycineti inhabiting the littoral zone of Yankicha Island (Kuriles), which is affected by the postvolcanic activity of the Ushishir Volcano were studied. The parameters varied within a wide range and appreciably depended on the distance mollusk populations were from underwater and land gas-hydrothermal vents, which caused strong heating (up to 48°C) of cold sea waters and the formation of an unusually acidic environment (pH approximately 3.5) in the Kraternaya Bay. The mollusks occurred at a water temperature below 24°C and a pH above 4, but formed multiage populations at a temperature of 2–14°C and a pH of approximately 6.4–8.2.  相似文献   

10.
Freezing of biologic drug substance at large scale is an important unit operation that enables manufacturing flexibility and increased use‐period for the material. Stability of the biologic in frozen solutions is associated with a number of issues including potentially destabilizing pH changes. The pH changes arise from temperature‐associated change in the pKas, solubility limitations, eutectic crystallization, and cryoconcentration. The pH changes for most of the common protein formulation buffers in the frozen state have not been systematically measured. Sodium phosphate buffer, a well‐studied system, shows the greatest change in pH when going from +25 to ?30°C. Among the other buffers, histidine hydrochloride, sodium acetate, histidine acetate, citrate, and succinate, less than 1 pH unit change (increase) was observed over the temperature range from +25 to ?30°C, whereas Tris‐hydrochloride had an ~1.2 pH unit increase. In general, a steady increase in pH was observed for all these buffers once cooled below 0°C. A formulated IgG2 monoclonal antibody in histidine buffer with added trehalose showed the same pH behavior as the buffer itself. This antibody in various formulations was subject to freeze/thaw cycling representing a wide process (phase transition) time range, reflective of practical situations. Measurement of soluble aggregates after repeated freeze–thaw cycles shows that the change in pH was not a factor for aggregate formation in this case, which instead is governed by the presence or absence of noncrystallizing cryoprotective excipients. In the absence of a cryoprotectant, longer phase transition times lead to higher aggregation. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2010  相似文献   

11.
Small heat shock proteins (sHsps) are a ubiquitous family of molecular chaperones that rescue misfolded proteins from irreversible aggregation during cellular stress. Many such sHsps exist as large polydisperse species in solution, and a rapid dynamic subunit exchange between oligomeric and dissociated forms modulates their function under a variety of stress conditions. Here, we investigated the structural and functional properties of Hsp20 from thermoacidophilic crenarchaeon Sulfolobus acidocaldarius. To provide a framework for investigating the structure-function relationship of Hsp20 and understanding its dynamic nature, we employed several biophysical and biochemical techniques. Our data suggested the existence of a ~24-mer of Hsp20 at room temperature (25 °C) and a higher oligomeric form at higher temperature (50 °C–70 °C) and lower pH (3.0–5.0). To our surprise, we identified a dimeric form of protein as the functional conformation in the presence of aggregating substrate proteins. The hydrophobic microenvironment mainly regulates the oligomeric plasticity of Hsp20, and it plays a key role in the protection of stress-induced protein aggregation. In Sulfolobus sp., Hsp20, despite being a non-secreted protein, has been reported to be present in secretory vesicles and it is still unclear whether it stabilizes substrate proteins or membrane lipids within the secreted vesicles. To address such an issue, we tested the ability of Hsp20 to interact with membrane lipids along with its ability to modulate membrane fluidity. Our data revealed that Hsp20 interacts with membrane lipids via a hydrophobic interaction and it lowers the propensity of in vitro phase transition of bacterial and archaeal lipids.  相似文献   

12.
氯嘧磺隆降解菌株LW-3的分离及生物学特性研究   总被引:3,自引:0,他引:3  
从长期受氯嘧磺隆污染的土壤中分离到一株氯嘧磺隆高效降解菌株,命名为LW-3,菌株LW-3可以氯嘧磺隆为唯一氮源生长,接种量为2%时,50 mg/L氯嘧磺隆经过7 d,降解率达70%~80%.生理生化实验和16S rDNA序列同源性分析,将菌株LW-3归属于假单胞茵属(Pseudomonas sp.);菌株的适宜降解条件为;温度30℃~35℃,pH 6.5~7.2,pH对菌株LW-3降解氯嘧磺隆有明显地影响.  相似文献   

13.
Xylem sap was collected from wheat and barley growing in a drying soil, and the effect of the sap on transpiration was detected by a bioassay with detached wheat leaves. The inhibitory activity of fresh sap was small, and could be largely accounted for by the abscisic acid content (about 2×10-5mol m-3). When fresh sap was stored at -20°C for several days, the activity increased. Maximum activity developed after a week. This increase in activity was due to a compound that increased in size with storage at -20°C. When fresh sap was fractionated with filters of different molecular size exclusion characteristics, and the separated fractions stored at -20°C for a week, activity developed only in the fraction containing compounds smaller than 0·3 kDa. However, when sap already stored at -20°C was fractionated, activity was only in fractions containing compounds larger than 0·3 kDa. The increase in activity and in size did not occur with storage in liquid nitrogen (-196°C) or at -80°C. These results suggest that storage at -20°C causes the aggregation or polymerization of a small compound with low activity to form a large compound with high activity. This change is not catalysed by an enzyme because it can occur in a fraction from which molecules larger than 0·3 kDa are removed. It is probably promoted by high solute concentrations when ice crystals form. Sap collected from plants in soils of high water potential had little or no activity after storage at -20°C.  相似文献   

14.
Proton binding to tobacco mosaic virus protein at 20 °C has been found to exhibit a reproducible hysteresis which results from the metastability of high molecular weight helical, virus-like rods. In a titration from pH 4 or 5 to 7, the time for depolymerization of such rods, as measured by ultracentrifugation, decreases from days to minutes over a range of about a tenth of a pH unit, near pH 6·6 at 20 °C. Relative to the extent of proton binding in the depolymerized state at 4 °C, the magnitude of the hysteresis near pH 6·2 corresponds to more than 50% of the protons bound per subunit in the equilibrium polymerized state.  相似文献   

15.
Lower induction temperature and polyoxyethylene sorbitan monolaurate (Tween-20) were successfully used to inhibit the aggregation of recombinant human consensus interferon-α mutant (cIFN) during Pichia pastoris fermentation. When the induction temperature was decreased from 30 to 20°C, the cIFN secreted into the medium was in the form of monomers instead of aggregates. The maximum specific activity at 20°C was 4.04 times as high as that at 30°C. There was no obvious effect on the cell growth at 20°C, but the total protein level was decreased. Similar inhibition effect on cIFN aggregation was observed when 0.2 g l−1 Tween-20 was added during induction. Furthermore, there was a synergistic effect found between induction temperature and Tween-20 on the inhibition of cIFN aggregation. The maximum specific activity with Tween-20 at 20°C was 19.9-fold higher than that without Tween-20 at 30°C.  相似文献   

16.
Concentration dependent and temperature dependent stopped-flow experiments on the transition A-protein→double disc show a triphasic reaction consisting of a nucleation phase, a propagation phase and a slow redistribution of polymer size which involves the dissociation of “overshoot” aggregates into double discs and smaller aggregates. No first order rate process is observed under the present experimental conditions. Equilibrium circular dichroism data and preliminary kinetic data at various temperatures indicate a change at about 21°C which might be correlated to a partial transition double disc→helix parallelled by a further shift in the equilibrium of double disc formation; from both data the thermodynamic and activation parameters for the A-protein→double disc transition are estimated.  相似文献   

17.
We searched for a new cell aggregation factor, and found one we called SAF in the mycelia of a strain of Streptomyces murinus. SAF was purified by active carbon and ion exchange column chromatographies and gel filtration of Sepharose 2B from the homogenized mycelia by sonication. SAF was stable from pH 7 to 9 at 37°C and up to 40°C at pH 8.0. The aggregation activity of SAF was maximum around pH 8.0 at 30°C, and the factor required for its activity metallic ions such as calcium and manganese. The aggregation activity of SAF was inhibited by laminarin, but it was not influenced by various other saccharides. SAF aggregated E. coli, S. aureus, M. luteus, sarcoma 180, and HeLa cells as well as S. marcescens, above all, its highest activity was toward B. subtilis, but P. vulgaris, P. aeruginosa, C. albicans, each type of human erythrocytes, and hepatoma 109A cells were quite resistant to SAF. These properties has proved that SAF is completely different from the other aggregation factors so far reported.  相似文献   

18.
Tannin acyl hydrolase (Tannase) from Asp. oryzae No. 7 was purified. The purified enzyme was homogenous on column chromatography (DEAE-Sephadex A50, Sephadex G100), ultra centrifugation and electrophoresis.

The molecular weight of the enzyme estimated by gel filtration method was about 200,000.

The enzyme was stable in the range of pH 3 to 7.5 for 12 hr at 5°C, and for 25 hr at the same temperature in the range of pH 4.5 to 6. The optimum pH for the reaction was 5.5. It was stable under 30°C (over one day, in 0.05 M-citrate buffer of pH 5.5), and the optimum temperature was 30~40°C (reaction for 20min). The activity was lost completely at 55°C in 20 min at pH 5.5, or at 85°C in 10 min at the same pH.

Any metal salt tested did not activate the enzyme, Zink chloride and cupric chloride inhibited the activity or denatured the enzyme. The activity was lost completely by dialysis against EDTA-solution at pH 7.25, although it was not affected by dialysis against deionized water.  相似文献   

19.
In order to select plant material for obtaining a high proportion of chlorophyll-free protein from Helianthus species by heat fractionation, leaf extracts from 11 field grown cultivars of H. annuus and H. debilis were investigated. In addition, press juices from H. annuus cv. Kinesisk were supplied with salts or urea before heat treatment in order to increase the proportion of chlorophyll-free protein during heat fractionation. The extracts were adjusted to pH 5.0, 6.0 and 8.0; then left for 20 min at 20°C, 50°C or 60°C and centrifuged at 2,500 g. The highest percentage chlorophyll-free protein of the total protein in the extract, more than 10%, was obtained for H. debilis cv. fl.pl. Sun Gold, H. annuus cv. Giganta, H. annuus cv. uniflorus and H. debilis cv. Stella, at pH 6.0 and 50°C. The low percentage chlorophyll-free protein obtained could be explained by the fact that a considerable part of the chlorophyll-free protein sedimented at original pH, low temperature and low centrifugation speed. This sedimentation at low temperature was probably due to interactions of phenolics. Besides, if the chlorophyll-associated membranes were highly fragmented during the disintegration of the plant material, the high temperature required to precipitate them completely led to further aggregation and thus to loss of chlorophyll-free proteins. Salts or urea present during heat treatment (pH 6.0/55°C for 20 min) did not considerably increase the proportion of chlorophyll-free proteins obtained.  相似文献   

20.
Egg maturation in Calliphora vicina is known to occur within a wide range of temperatures, from 12°C to nearly 30°C (Vinogradova, 1991). Photoperiodism has no effect on this process. Some females enter diapause already at 20°C; their fraction increases at lower temperatures and reaches 100% at 6°C. Reproducing females with eggs can survive for a long time and even lay eggs at low temperatures (4–5°C). Experiments with C. vicina from Leningrad Province revealed some effects of the diet (liver or fish) and temperature on the fly reproduction. At 20 and 25°C, 7–10-day old females begin to oviposit, but at 20°C egg maturation is observed in 98% of females feeding on liver and in only 5% of females feeding on fish. On the liver diet, the mean daily fecundity is significantly correlated with the day of oviposition but not with the temperature. At 20°C a significant correlation is observed between the mean daily fecundity and both the day of oviposition and food. The total number of eggs laid by flies after feeding on fish is half that produced after feeding on liver. The optimal conditions for Calliphora vicina cultivation are a 16-h light day, temperatures within the range from 20 to 25°C, and liver as food.  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